fbpx
Wikipedia

Haber process

The Haber process,[1] also called the Haber–Bosch process, is the main industrial procedure for the production of ammonia.[2][3] It is named after its inventors, the German chemists Fritz Haber and Carl Bosch, who developed it in the first decade of the 20th century. The process converts atmospheric nitrogen (N2) to ammonia (NH3) by a reaction with hydrogen (H2) using a metal catalyst under high temperatures and pressures. This reaction is slightly exothermic (i.e. it releases energy), meaning that the reaction is favoured at lower temperatures[4] and higher pressures.[5] It decreases entropy, complicating the process. Hydrogen is produced via steam reforming, followed by an iterative closed cycle to react hydrogen with nitrogen to produce ammonia.

The primary reaction is:

Before the development of the Haber process, it had been difficult to produce ammonia on an industrial scale,[6][7][8] because earlier methods, such as the Birkeland–Eyde process and the Frank–Caro process, were too inefficient.

History

 

During the 19th century, the demand for nitrates and ammonia for use as fertilizers and industrial feedstocks rapidly increased. The main source was mining niter deposits and guano from tropical islands.[9] At the beginning of the 20th century these reserves were thought insufficient to satisfy future demands,[10] and research into new potential sources of ammonia increased. Although atmospheric nitrogen (N2) is abundant, comprising ~78% of the air, it is exceptionally stable and does not readily react with other chemicals.

Haber, with his assistant Robert Le Rossignol, developed the high-pressure devices and catalysts needed to demonstrate the Haber process at a laboratory scale.[11][12] They demonstrated their process in the summer of 1909 by producing ammonia from the air, drop by drop, at the rate of about 125 mL (4 US fl oz) per hour. The process was purchased by the German chemical company BASF, which assigned Carl Bosch the task of scaling up Haber's tabletop machine to industrial scale.[7][13] He succeeded in 1910. Haber and Bosch were later awarded Nobel Prizes, in 1918 and 1931 respectively, for their work in overcoming the chemical and engineering problems of large-scale, continuous-flow, high-pressure technology.[7]

Ammonia was first manufactured using the Haber process on an industrial scale in 1913 in BASF's Oppau plant in Germany, reaching 20 tonnes/day in 1914.[14] During World War I, the production of munitions required large amounts of nitrate. The Allies had access to large deposits of sodium nitrate in Chile (Chile saltpetre) controlled by British companies. Germany had no such resources, so the Haber process proved essential to the German war effort.[7][15] Synthetic ammonia from the Haber process was used for the production of nitric acid, a precursor to the nitrates used in explosives.

The original Haber–Bosch reaction chambers used osmium as the catalyst, but it was available in extremely small quantities. Haber noted uranium was almost as effective and easier to obtain than osmium. In 1909, BASF researcher Alwin Mittasch discovered a much less expensive iron-based catalyst that is still used. A major contributor to the elucidation of this catalysis was Gerhard Ertl.[16][17][18][19] The most popular catalysts are based on iron promoted with K2O, CaO, SiO2, and Al2O3.

During the interwar years, alternative processes were developed, most notably the Casale process, Claude process, and the Mont-Cenis process developed by Friedrich Uhde Ingenieurbüro.[20] Luigi Casale and Georges Claude proposed to increase the pressure of the synthesis loop to 80–100 MPa (800–1,000 bar; 12,000–15,000 psi), thereby increasing the single-pass ammonia conversion and making nearly complete liquefaction at ambient temperature feasible. Claude proposed to have three or four converters with liquefaction steps in series, thereby avoiding recycling. Most plants continue to use the original Haber process (20 MPa (200 bar; 2,900 psi) and 500 °C (932 °F)), albeit with improved single-pass conversion and lower energy consumption due to process and catalyst optimization.

Process

 
A historical (1921) high-pressure steel reactor for the production of ammonia via the Haber process is displayed at the Karlsruhe Institute of Technology, Germany

Combined with the energy needed to produce hydrogen[note 1] and purified atmospheric nitrogen, ammonia production is energy-intensive, accounting for 1% to 2% of global energy consumption, 3% of global carbon emissions,[22] and 3% to 5% of natural gas consumption.[23]

The choice of catalyst is important for synthesizing ammonia. In 2012, Hideo Hosono's group found that Ru-loaded calcium-aluminum oxide C12A7:e electride works well as a catalyst and pursued more efficient formation.[24][25] This method is implemented in a small plant for ammonia synthesis in Japan.[26][27] In 2019, Hosono's group found another catalyst, a novel perovskite oxynitride-hydride BaCeO3−xNyHz, that works at lower temperature and without costly ruthenium.[28]

Hydrogen production

The major source of hydrogen is methane. Steam reforming extracts hydrogen from methane in a high-temperature and pressure tube inside a reformer with a nickel catalyst. Other fossil fuel sources include coal, heavy fuel oil and naphtha.

Green hydrogen is produced without fossil fuels or carbon dioxide emissions from biomass, water electrolysis and thermochemical (solar or another heat source) water splitting. However, these hydrogen sources are not economically competitive with steam reforming.[29][30][31]

Starting with a natural gas (CH
4
) feedstock, the steps are:

  • Remove sulfur compounds from the feedstock, because sulfur deactivates the catalysts used in subsequent steps. Sulfur removal requires catalytic hydrogenation to convert sulfur compounds in the feedstocks to gaseous hydrogen sulfide:
H2 + RSH → RH + H2S(gas)
  • Hydrogen sulfide is adsorbed and removed by passing it through beds of zinc oxide where it is converted to solid zinc sulfide:
 
Illustrating inputs and outputs of steam reforming of natural gas, a process to produce hydrogen
H2S + ZnO → ZnS + H2O
CH4 + H2O → CO + 3 H2
CO + H2O → CO2 + H2
  • Carbon dioxide is removed either by absorption in aqueous ethanolamine solutions or by adsorption in pressure swing adsorbers (PSA) using proprietary solid adsorption media.
  • The final step in producing hydrogen is to use catalytic methanation to remove residual carbon monoxide or carbon dioxide:
CO  + 3 H2 → CH4 + H2O
CO2 + 4 H2 → CH4 + 2 H2O

Ammonia production

The hydrogen is catalytically reacted with nitrogen (derived from process air) to form anhydrous liquid ammonia. It is difficult and expensive, as lower temperatures result in slower reaction kinetics (hence a slower reaction rate)[32] and high pressure requires high-strength pressure vessels[33] that resist hydrogen embrittlement. Diatomic nitrogen is bound together by a triple bond, which makes it relatively inert.[34] Yield and efficiency are low, meaning that the ammonia must be extracted and the gases reprocessed for the reaction to proceed at an acceptable pace.[35]

This step is known as the ammonia synthesis loop:

3 H2 + N2 → 2 NH3

The gases (nitrogen and hydrogen) are passed over four beds of catalyst, with cooling between each pass to maintain a reasonable equilibrium constant. On each pass, only about 15% conversion occurs, but unreacted gases are recycled, and eventually conversion of 97% is achieved.[3]

Due to the nature of the (typically multi-promoted magnetite) catalyst used in the ammonia synthesis reaction, only low levels of oxygen-containing (especially CO, CO2 and H2O) compounds can be tolerated in the hydrogen/nitrogen mixture. Relatively pure nitrogen can be obtained by air separation, but additional oxygen removal may be required.

Because of relatively low single pass conversion rates (typically less than 20%), a large recycle stream is required. This can lead to the accumulation of inerts in the gas.

Nitrogen gas (N2) is unreactive because the atoms are held together by triple bonds. The Haber process relies on catalysts that accelerate the scission of these bonds.

Two opposing considerations are relevant: the equilibrium position and the reaction rate. At room temperature, the equilibrium is in favor of ammonia, but the reaction doesn't proceed at a detectable rate due to its high activation energy. Because the reaction is exothermic, the equilibrium constant becomes unity at around 150–200 °C (302–392 °F) following Le Châtelier's principle.[3]

K(T) for N
2
+ 3 H
2
⇌ 2 NH
3
[36]
Temperature (°C) Kp
300 4.34 × 10−3
400 1.64 × 10−4
450 4.51 × 10−5
500 1.45 × 10−5
550 5.38 × 10−6
600 2.25 × 10−6

Above this temperature, the equilibrium quickly becomes unfavorable at atmospheric pressure, according to the Van 't Hoff equation. Lowering the temperature is unhelpful because the catalyst requires a temperature of at least 400 °C to be efficient.[3]

Increased pressure favors the forward reaction because 4 moles of reactant produce 2 moles of product, and the pressure used (15–25 MPa (150–250 bar; 2,200–3,600 psi)) alters the equilibrium concentrations to give a substantial ammonia yield. The reason for this is evident in the equilibrium relationship:

 

where   is the fugacity coefficient of species  ,   is the mole fraction of the same species,   is the reactor pressure, and   is standard pressure, typically 1 bar (0.10 MPa).

Economically, reactor pressurization is expensive: pipes, valves, and reaction vessels need to be strong enough, and safety considerations affect operating at 20 MPa. Compressors take considerable energy, as work must be done on the (compressible) gas. Thus, the compromise used gives a single-pass yield of around 15%.[3]

While removing the ammonia from the system increases the reaction yield, this step is not used in practice, since the temperature is too high; instead it is removed from the gases leaving the reaction vessel. The hot gases are cooled under high pressure, allowing the ammonia to condense and be removed as a liquid. Unreacted hydrogen and nitrogen gases are returned to the reaction vessel for another round.[3] While most ammonia is removed (typically down to 2–5 mol.%), some ammonia remains in the recycle stream. In academic literature, a more complete separation of ammonia has been proposed by absorption in metal halides or zeolites. Such a process is called an absorbent-enhanced Haber process or adsorbent-enhanced Haber–Bosch process.[37]

Pressure/temperature

The steam reforming, shift conversion, carbon dioxide removal, and methanation steps each operate at absolute pressures of about 25 to 35 bar, while the ammonia synthesis loop operates at temperatures of 300–500 °C (572–932 °F) and pressures ranging from 60 to 180 bar depending upon the method used. The resulting ammonia must then be separated from the residual hydrogen and nitrogen at temperatures of −20 °C (−4 °F).[38][3]

Catalysts

 
First reactor at the Oppau plant in 1913
 
Profiles of the active components of heterogeneous catalysts; the top right figure shows the profile of a shell catalyst.

The Haber–Bosch process relies on catalysts to accelerate N2 hydrogenation. The catalysts are heterogeneous, solids that interact with gaseous reagents.[39]

The catalyst typically consists of finely divided iron bound to an iron oxide carrier containing promoters possibly including aluminium oxide, potassium oxide, calcium oxide, potassium hydroxide,[40] molybdenum,[41] and magnesium oxide.

Iron-based catalysts

The iron catalyst is obtained from finely ground iron powder, which is usually obtained by reduction of high-purity magnetite (Fe3O4). The pulverized iron is oxidized to give magnetite or wüstite (FeO, ferrous oxide) particles of a specific size. The magnetite (or wüstite) particles are then partially reduced, removing some of the oxygen. The resulting catalyst particles consist of a core of magnetite, encased in a shell of wüstite, which in turn is surrounded by an outer shell of metallic iron. The catalyst maintains most of its bulk volume during the reduction, resulting in a highly porous high-surface-area material, which enhances its catalytic effectiveness. Minor components include calcium and aluminium oxides, which support the iron catalyst and help it maintain its surface area. These oxides of Ca, Al, K, and Si are unreactive to reduction by hydrogen.[3]

The production of the catalyst requires a particular melting process in which used raw materials must be free of catalyst poisons and the promoter aggregates must be evenly distributed in the magnetite melt. Rapid cooling of the magnetite, which has an initial temperature of about 3500 °C, produces the desired precursor. Unfortunately, the rapid cooling ultimately forms a catalyst of reduced abrasion resistance. Despite this disadvantage, the method of rapid cooling is often employed.[3]

The reduction of the precursor magnetite to α-iron is carried out directly in the production plant with synthesis gas. The reduction of the magnetite proceeds via the formation of wüstite (FeO) so that particles with a core of magnetite become surrounded by a shell of wüstite. The further reduction of magnetite and wüstite leads to the formation of α-iron, which forms together with the promoters of the outer shell.[42] The involved processes are complex and depend on the reduction temperature: At lower temperatures, wüstite disproportionates into an iron phase and a magnetite phase; at higher temperatures, the reduction of the wüstite and magnetite to iron dominates.[43]

The α-iron forms primary crystallites with a diameter of about 30 nanometers. These crystallites form a bimodal pore system with pore diameters of about 10 nanometers (produced by the reduction of the magnetite phase) and of 25 to 50 nanometers (produced by the reduction of the wüstite phase).[42] With the exception of cobalt oxide, the promoters are not reduced.

During the reduction of the iron oxide with synthesis gas, water vapor is formed. This water vapor must be considered for high catalyst quality as contact with the finely divided iron would lead to premature aging of the catalyst through recrystallization, especially in conjunction with high temperatures. The vapor pressure of the water in the gas mixture produced during catalyst formation is thus kept as low as possible, target values are below 3 gm−3. For this reason, the reduction is carried out at high gas exchange, low pressure, and low temperatures. The exothermic nature of the ammonia formation ensures a gradual increase in temperature.[3]

The reduction of fresh, fully oxidized catalyst or precursor to full production capacity takes four to ten days.[3] The wüstite phase is reduced faster and at lower temperatures than the magnetite phase (Fe3O4). After detailed kinetic, microscopic, and X-ray spectroscopic investigations it was shown that wüstite reacts first to metallic iron. This leads to a gradient of iron(II) ions, whereby these diffuse from the magnetite through the wüstite to the particle surface and precipitate there as iron nuclei.

Pre-reduced, stabilized catalysts occupy a significant market share. They are delivered showing the fully developed pore structure, but have been oxidized again on the surface after manufacture and are therefore no longer pyrophoric. The reactivation of such pre-reduced catalysts requires only 30 to 40 hours instead of several days. In addition to the short start-up time, they have other advantages such as higher water resistance and lower weight.[3]

Typical catalyst composition[44] Iron (%) Potassium (%) Aluminium (%) Calcium (%) Oxygen (%)
Volume composition 40.5 00.35 02.0 1.7 53.2
Surface composition before reduction 08.6 36.1 10.7 4.7 40.0
Surface composition after reduction 11.0 27.0 17.0 4.0 41.0

Catalysts other than iron

Many efforts have been made to improve the Haber–Bosch process. Many metals were tested as catalysts. The requirement for suitability is the dissociative adsorption of nitrogen (i. e. the nitrogen molecule must be split into nitrogen atoms upon absorption). If the binding of the nitrogen is too strong, the catalyst is blocked and the catalytic ability is reduced (self-poisoning). The elements in the periodic table to the left of the iron group show such strong bonds. Further, the formation of surface nitrides makes, for example, chromium catalysts ineffective. Metals to the right of the iron group, in contrast, adsorb nitrogen too weakly for ammonia synthesis. Haber initially used catalysts based on osmium and uranium. Uranium reacts to its nitride during catalysis, while osmium oxide is rare.[45]

According to theoretical and practical studies, improvements over pure iron are limited. The activity of iron catalysts is increased by the inclusion of cobalt.[46]

Ruthenium

Ruthenium forms highly active catalysts. Allowing milder operating pressures and temperatures, Ru-based materials are referred to as second-generation catalysts. Such catalysts are prepared by the decomposition of triruthenium dodecacarbonyl on graphite.[3] A drawback of activated-carbon-supported ruthenium-based catalysts is the methanation of the support in the presence of hydrogen. Their activity is strongly dependent on the catalyst carrier and the promoters. A wide range of substances can be used as carriers, including carbon, magnesium oxide, aluminium oxide, zeolites, spinels, and boron nitride.[47]

Ruthenium-activated carbon-based catalysts have been used industrially in the KBR Advanced Ammonia Process (KAAP) since 1992.[48] The carbon carrier is partially degraded to methane; however, this can be mitigated by a special treatment of the carbon at 1500 °C, thus prolonging the catalyst lifetime. In addition, the finely dispersed carbon poses a risk of explosion. For these reasons and due to its low acidity, magnesium oxide has proven to be a good choice of carrier. Carriers with acidic properties extract electrons from ruthenium, make it less reactive, and have the undesirable effect of binding ammonia to the surface.[47]

Catalyst poisons

Catalyst poisons lower catalyst activity. They are usually impurities in the synthesis gas. Permanent poisons cause irreversible loss of catalytic activity and, while temporary poisons lower the activity while present. Sulfur compounds, phosphorus compounds, arsenic compounds, and chlorine compounds are permanent poisons. Oxygenic compounds like water, carbon monoxide, carbon dioxide, and oxygen are temporary poisons.[3][49]

Although chemically inert components of the synthesis gas mixture such as noble gases or methane are not strictly poisons, they accumulate through the recycling of the process gases and thus lower the partial pressure of the reactants, which in turn slows conversion.[50]

Industrial production

Synthesis parameters

Change of the equilibrium constant Keq as a function of temperature[51]
temperature (°C) Keq
300 4.34 × 10−3
400 1.64 × 10−4
450 4.51 × 10−5
500 1.45 × 10−5
550 5.38 × 10−6
600 2.25 × 10−6

The reaction is:

 [52]

The reaction is an exothermic equilibrium reaction in which the gas volume is reduced. The equilibrium constant Keq of the reaction (see table) is obtained from:

 

Since the reaction is exothermic, the equilibrium of the reaction shifts at lower temperatures to the ammonia side. Furthermore, four volumetric units of the raw materials produce two volumetric units of ammonia. According to Le Chatelier's principle, higher pressure favours ammonia. High pressure is necessary to ensure sufficient surface coverage of the catalyst with nitrogen.[53] For this reason, a ratio of nitrogen to hydrogen of 1 to 3, a pressure of 250 to 350 bar, a temperature of 450 to 550 °C and α iron are optimal.

The catalyst ferrite (α-Fe) is produced in the reactor by the reduction of magnetite with hydrogen. The catalyst has its highest efficiency at temperatures of about 400 to 500 °C. Even though the catalyst greatly lowers the activation energy for the cleavage of the triple bond of the nitrogen molecule, high temperatures are still required for an appropriate reaction rate. At the industrially used reaction temperature of 450 to 550 °C an optimum between the decomposition of ammonia into the starting materials and the effectiveness of the catalyst is achieved.[54] The formed ammonia is continuously removed from the system. The volume fraction of ammonia in the gas mixture is about 20%.

The inert components, especially the noble gases such as argon, should not exceed a certain content in order not to reduce the partial pressure of the reactants too much. To remove the inert gas components, part of the gas is removed and the argon is separated in a gas separation plant. The extraction of pure argon from the circulating gas is carried out using the Linde process.[55]

Large-scale implementation

Modern ammonia plants produce more than 3000 tons per day in one production line. The following diagram shows the set-up of a Haber–Bosch plant:

 
 primary reformer  air feed  secondary reformer  CO conversion  washing tower  ammonia reactor  heat exchanger  ammonia condenser

Depending on its origin, the synthesis gas must first be freed from impurities such as hydrogen sulfide or organic sulphur compounds, which act as a catalyst poison. High concentrations of hydrogen sulfide, which occur in synthesis gas from carbonization coke, are removed in a wet cleaning stage such as the sulfosolvan process, while low concentrations are removed by adsorption on activated carbon.[56] Organosulfur compounds are separated by pressure swing adsorption together with carbon dioxide after CO conversion.

To produce hydrogen by steam reforming, methane reacts with water vapor using a nickel oxide-alumina catalyst in the primary reformer to form carbon monoxide and hydrogen. The energy required for this, the enthalpy ΔH, is 206 kJ/mol.[57]

 

The methane gas reacts in the primary reformer only partially. To increase the hydrogen yield and keep the content of inert components (i. e. methane) as low as possible, the remaining methane gas is converted in a second step with oxygen to hydrogen and carbon monoxide in the secondary reformer. The secondary reformer is supplied with air as the oxygen source. Also, the required nitrogen for the subsequent ammonia synthesis is added to the gas mixture.

 

In the third step, the carbon monoxide is oxidized to carbon dioxide, which is called CO conversion or water-gas shift reaction.

 

Carbon monoxide and carbon dioxide would form carbamates with ammonia, which would clog (as solids) pipelines and apparatus within a short time. In the following process step, the carbon dioxide must therefore be removed from the gas mixture. In contrast to carbon monoxide, carbon dioxide can easily be removed from the gas mixture by gas scrubbing with triethanolamine. The gas mixture then still contains methane and noble gases such as argon, which, however, behave inertly.[50]

The gas mixture is then compressed to operating pressure by turbo compressors. The resulting compression heat is dissipated by heat exchangers; it is used to preheat raw gases.

The actual production of ammonia takes place in the ammonia reactor. The first reactors were bursting under high pressure because the atomic hydrogen in the carbonaceous steel partially recombined into methane and produced cracks in the steel. Bosch, therefore, developed tube reactors consisting of a pressure-bearing steel tube in which a low-carbon iron lining tube was inserted and filled with the catalyst. Hydrogen that diffused through the inner steel pipe escaped to the outside via thin holes in the outer steel jacket, the so-called Bosch holes.[52] A disadvantage of the tubular reactors was the relatively high-pressure loss, which had to be applied again by compression. The development of hydrogen-resistant chromium-molybdenum steels made it possible to construct single-walled pipes.[58]

 
Modern ammonia reactor with heat exchanger modules: The cold gas mixture is preheated to reaction temperature in heat exchangers by the reaction heat and cools in turn the produced ammonia.

Modern ammonia reactors are designed as multi-storey reactors with a low-pressure drop, in which the catalysts are distributed as fills over about ten storeys one above the other. The gas mixture flows through them one after the other from top to bottom. Cold gas is injected from the side for cooling. A disadvantage of this reactor type is the incomplete conversion of the cold gas mixture in the last catalyst bed.[58]

Alternatively, the reaction mixture between the catalyst layers is cooled using heat exchangers, whereby the hydrogen-nitrogen mixture is preheated to the reaction temperature. Reactors of this type have three catalyst beds. In addition to good temperature control, this reactor type has the advantage of better conversion of the raw material gases compared to reactors with cold gas injection.

Uhde has developed and is using an ammonia converter with three radial flow catalyst beds and two internal heat exchangers instead of axial flow catalyst beds. This further reduces the pressure drop in the converter.[59]

The reaction product is continuously removed for maximum yield. The gas mixture is cooled to 450 °C in a heat exchanger using water, freshly supplied gases, and other process streams. The ammonia also condenses and is separated in a pressure separator. Unreacted nitrogen and hydrogen are then compressed back to the process by a circulating gas compressor, supplemented with fresh gas, and fed to the reactor.[58] In a subsequent distillation, the product ammonia is purified.

Mechanism

Elementary steps

The mechanism of ammonia synthesis contains the following seven elementary steps:

  1. transport of the reactants from the gas phase through the boundary layer to the surface of the catalyst.
  2. pore diffusion to the reaction center
  3. adsorption of reactants
  4. reaction
  5. desorption of product
  6. transport of the product through the pore system back to the surface
  7. transport of the product into the gas phase

Transport and diffusion (the first and last two steps) are fast compared to adsorption, reaction, and desorption because of the shell structure of the catalyst. It is known from various investigations that the rate-determining step of the ammonia synthesis is the dissociation of nitrogen.[3] In contrast, exchange reactions between hydrogen and deuterium on the Haber–Bosch catalysts still take place at temperatures of −196 °C (−320.8 °F) at a measurable rate; the exchange between deuterium and hydrogen on the ammonia molecule also takes place at room temperature. Since the adsorption of both molecules is rapid, it cannot determine the speed of ammonia synthesis.[60]

In addition to the reaction conditions, the adsorption of nitrogen on the catalyst surface depends on the microscopic structure of the catalyst surface. Iron has different crystal surfaces, whose reactivity is very different. The Fe(111) and Fe(211) surfaces have by far the highest activity. The explanation for this is that only these surfaces have so-called C7 sites – these are iron atoms with seven closest neighbours.[3]

The dissociative adsorption of nitrogen on the surface follows the following scheme, where S* symbolizes an iron atom on the surface of the catalyst:[42]

N2 → S*–N2 (γ-species) → S*–N2–S* (α-species) → 2 S*–N (β-species, surface nitride)

The adsorption of nitrogen is similar to the chemisorption of carbon monoxide. On a Fe(111) surface, the adsorption of nitrogen first leads to an adsorbed γ-species with an adsorption energy of 24 kJmol−1 and an N-N stretch vibration of 2100 cm−1. Since the nitrogen is isoelectronic to carbon monoxide, it adsorbs in an on-end configuration in which the molecule is bound perpendicular to the metal surface at one nitrogen atom.[18][61][3] This has been confirmed by photoelectron spectroscopy.[62]

Ab-initio-MO calculations have shown that, in addition to the σ binding of the free electron pair of nitrogen to the metal, there is a π binding from the d orbitals of the metal to the π* orbitals of nitrogen, which strengthens the iron-nitrogen bond. The nitrogen in the α state is more strongly bound with 31 kJmol−1. The resulting N–N bond weakening could be experimentally confirmed by a reduction of the wave numbers of the N–N stretching oscillation to 1490 cm−1.[61]

Further heating of the Fe(111) area covered by α-N2 leads to both desorption and the emergence of a new band at 450 cm−1. This represents a metal-nitrogen oscillation, the β state. A comparison with vibration spectra of complex compounds allows the conclusion that the N2 molecule is bound "side-on", with an N atom in contact with a C7 site. This structure is called "surface nitride". The surface nitride is very strongly bound to the surface.[62] Hydrogen atoms (Hads), which are very mobile on the catalyst surface, quickly combine with it.

Infrared spectroscopically detected surface imides (NHad), surface amides (NH2,ad) and surface ammoniacates (NH3,ad) are formed, the latter decay under NH3 release (desorption).[52] The individual molecules were identified or assigned by X-ray photoelectron spectroscopy (XPS), high-resolution electron energy loss spectroscopy (HREELS) and Ir Spectroscopy.

 
Drawn reaction scheme

On the basis of these experimental findings, the reaction mechanism is believed to involve the following steps (see also figure):[63]

  1. N2 (g) → N2 (adsorbed)
  2. N2 (adsorbed) → 2 N (adsorbed)
  3. H2 (g) → H2 (adsorbed)
  4. H2 (adsorbed) → 2 H (adsorbed)
  5. N (adsorbed) + 3 H (adsorbed) → NH3 (adsorbed)
  6. NH3 (adsorbed) → NH3 (g)

Reaction 5 occurs in three steps, forming NH, NH2, and then NH3. Experimental evidence points to reaction 2 as being slow, rate-determining step. This is not unexpected, since the bond is broken, the nitrogen triple bond is the strongest of the bonds that must be broken.

As with all Haber–Bosch catalysts, nitrogen dissociation is the rate-determining step for ruthenium-activated carbon catalysts. The active center for ruthenium is a so-called B5 site, a 5-fold coordinated position on the Ru(0001) surface where two ruthenium atoms form a step edge with three ruthenium atoms on the Ru(0001) surface.[64] The number of B5 sites depends on the size and shape of the ruthenium particles, the ruthenium precursor and the amount of ruthenium used.[47] The reinforcing effect of the basic carrier used in the ruthenium catalyst is similar to the promoter effect of alkali metals used in the iron catalyst.[47]

Energy diagram

An energy diagram can be created based on the Enthalpy of Reaction of the individual steps. The energy diagram can be used to compare homogeneous and heterogeneous reactions: Due to the high activation energy of the dissociation of nitrogen, the homogeneous gas phase reaction is not realizable. The catalyst avoids this problem as the energy gain resulting from the binding of nitrogen atoms to the catalyst surface overcompensates for the necessary dissociation energy so that the reaction is finally exothermic. Nevertheless, the dissociative adsorption of nitrogen remains the rate-determining step: not because of the activation energy, but mainly because of the unfavorable pre-exponential factor of the rate constant. Although hydrogenation is endothermic, this energy can easily be applied by the reaction temperature (about 700 K).[3]

Economic and environmental aspects

External video
  How Earth's Population Exploded Bloomberg Quicktake
 
Severnside fertilizer plant northwest of Bristol, UK

When first invented, the Haber process competed against another industrial process, the cyanamide process. However, the cyanamide process consumed large amounts of electrical power and was more labor-intensive than the Haber process.[7]: 137–143 

As of 2018, the Haber process produces 230 million tonnes of anhydrous ammonia per year.[65] The ammonia is used mainly as a nitrogen fertilizer as ammonia itself, in the form of ammonium nitrate, and as urea. The Haber process consumes 3–5% of the world's natural gas production (around 1–2% of the world's energy supply).[6][66][67][68] In combination with advances in breeding, herbicides, and pesticides, these fertilizers have helped to increase the productivity of agricultural land:

With average crop yields remaining at the 1900 level the crop harvest in the year 2000 would have required nearly four times more land and the cultivated area would have claimed nearly half of all ice-free continents, rather than under 15% of the total land area that is required today.[69]

The energy-intensity of the process contributes to climate change and other environmental problems such as the leaching of nitrates into groundwater, rivers, ponds, and lakes; expanding dead zones in coastal ocean waters, resulting from recurrent eutrophication; atmospheric deposition of nitrates and ammonia affecting natural ecosystems; higher emissions of nitrous oxide (N2O), now the third most important greenhouse gas following CO2 and CH4.[69] The Haber–Bosch process is one of the largest contributors to a buildup of reactive nitrogen in the biosphere, causing an anthropogenic disruption to the nitrogen cycle.[70]

Since nitrogen use efficiency is typically less than 50%,[71] farm runoff from heavy use of fixed industrial nitrogen disrupts biological habitats.[6][72]

Nearly 50% of the nitrogen found in human tissues originated from the Haber–Bosch process.[73] Thus, the Haber process serves as the "detonator of the population explosion", enabling the global population to increase from 1.6 billion in 1900 to 7.7 billion by November 2018.[74]

Reverse fuel cell technology converts renewable energy, water and air into ammonia without a separate hydrogen electrolysis process.[75]

Research

Researchers designed and synthesized metal-organic frameworks (MOF) that bind and release ammonia at lower pressures and temperatures (~175 C). The MOF does not bind the other reactants. It uses copper atoms linked by cyclohexanedicarboxylate. Ammonia converted the MOF into a bundle copper and ammonia-containing polymer strands storing large amounts of ammonia. Lower temperatures and pressures make possible the use of smaller, local facilities.[76]

Other researchers are working to produce hydrogen using renewable energy and other catalysts that work at lower temperatures and pressures. Still others are working on adapting carbon capture and storage technology for use in ammonia plants.

It is expected that an early use of quantum computing will be modeling that improves the efficiency of the Haber–Bosch process[77] by the mid 2020s[78] although some have predicted it will take longer.[79]

See also

Notes

  1. ^ Hydrogen required for ammonia synthesis is most often produced through gasification of carbon-containing material, mostly natural gas, but other potential carbon sources include coal, petroleum, peat, biomass, or waste. As of 2012, the global production of ammonia produced from natural gas using the steam reforming process was 72%.[21] Hydrogen can also be produced from water and electricity using electrolysis: at one time, most of Europe's ammonia was produced from the Hydro plant at Vemork. Other possibilities include biological hydrogen production or photolysis, but at present, steam reforming of natural gas is the most economical means of mass-producing hydrogen.

References

  1. ^ Habers process chemistry. India: Arihant publications. 2018. p. 264. ISBN 978-93-131-6303-9.
  2. ^ Appl, M. (1982). "The Haber–Bosch Process and the Development of Chemical Engineering". A Century of Chemical Engineering. New York: Plenum Press. pp. 29–54. ISBN 978-0-306-40895-3.
  3. ^ a b c d e f g h i j k l m n o p q r Appl, Max (2006). "Ammonia". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1002/14356007.a02_143.pub2.
  4. ^ Clark 2013, "The forward reaction (the production of ammonia) is exothermic. According to Le Chatelier's Principle, this will be favoured at a lower temperature. The system will respond by moving the position of equilibrium to counteract this – in other words by producing more heat. To obtain as much ammonia as possible in the equilibrium mixture, as low a temperature as possible is needed".
  5. ^ Clark 2013, "Notice that there are 4 molecules on the left-hand side of the equation, but only 2 on the right. According to Le Chatelier's Principle, by increasing the pressure the system will respond by favouring the reaction which produces fewer molecules. That will cause the pressure to fall again. To get as much ammonia as possible in the equilibrium mixture, as high a pressure as possible is needed. 200 atmospheres is a high pressure, but not amazingly high".
  6. ^ a b c Smil, Vaclav (2004). Enriching the Earth: Fritz Haber, Carl Bosch, and the Transformation of World Food Production (1st ed.). Cambridge, MA: MIT. ISBN 978-0-262-69313-4.
  7. ^ a b c d e Hager, Thomas (2008). The Alchemy of Air: A Jewish genius, a doomed tycoon, and the scientific discovery that fed the world but fueled the rise of Hitler (1st ed.). New York, NY: Harmony Books. ISBN 978-0-307-35178-4.
  8. ^ Sittig, Marshall (1979). Fertilizer Industry: Processes, Pollution Control, and Energy Conservation. Park Ridge, NJ: Noyes Data Corp. ISBN 978-0-8155-0734-5.
  9. ^ Vandermeer, John (2011). The Ecology of Agroecosystems. Jones & Bartlett Learning. p. 149. ISBN 978-0-7637-7153-9.
  10. ^ James, Laylin K. (1993). Nobel Laureates in Chemistry 1901–1992 (3rd ed.). Washington, DC: American Chemical Society. p. 118. ISBN 978-0-8412-2690-6.
  11. ^ Haber, Fritz (1905). Thermodynamik technischer Gasreaktionen (in German) (1st ed.). Paderborn: Salzwasser Verlag. ISBN 978-3-86444-842-3.
  12. ^ (PDF), ChemUCL Newsletter, p. 8, 2009, archived from the original (PDF) on 13 January 2011.
  13. ^ Bosch, Carl (2 March 1908) "Process of producing ammonia." U.S. Patent 990,191.
  14. ^ Philip, Phylis Morrison (2001). . American Scientist. Archived from the original on 2 July 2012.
  15. ^ (PDF). The New York Times. 3 February 1920. Archived from the original (PDF) on 24 February 2021. Retrieved 11 October 2010.
  16. ^ Bozso, F.; Ertl, G.; Grunze, M.; Weiss, M. (1977). "Interaction of nitrogen with iron surfaces: I. Fe(100) and Fe(111)". Journal of Catalysis. 49 (1): 18–41. doi:10.1016/0021-9517(77)90237-8.
  17. ^ Imbihl, R.; Behm, R. J.; Ertl, G.; Moritz, W. (1982). "The structure of atomic nitrogen adsorbed on Fe(100)" (PDF). Surface Science. 123 (1): 129–140. Bibcode:1982SurSc.123..129I. doi:10.1016/0039-6028(82)90135-2.
  18. ^ a b Ertl, G.; Lee, S. B.; Weiss, M. (1982). "Kinetics of nitrogen adsorption on Fe(111)". Surface Science. 114 (2–3): 515–526. Bibcode:1982SurSc.114..515E. doi:10.1016/0039-6028(82)90702-6.
  19. ^ Ertl, G. (1983). "Primary steps in catalytic synthesis of ammonia". Journal of Vacuum Science and Technology A. 1 (2): 1247–1253. Bibcode:1983JVSTA...1.1247E. doi:10.1116/1.572299.
  20. ^ "100 years of thyssenkrupp Uhde". Industrial Solutions (in German). Retrieved 8 December 2021.
  21. ^ "Ammonia". Industrial Efficiency Technology & Measures. 30 April 2013. Retrieved 6 April 2018.
  22. ^ "Electrochemically-produced ammonia could revolutionize food production". 9 July 2018. Retrieved 15 December 2018. Ammonia manufacturing consumes 1 to 2% of total global energy and is responsible for approximately 3% of global carbon dioxide emissions.
  23. ^ Song, Yang; Hensley, Dale; Bonnesen, Peter; Liang, Liango; Huang, Jingsong; Baddorf, Arthur; Tschaplinski, Timothy; Engle, Nancy; Wu, Zili; Cullen, David; Meyer, Harry III; Sumpter, Bobby; Rondinone, Adam (2 May 2018). "A physical catalyst for the electrolysis of nitrogen to ammonia". Science Advances. Oak Ridge National Laboratory. 4 (4): e1700336. Bibcode:2018SciA....4E0336S. doi:10.1126/sciadv.1700336. PMC 5922794. PMID 29719860. Retrieved 15 December 2018. Ammonia synthesis consumes 3 to 5% of the world's natural gas, making it a significant contributor to greenhouse gas emissions.
  24. ^ Kuganathan, Navaratnarajah; Hosono, Hideo; Shluger, Alexander L.; Sushko, Peter V. (January 2014). "Enhanced N2 Dissociation on Ru-Loaded Inorganic Electride". Journal of the American Chemical Society. 136 (6): 2216–2219. doi:10.1021/ja410925g. PMID 24483141.
  25. ^ Hara, Michikazu; Kitano, Masaaki; Hosono, Hideo; Sushko, Peter V. (2017). "Ru-Loaded C12A7:e– Electride as a Catalyst for Ammonia Synthesi". ACS Catalysis. 7 (4): 2313–2324. doi:10.1021/acscatal.6b03357.
  26. ^ "Ajinomoto Co., Inc., UMI, and Tokyo Institute of Technology Professors Establish New Company to implement the World's First On Site Production of Ammonia". Ajinomoto. 27 April 2017. Retrieved 9 November 2021.
  27. ^ Stephen H. Crolius (17 December 2020). "Tsubame BHB Launches Joint Evaluation with Mitsubishi Chemical". Ammonia Energy Association. Retrieved 9 November 2021.
  28. ^ Kitano, Masaaki; Kujirai, Jun; Ogasawara, Kiya; Matsuishi, Satoru; Tada, Tomofumi; Abe, Hitoshi; Niwa, Yasuhiro; Hosono, Hideo (2019). "Low-Temperature Synthesis of Perovskite Oxynitride-Hydrides as Ammonia Synthesis Catalysts". Journal of the American Chemical Society. 141 (51): 20344–20353. doi:10.1021/jacs.9b10726. PMID 31755269. S2CID 208227325.
  29. ^ Wang, Ying; Meyer, Thomas J. (14 March 2019). "A Route to Renewable Energy Triggered by the Haber–Bosch Process". Chem. 5 (3): 496–497. doi:10.1016/j.chempr.2019.02.021. S2CID 134713643.
  30. ^ Schneider, Stefan; Bajohr, Siegfried; Graf, Frank; Kolb, Thomas (13 January 2020). "State of the Art of Hydrogen Production via Pyrolysis of Natural Gas". ChemBioEng Reviews. 7 (5): 150–158. doi:10.1002/cben.202000014. S2CID 221708661 – via Wiley Online Library.
  31. ^ "Progress in the Electrochemical Synthesis of Ammonia | Request PDF".
  32. ^ Clark 2013, However, 400–450 °C isn't a low temperature! Rate considerations: The lower the temperature you use, the slower the reaction becomes. A manufacturer is trying to produce as much ammonia as possible per day. It makes no sense to try to achieve an equilibrium mixture which contains a very high proportion of ammonia if it takes several years for the reaction to reach that equilibrium"..
  33. ^ Clark 2013, "Rate considerations: Increasing the pressure brings the molecules closer together. In this particular instance, it will increase their chances of hitting and sticking to the surface of the catalyst where they can react. The higher the pressure the better in terms of the rate of a gas reaction. Economic considerations: Very high pressures are expensive to produce on two counts. Extremely strong pipes and containment vessels are needed to withstand the very high pressure. That increases capital costs when the plant is built".
  34. ^ "Chemistry of Nitrogen". Compounds. Chem.LibreTexts.org. 5 June 2019. Retrieved 7 July 2019.
  35. ^ Clark 2013, "At each pass of the gases through the reactor, only about 15% of the nitrogen and hydrogen converts to ammonia. (This figure also varies from plant to plant.) By continual recycling of the unreacted nitrogen and hydrogen, the overall conversion is about 98%".
  36. ^ Brown, Theodore L.; LeMay, H. Eugene Jr.; Bursten, Bruce E. (2006). "Table 15.2". Chemistry: The Central Science (10th ed.). Upper Saddle River, NJ: Pearson. ISBN 978-0-13-109686-8.
  37. ^ Abild-pedersen, Frank; Bligaard, Thomas (1 January 2014). "Exploring the limits: A low-pressure, low-temperature Haber–Bosch process". Chemical Physics Letters. 598: 108. Bibcode:2014CPL...598..108V. doi:10.1016/j.cplett.2014.03.003 – via academia.edu.
  38. ^ Koop, Fermin (13 January 2023). "Green ammonia (and fertilizer) may finally be in sight -- and it would be huge". ZME Science. Retrieved 21 March 2023.
  39. ^ Mittasch, Alwin (1926). "Bemerkungen zur Katalyse". Berichte der Deutschen Chemischen Gesellschaft (A and B Series). 59: 13–36. doi:10.1002/cber.19260590103.
  40. ^ "3.1 Ammonia synthesis". resources.schoolscience.co.uk. from the original on 6 July 2020.
  41. ^ Rock, Peter A. (19 June 2013). Chemical Thermodynamics. p. 317. ISBN 978-1-891389-32-0.
  42. ^ a b c Max Appl (2006). "Ammonia". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1002/14356007.a02_143.pub2.
  43. ^ Jozwiak, W. K.; Kaczmarek, E. (2007). "Reduction behavior of iron oxides in hydrogen and carbon monoxide atmospheres". Applied Catalysis A: General. 326: 17–27. doi:10.1016/j.apcata.2007.03.021.
  44. ^ Ertl, Gerhard (1983). "Zum Mechanismus der Ammoniak-Synthese". Nachrichten aus Chemie, Technik und Laboratorium. 31 (3): 178–182. doi:10.1002/nadc.19830310307.
  45. ^ Bowker, Michael (1993). "Chapter 7". In King, D. A.; Woodruff, D. P. (eds.). The Chemical Physics of Solid Surfaces. Vol. 6: Coadsorption, promoters and poisons. Elsevier. pp. 225–268. ISBN 978-0-444-81468-5.
  46. ^ Tavasoli, Ahmad; Trépanier, Mariane; Malek Abbaslou, Reza M.; Dalai, Ajay K.; Abatzoglou, Nicolas (1 December 2009). "Fischer–Tropsch synthesis on mono- and bimetallic Co and Fe catalysts supported on carbon nanotubes". Fuel Processing Technology. 90 (12): 1486–1494. doi:10.1016/j.fuproc.2009.07.007. ISSN 0378-3820.
  47. ^ a b c d You, Zhixiong; Inazu, Koji; Aika, Ken-ichi; Baba, Toshihide (October 2007). "Electronic and structural promotion of barium hexaaluminate as a ruthenium catalyst support for ammonia synthesis". Journal of Catalysis. 251 (2): 321–331. doi:10.1016/j.jcat.2007.08.006.
  48. ^ Rosowski, F.; Hornung, A.; Hinrichsen, O.; Herein, D.; Muhler, M. (April 1997). "Ruthenium catalysts for ammonia synthesis at high pressures: Preparation, characterization, and power-law kinetics". Applied Catalysis A: General. 151 (2): 443–460. doi:10.1016/S0926-860X(96)00304-3.
  49. ^ Højlund Nielsen, P. E. (1995), Nielsen, Anders (ed.), "Poisoning of Ammonia Synthesis Catalysts", Ammonia: Catalysis and Manufacture, Berlin, Heidelberg: Springer, pp. 191–198, doi:10.1007/978-3-642-79197-0_5, ISBN 978-3-642-79197-0, retrieved 30 July 2022
  50. ^ a b Falbe, Jürgen (1997). Römpp-Lexikon Chemie (H–L). Georg Thieme Verlag. pp. 1644–1646. ISBN 978-3-13-107830-8.
  51. ^ Brown, Theodore L.; LeMay, H. Eugene; Bursten, Bruce Edward (2003). Brunauer, Linda Sue (ed.). Chemistry the Central Science (9th ed.). Upper Saddle River, NJ: Prentice Hall. ISBN 978-0-13-038168-2.
  52. ^ a b c Holleman, Arnold Frederik; Wiberg, Egon (2001), Wiberg, Nils (ed.), Inorganic Chemistry, translated by Eagleson, Mary; Brewer, William, San Diego/Berlin: Academic Press/De Gruyter, pp. 662–65, ISBN 0-12-352651-5
  53. ^ Cornils, Boy; Herrmann, Wolfgang A.; Muhler, M.; Wong, C. (2007). Catalysis from A to Z: A Concise Encyclopedia. Verlag Wiley-VCH. p. 31. ISBN 978-3-527-31438-6.
  54. ^ Fokus Chemie Oberstufe Einführungsphase. Berlin: Cornelsen-Verlag. 2010. p. 79. ISBN 978-3-06-013953-8.
  55. ^ P. Häussinger u. a.: Noble Gases. In: Ullmann’s Encyclopedia of Industrial Chemistry. Wiley-VCH, Weinheim 2006. doi:10.1002/14356007.a17_485
  56. ^ Leibnitz, E.; Koch, H.; Götze, A. (1961). "Über die drucklose Aufbereitung von Braunkohlenkokereigas auf Starkgas nach dem Girbotol-Verfahren". Journal für Praktische Chemie. 13 (3–4): 215–236. doi:10.1002/prac.19610130315.
  57. ^ Steinborn, Dirk (2007). Grundlagen der metallorganischen Komplexkatalyse. Wiesbaden: Teubner. pp. 319–321. ISBN 978-3-8351-0088-6.
  58. ^ a b c Forst, Detlef; Kolb, Maximillian; Roßwag, Helmut (1993). Chemie für Ingenieure. Springer Verlag. pp. 234–238. ISBN 978-3-662-00655-9.
  59. ^ "Ammoniakkonverter - Düngemittelanlagen". Industrial Solutions (in German). Retrieved 8 December 2021.
  60. ^ Moore, Walter J.; Hummel, Dieter O. (1983). Physikalische Chemie. Berlin: Walter de Gruyter. p. 604. ISBN 978-3-11-008554-9.
  61. ^ a b Lee, S. B.; Weiss, M. (1982). "Adsorption of nitrogen on potassium promoted Fe(111) and (100) surfaces". Surface Science. 114 (2–3): 527–545. Bibcode:1982SurSc.114..527E. doi:10.1016/0039-6028(82)90703-8.
  62. ^ a b Ertl, Gerhard (2010). Reactions at Solid Surfaces. John Wiley & Sons. p. 123. ISBN 978-0-470-26101-9.
  63. ^ Wennerström, Håkan; Lidin, Sven. "Scientific Background on the Nobel Prize in Chemistry 2007 Chemical Processes on Solid Surfaces" (PDF). Nobel Foundation. Retrieved 17 September 2015.
  64. ^ Gavnholt, Jeppe; Schiøtz, Jakob (2008). "Structure and reactivity of ruthenium nanoparticles" (PDF). Physical Review B. 77 (3): 035404. Bibcode:2008PhRvB..77c5404G. doi:10.1103/PhysRevB.77.035404. S2CID 49236953.
  65. ^ "Ammonia annual production capacity globally 2030". Statista. Retrieved 7 May 2020.
  66. ^ "International Energy Outlook 2007". eia.gov. U.S. Energy Information Administration.
  67. ^ Fertilizer statistics. . www.fertilizer.org. International Fertilizer Industry Association. Archived from the original on 24 April 2008.
  68. ^ Smith, Barry E. (September 2002). "Structure. Nitrogenase reveals its inner secrets". Science. 297 (5587): 1654–5. doi:10.1126/science.1076659. PMID 12215632. S2CID 82195088.
  69. ^ a b Smil, Vaclav (2011). "Nitrogen cycle and world food production" (PDF). World Agriculture. 2: 9–13.
  70. ^ Kanter, David R.; Bartolini, Fabio; Kugelberg, Susanna; Leip, Adrian; Oenema, Oene; Uwizeye, Aimable (2 December 2019). "Nitrogen pollution policy beyond the farm". Nature Food. 1: 27–32. doi:10.1038/s43016-019-0001-5. ISSN 2662-1355.
  71. ^ Oenema, O.; Witzke, H. P.; Klimont, Z.; Lesschen, J. P.; Velthof, G. L. (2009). "Integrated assessment of promising measures to decrease nitrogen losses in agriculture in EU-27". Agriculture, Ecosystems and Environment. 133 (3–4): 280–288. doi:10.1016/j.agee.2009.04.025.
  72. ^ Howarth, R. W. (2008). "Coastal nitrogen pollution: a review of sources and trends globally and regionally". Harmful Algae. 8: 14–20. doi:10.1016/j.hal.2008.08.015.
  73. ^ Ritter, Steven K. (18 August 2008). "The Haber–Bosch Reaction: An Early Chemical Impact On Sustainability". Chemical & Engineering News. 86 (33).
  74. ^ Smil, Vaclav (1999). "Detonator of the population explosion" (PDF). Nature. 400 (6743): 415. Bibcode:1999Natur.400..415S. doi:10.1038/22672. S2CID 4301828.
  75. ^ Blain, Loz (3 September 2021). "Green ammonia: The rocky pathway to a new clean fuel". New Atlas. Retrieved 23 March 2023.
  76. ^ Koop, Fermin (13 January 2023). "Green ammonia (and fertilizer) may finally be in sight -- and it would be huge". ZME Science. Retrieved 23 March 2023.
  77. ^ Ruane, Jonathan; McAfee, Andrew; Oliver, William D. (1 January 2022). "Quantum Computing for Business Leaders". Harvard Business Review. ISSN 0017-8012. Retrieved 12 April 2023.
  78. ^ Budde, Florian; Volz, Daniel (12 July 2019). "Quantum computing and the chemical industry | McKinsey". www.mckinsey.com. McKinsey and Company. Retrieved 12 April 2023.
  79. ^ Bourzac, Katherine (30 October 2017). "Chemistry is quantum computing's killer app". cen.acs.org. American Chemical Society. Retrieved 12 April 2023.

Sources

  • Clark, Jim (April 2013) [2002]. "The Haber Process". Retrieved 15 December 2018.

External links

  • "Detonator of the population explosion" (PDF). Vaclav Smil, Department of Geography, University of Manitoba. Macmillan Magazines Ltd.
  • "Review of "Between Genius and Genocide: The Tragedy of Fritz Haber, Father of Chemical Warfare" (PDF). Daniel Charles.
  • "The Haber Process". Chemguide.co.uk.
  • BASF – Fertilizer out of thin air
  • Britannica guide to Nobel Prizes: Fritz Haber
  • Haber Process for Ammonia Synthesis
  • Haber–Bosch process, most important invention of the 20th century, according to V. Smil, Nature, 29 July 1999, p. 415 (by Jürgen Schmidhuber)
  • Nobel e-Museum – Biography of Fritz Haber
  • Uses and Production of Ammonia

haber, process, also, called, haber, bosch, process, main, industrial, procedure, production, ammonia, named, after, inventors, german, chemists, fritz, haber, carl, bosch, developed, first, decade, 20th, century, process, converts, atmospheric, nitrogen, ammo. The Haber process 1 also called the Haber Bosch process is the main industrial procedure for the production of ammonia 2 3 It is named after its inventors the German chemists Fritz Haber and Carl Bosch who developed it in the first decade of the 20th century The process converts atmospheric nitrogen N2 to ammonia NH3 by a reaction with hydrogen H2 using a metal catalyst under high temperatures and pressures This reaction is slightly exothermic i e it releases energy meaning that the reaction is favoured at lower temperatures 4 and higher pressures 5 It decreases entropy complicating the process Hydrogen is produced via steam reforming followed by an iterative closed cycle to react hydrogen with nitrogen to produce ammonia Fritz Haber 1918 The primary reaction is N 2 3 H 2 2 NH 3 D H 91 8 kJ mol displaystyle ce N2 3 H2 gt 2 NH3 quad Delta H circ 91 8 text kJ mol Before the development of the Haber process it had been difficult to produce ammonia on an industrial scale 6 7 8 because earlier methods such as the Birkeland Eyde process and the Frank Caro process were too inefficient Contents 1 History 2 Process 2 1 Hydrogen production 2 2 Ammonia production 2 3 Pressure temperature 3 Catalysts 3 1 Iron based catalysts 3 2 Catalysts other than iron 3 2 1 Ruthenium 3 3 Catalyst poisons 4 Industrial production 4 1 Synthesis parameters 4 2 Large scale implementation 5 Mechanism 5 1 Elementary steps 5 2 Energy diagram 6 Economic and environmental aspects 7 Research 8 See also 9 Notes 10 References 11 Sources 12 External linksHistory EditMain article History of the Haber process Carl Bosch 1927 During the 19th century the demand for nitrates and ammonia for use as fertilizers and industrial feedstocks rapidly increased The main source was mining niter deposits and guano from tropical islands 9 At the beginning of the 20th century these reserves were thought insufficient to satisfy future demands 10 and research into new potential sources of ammonia increased Although atmospheric nitrogen N2 is abundant comprising 78 of the air it is exceptionally stable and does not readily react with other chemicals Haber with his assistant Robert Le Rossignol developed the high pressure devices and catalysts needed to demonstrate the Haber process at a laboratory scale 11 12 They demonstrated their process in the summer of 1909 by producing ammonia from the air drop by drop at the rate of about 125 mL 4 US fl oz per hour The process was purchased by the German chemical company BASF which assigned Carl Bosch the task of scaling up Haber s tabletop machine to industrial scale 7 13 He succeeded in 1910 Haber and Bosch were later awarded Nobel Prizes in 1918 and 1931 respectively for their work in overcoming the chemical and engineering problems of large scale continuous flow high pressure technology 7 Ammonia was first manufactured using the Haber process on an industrial scale in 1913 in BASF s Oppau plant in Germany reaching 20 tonnes day in 1914 14 During World War I the production of munitions required large amounts of nitrate The Allies had access to large deposits of sodium nitrate in Chile Chile saltpetre controlled by British companies Germany had no such resources so the Haber process proved essential to the German war effort 7 15 Synthetic ammonia from the Haber process was used for the production of nitric acid a precursor to the nitrates used in explosives The original Haber Bosch reaction chambers used osmium as the catalyst but it was available in extremely small quantities Haber noted uranium was almost as effective and easier to obtain than osmium In 1909 BASF researcher Alwin Mittasch discovered a much less expensive iron based catalyst that is still used A major contributor to the elucidation of this catalysis was Gerhard Ertl 16 17 18 19 The most popular catalysts are based on iron promoted with K2O CaO SiO2 and Al2O3 During the interwar years alternative processes were developed most notably the Casale process Claude process and the Mont Cenis process developed by Friedrich Uhde Ingenieurburo 20 Luigi Casale and Georges Claude proposed to increase the pressure of the synthesis loop to 80 100 MPa 800 1 000 bar 12 000 15 000 psi thereby increasing the single pass ammonia conversion and making nearly complete liquefaction at ambient temperature feasible Claude proposed to have three or four converters with liquefaction steps in series thereby avoiding recycling Most plants continue to use the original Haber process 20 MPa 200 bar 2 900 psi and 500 C 932 F albeit with improved single pass conversion and lower energy consumption due to process and catalyst optimization Process Edit A historical 1921 high pressure steel reactor for the production of ammonia via the Haber process is displayed at the Karlsruhe Institute of Technology Germany Combined with the energy needed to produce hydrogen note 1 and purified atmospheric nitrogen ammonia production is energy intensive accounting for 1 to 2 of global energy consumption 3 of global carbon emissions 22 and 3 to 5 of natural gas consumption 23 The choice of catalyst is important for synthesizing ammonia In 2012 Hideo Hosono s group found that Ru loaded calcium aluminum oxide C12A7 e electride works well as a catalyst and pursued more efficient formation 24 25 This method is implemented in a small plant for ammonia synthesis in Japan 26 27 In 2019 Hosono s group found another catalyst a novel perovskite oxynitride hydride BaCeO3 xNyHz that works at lower temperature and without costly ruthenium 28 Hydrogen production Edit The major source of hydrogen is methane Steam reforming extracts hydrogen from methane in a high temperature and pressure tube inside a reformer with a nickel catalyst Other fossil fuel sources include coal heavy fuel oil and naphtha Green hydrogen is produced without fossil fuels or carbon dioxide emissions from biomass water electrolysis and thermochemical solar or another heat source water splitting However these hydrogen sources are not economically competitive with steam reforming 29 30 31 Starting with a natural gas CH4 feedstock the steps are Remove sulfur compounds from the feedstock because sulfur deactivates the catalysts used in subsequent steps Sulfur removal requires catalytic hydrogenation to convert sulfur compounds in the feedstocks to gaseous hydrogen sulfide H2 RSH RH H2S gas dd Hydrogen sulfide is adsorbed and removed by passing it through beds of zinc oxide where it is converted to solid zinc sulfide Illustrating inputs and outputs of steam reforming of natural gas a process to produce hydrogen H2S ZnO ZnS H2O dd Catalytic steam reforming of the sulfur free feedstock forms hydrogen plus carbon monoxide CH4 H2O CO 3 H2 dd Catalytic shift conversion converts the carbon monoxide to carbon dioxide and more hydrogen CO H2O CO2 H2 dd Carbon dioxide is removed either by absorption in aqueous ethanolamine solutions or by adsorption in pressure swing adsorbers PSA using proprietary solid adsorption media The final step in producing hydrogen is to use catalytic methanation to remove residual carbon monoxide or carbon dioxide CO 3 H2 CH4 H2O CO2 4 H2 CH4 2 H2O dd Ammonia production Edit The hydrogen is catalytically reacted with nitrogen derived from process air to form anhydrous liquid ammonia It is difficult and expensive as lower temperatures result in slower reaction kinetics hence a slower reaction rate 32 and high pressure requires high strength pressure vessels 33 that resist hydrogen embrittlement Diatomic nitrogen is bound together by a triple bond which makes it relatively inert 34 Yield and efficiency are low meaning that the ammonia must be extracted and the gases reprocessed for the reaction to proceed at an acceptable pace 35 This step is known as the ammonia synthesis loop 3 H2 N2 2 NH3 dd The gases nitrogen and hydrogen are passed over four beds of catalyst with cooling between each pass to maintain a reasonable equilibrium constant On each pass only about 15 conversion occurs but unreacted gases are recycled and eventually conversion of 97 is achieved 3 Due to the nature of the typically multi promoted magnetite catalyst used in the ammonia synthesis reaction only low levels of oxygen containing especially CO CO2 and H2O compounds can be tolerated in the hydrogen nitrogen mixture Relatively pure nitrogen can be obtained by air separation but additional oxygen removal may be required Because of relatively low single pass conversion rates typically less than 20 a large recycle stream is required This can lead to the accumulation of inerts in the gas Nitrogen gas N2 is unreactive because the atoms are held together by triple bonds The Haber process relies on catalysts that accelerate the scission of these bonds Two opposing considerations are relevant the equilibrium position and the reaction rate At room temperature the equilibrium is in favor of ammonia but the reaction doesn t proceed at a detectable rate due to its high activation energy Because the reaction is exothermic the equilibrium constant becomes unity at around 150 200 C 302 392 F following Le Chatelier s principle 3 K T for N2 3 H2 2 NH3 36 Temperature C Kp300 4 34 10 3400 1 64 10 4450 4 51 10 5500 1 45 10 5550 5 38 10 6600 2 25 10 6Above this temperature the equilibrium quickly becomes unfavorable at atmospheric pressure according to the Van t Hoff equation Lowering the temperature is unhelpful because the catalyst requires a temperature of at least 400 C to be efficient 3 Increased pressure favors the forward reaction because 4 moles of reactant produce 2 moles of product and the pressure used 15 25 MPa 150 250 bar 2 200 3 600 psi alters the equilibrium concentrations to give a substantial ammonia yield The reason for this is evident in the equilibrium relationship K y NH 3 2 y H 2 3 y N 2 ϕ NH 3 2 ϕ H 2 3 ϕ N 2 P P 2 displaystyle K frac y ce NH3 2 y ce H2 3 y ce N2 frac hat phi ce NH3 2 hat phi ce H2 3 hat phi ce N2 left frac P circ P right 2 where ϕ i displaystyle hat phi i is the fugacity coefficient of species i displaystyle i y i displaystyle y i is the mole fraction of the same species P displaystyle P is the reactor pressure and P displaystyle P circ is standard pressure typically 1 bar 0 10 MPa Economically reactor pressurization is expensive pipes valves and reaction vessels need to be strong enough and safety considerations affect operating at 20 MPa Compressors take considerable energy as work must be done on the compressible gas Thus the compromise used gives a single pass yield of around 15 3 While removing the ammonia from the system increases the reaction yield this step is not used in practice since the temperature is too high instead it is removed from the gases leaving the reaction vessel The hot gases are cooled under high pressure allowing the ammonia to condense and be removed as a liquid Unreacted hydrogen and nitrogen gases are returned to the reaction vessel for another round 3 While most ammonia is removed typically down to 2 5 mol some ammonia remains in the recycle stream In academic literature a more complete separation of ammonia has been proposed by absorption in metal halides or zeolites Such a process is called an absorbent enhanced Haber process or adsorbent enhanced Haber Bosch process 37 Pressure temperature Edit The steam reforming shift conversion carbon dioxide removal and methanation steps each operate at absolute pressures of about 25 to 35 bar while the ammonia synthesis loop operates at temperatures of 300 500 C 572 932 F and pressures ranging from 60 to 180 bar depending upon the method used The resulting ammonia must then be separated from the residual hydrogen and nitrogen at temperatures of 20 C 4 F 38 3 Catalysts Edit First reactor at the Oppau plant in 1913 Profiles of the active components of heterogeneous catalysts the top right figure shows the profile of a shell catalyst The Haber Bosch process relies on catalysts to accelerate N2 hydrogenation The catalysts are heterogeneous solids that interact with gaseous reagents 39 The catalyst typically consists of finely divided iron bound to an iron oxide carrier containing promoters possibly including aluminium oxide potassium oxide calcium oxide potassium hydroxide 40 molybdenum 41 and magnesium oxide Iron based catalysts Edit The iron catalyst is obtained from finely ground iron powder which is usually obtained by reduction of high purity magnetite Fe3O4 The pulverized iron is oxidized to give magnetite or wustite FeO ferrous oxide particles of a specific size The magnetite or wustite particles are then partially reduced removing some of the oxygen The resulting catalyst particles consist of a core of magnetite encased in a shell of wustite which in turn is surrounded by an outer shell of metallic iron The catalyst maintains most of its bulk volume during the reduction resulting in a highly porous high surface area material which enhances its catalytic effectiveness Minor components include calcium and aluminium oxides which support the iron catalyst and help it maintain its surface area These oxides of Ca Al K and Si are unreactive to reduction by hydrogen 3 The production of the catalyst requires a particular melting process in which used raw materials must be free of catalyst poisons and the promoter aggregates must be evenly distributed in the magnetite melt Rapid cooling of the magnetite which has an initial temperature of about 3500 C produces the desired precursor Unfortunately the rapid cooling ultimately forms a catalyst of reduced abrasion resistance Despite this disadvantage the method of rapid cooling is often employed 3 The reduction of the precursor magnetite to a iron is carried out directly in the production plant with synthesis gas The reduction of the magnetite proceeds via the formation of wustite FeO so that particles with a core of magnetite become surrounded by a shell of wustite The further reduction of magnetite and wustite leads to the formation of a iron which forms together with the promoters of the outer shell 42 The involved processes are complex and depend on the reduction temperature At lower temperatures wustite disproportionates into an iron phase and a magnetite phase at higher temperatures the reduction of the wustite and magnetite to iron dominates 43 The a iron forms primary crystallites with a diameter of about 30 nanometers These crystallites form a bimodal pore system with pore diameters of about 10 nanometers produced by the reduction of the magnetite phase and of 25 to 50 nanometers produced by the reduction of the wustite phase 42 With the exception of cobalt oxide the promoters are not reduced During the reduction of the iron oxide with synthesis gas water vapor is formed This water vapor must be considered for high catalyst quality as contact with the finely divided iron would lead to premature aging of the catalyst through recrystallization especially in conjunction with high temperatures The vapor pressure of the water in the gas mixture produced during catalyst formation is thus kept as low as possible target values are below 3 gm 3 For this reason the reduction is carried out at high gas exchange low pressure and low temperatures The exothermic nature of the ammonia formation ensures a gradual increase in temperature 3 The reduction of fresh fully oxidized catalyst or precursor to full production capacity takes four to ten days 3 The wustite phase is reduced faster and at lower temperatures than the magnetite phase Fe3O4 After detailed kinetic microscopic and X ray spectroscopic investigations it was shown that wustite reacts first to metallic iron This leads to a gradient of iron II ions whereby these diffuse from the magnetite through the wustite to the particle surface and precipitate there as iron nuclei Pre reduced stabilized catalysts occupy a significant market share They are delivered showing the fully developed pore structure but have been oxidized again on the surface after manufacture and are therefore no longer pyrophoric The reactivation of such pre reduced catalysts requires only 30 to 40 hours instead of several days In addition to the short start up time they have other advantages such as higher water resistance and lower weight 3 Typical catalyst composition 44 Iron Potassium Aluminium Calcium Oxygen Volume composition 40 5 0 0 35 0 2 0 1 7 53 2Surface composition before reduction 0 8 6 36 1 10 7 4 7 40 0Surface composition after reduction 11 0 27 0 17 0 4 0 41 0Catalysts other than iron Edit Many efforts have been made to improve the Haber Bosch process Many metals were tested as catalysts The requirement for suitability is the dissociative adsorption of nitrogen i e the nitrogen molecule must be split into nitrogen atoms upon absorption If the binding of the nitrogen is too strong the catalyst is blocked and the catalytic ability is reduced self poisoning The elements in the periodic table to the left of the iron group show such strong bonds Further the formation of surface nitrides makes for example chromium catalysts ineffective Metals to the right of the iron group in contrast adsorb nitrogen too weakly for ammonia synthesis Haber initially used catalysts based on osmium and uranium Uranium reacts to its nitride during catalysis while osmium oxide is rare 45 According to theoretical and practical studies improvements over pure iron are limited The activity of iron catalysts is increased by the inclusion of cobalt 46 Ruthenium Edit Ruthenium forms highly active catalysts Allowing milder operating pressures and temperatures Ru based materials are referred to as second generation catalysts Such catalysts are prepared by the decomposition of triruthenium dodecacarbonyl on graphite 3 A drawback of activated carbon supported ruthenium based catalysts is the methanation of the support in the presence of hydrogen Their activity is strongly dependent on the catalyst carrier and the promoters A wide range of substances can be used as carriers including carbon magnesium oxide aluminium oxide zeolites spinels and boron nitride 47 Ruthenium activated carbon based catalysts have been used industrially in the KBR Advanced Ammonia Process KAAP since 1992 48 The carbon carrier is partially degraded to methane however this can be mitigated by a special treatment of the carbon at 1500 C thus prolonging the catalyst lifetime In addition the finely dispersed carbon poses a risk of explosion For these reasons and due to its low acidity magnesium oxide has proven to be a good choice of carrier Carriers with acidic properties extract electrons from ruthenium make it less reactive and have the undesirable effect of binding ammonia to the surface 47 Catalyst poisons Edit Catalyst poisons lower catalyst activity They are usually impurities in the synthesis gas Permanent poisons cause irreversible loss of catalytic activity and while temporary poisons lower the activity while present Sulfur compounds phosphorus compounds arsenic compounds and chlorine compounds are permanent poisons Oxygenic compounds like water carbon monoxide carbon dioxide and oxygen are temporary poisons 3 49 Although chemically inert components of the synthesis gas mixture such as noble gases or methane are not strictly poisons they accumulate through the recycling of the process gases and thus lower the partial pressure of the reactants which in turn slows conversion 50 Industrial production EditSynthesis parameters Edit Change of the equilibrium constant Keq as a function of temperature 51 temperature C Keq300 4 34 10 3400 1 64 10 4450 4 51 10 5500 1 45 10 5550 5 38 10 6600 2 25 10 6The reaction is N 2 3 H 2 2 NH 3 D H 92 28 kJ D H 298 K 46 14 k J m o l displaystyle ce N2 3H2 lt gt 2NH3 qquad Delta H circ 92 28 ce kJ Delta H 298 mathrm K circ 46 14 mathrm kJ mol 52 The reaction is an exothermic equilibrium reaction in which the gas volume is reduced The equilibrium constant Keq of the reaction see table is obtained from K e q p 2 NH 3 p N 2 p 3 H 2 displaystyle K eq frac p 2 ce NH3 p ce N2 cdot p 3 ce H2 Since the reaction is exothermic the equilibrium of the reaction shifts at lower temperatures to the ammonia side Furthermore four volumetric units of the raw materials produce two volumetric units of ammonia According to Le Chatelier s principle higher pressure favours ammonia High pressure is necessary to ensure sufficient surface coverage of the catalyst with nitrogen 53 For this reason a ratio of nitrogen to hydrogen of 1 to 3 a pressure of 250 to 350 bar a temperature of 450 to 550 C and a iron are optimal The catalyst ferrite a Fe is produced in the reactor by the reduction of magnetite with hydrogen The catalyst has its highest efficiency at temperatures of about 400 to 500 C Even though the catalyst greatly lowers the activation energy for the cleavage of the triple bond of the nitrogen molecule high temperatures are still required for an appropriate reaction rate At the industrially used reaction temperature of 450 to 550 C an optimum between the decomposition of ammonia into the starting materials and the effectiveness of the catalyst is achieved 54 The formed ammonia is continuously removed from the system The volume fraction of ammonia in the gas mixture is about 20 The inert components especially the noble gases such as argon should not exceed a certain content in order not to reduce the partial pressure of the reactants too much To remove the inert gas components part of the gas is removed and the argon is separated in a gas separation plant The extraction of pure argon from the circulating gas is carried out using the Linde process 55 Large scale implementation Edit Modern ammonia plants produce more than 3000 tons per day in one production line The following diagram shows the set up of a Haber Bosch plant primary reformer air feed secondary reformer CO conversion washing tower ammonia reactor heat exchanger ammonia condenser Depending on its origin the synthesis gas must first be freed from impurities such as hydrogen sulfide or organic sulphur compounds which act as a catalyst poison High concentrations of hydrogen sulfide which occur in synthesis gas from carbonization coke are removed in a wet cleaning stage such as the sulfosolvan process while low concentrations are removed by adsorption on activated carbon 56 Organosulfur compounds are separated by pressure swing adsorption together with carbon dioxide after CO conversion To produce hydrogen by steam reforming methane reacts with water vapor using a nickel oxide alumina catalyst in the primary reformer to form carbon monoxide and hydrogen The energy required for this the enthalpy DH is 206 kJ mol 57 CH 4 g H 2 O g CO g 3 H 2 g D H 206 kJ mol displaystyle ce CH4 g H2O g gt CO g 3H2 g qquad Delta H 206 ce kJ mol The methane gas reacts in the primary reformer only partially To increase the hydrogen yield and keep the content of inert components i e methane as low as possible the remaining methane gas is converted in a second step with oxygen to hydrogen and carbon monoxide in the secondary reformer The secondary reformer is supplied with air as the oxygen source Also the required nitrogen for the subsequent ammonia synthesis is added to the gas mixture 2 CH 4 g O 2 g 2 CO g 4 H 2 g D H 71 kJ mol displaystyle ce 2CH4 g O2 g gt 2CO g 4H2 g qquad Delta H 71 ce kJ mol In the third step the carbon monoxide is oxidized to carbon dioxide which is called CO conversion or water gas shift reaction CO g H 2 O g CO 2 g H 2 g D H 41 kJ mol displaystyle ce CO g H2O g gt CO2 g H2 g qquad Delta H 41 ce kJ mol Carbon monoxide and carbon dioxide would form carbamates with ammonia which would clog as solids pipelines and apparatus within a short time In the following process step the carbon dioxide must therefore be removed from the gas mixture In contrast to carbon monoxide carbon dioxide can easily be removed from the gas mixture by gas scrubbing with triethanolamine The gas mixture then still contains methane and noble gases such as argon which however behave inertly 50 The gas mixture is then compressed to operating pressure by turbo compressors The resulting compression heat is dissipated by heat exchangers it is used to preheat raw gases The actual production of ammonia takes place in the ammonia reactor The first reactors were bursting under high pressure because the atomic hydrogen in the carbonaceous steel partially recombined into methane and produced cracks in the steel Bosch therefore developed tube reactors consisting of a pressure bearing steel tube in which a low carbon iron lining tube was inserted and filled with the catalyst Hydrogen that diffused through the inner steel pipe escaped to the outside via thin holes in the outer steel jacket the so called Bosch holes 52 A disadvantage of the tubular reactors was the relatively high pressure loss which had to be applied again by compression The development of hydrogen resistant chromium molybdenum steels made it possible to construct single walled pipes 58 Modern ammonia reactor with heat exchanger modules The cold gas mixture is preheated to reaction temperature in heat exchangers by the reaction heat and cools in turn the produced ammonia Modern ammonia reactors are designed as multi storey reactors with a low pressure drop in which the catalysts are distributed as fills over about ten storeys one above the other The gas mixture flows through them one after the other from top to bottom Cold gas is injected from the side for cooling A disadvantage of this reactor type is the incomplete conversion of the cold gas mixture in the last catalyst bed 58 Alternatively the reaction mixture between the catalyst layers is cooled using heat exchangers whereby the hydrogen nitrogen mixture is preheated to the reaction temperature Reactors of this type have three catalyst beds In addition to good temperature control this reactor type has the advantage of better conversion of the raw material gases compared to reactors with cold gas injection Uhde has developed and is using an ammonia converter with three radial flow catalyst beds and two internal heat exchangers instead of axial flow catalyst beds This further reduces the pressure drop in the converter 59 The reaction product is continuously removed for maximum yield The gas mixture is cooled to 450 C in a heat exchanger using water freshly supplied gases and other process streams The ammonia also condenses and is separated in a pressure separator Unreacted nitrogen and hydrogen are then compressed back to the process by a circulating gas compressor supplemented with fresh gas and fed to the reactor 58 In a subsequent distillation the product ammonia is purified Mechanism EditElementary steps Edit The mechanism of ammonia synthesis contains the following seven elementary steps transport of the reactants from the gas phase through the boundary layer to the surface of the catalyst pore diffusion to the reaction center adsorption of reactants reaction desorption of product transport of the product through the pore system back to the surface transport of the product into the gas phaseTransport and diffusion the first and last two steps are fast compared to adsorption reaction and desorption because of the shell structure of the catalyst It is known from various investigations that the rate determining step of the ammonia synthesis is the dissociation of nitrogen 3 In contrast exchange reactions between hydrogen and deuterium on the Haber Bosch catalysts still take place at temperatures of 196 C 320 8 F at a measurable rate the exchange between deuterium and hydrogen on the ammonia molecule also takes place at room temperature Since the adsorption of both molecules is rapid it cannot determine the speed of ammonia synthesis 60 In addition to the reaction conditions the adsorption of nitrogen on the catalyst surface depends on the microscopic structure of the catalyst surface Iron has different crystal surfaces whose reactivity is very different The Fe 111 and Fe 211 surfaces have by far the highest activity The explanation for this is that only these surfaces have so called C7 sites these are iron atoms with seven closest neighbours 3 The dissociative adsorption of nitrogen on the surface follows the following scheme where S symbolizes an iron atom on the surface of the catalyst 42 N2 S N2 g species S N2 S a species 2 S N b species surface nitride The adsorption of nitrogen is similar to the chemisorption of carbon monoxide On a Fe 111 surface the adsorption of nitrogen first leads to an adsorbed g species with an adsorption energy of 24 kJmol 1 and an N N stretch vibration of 2100 cm 1 Since the nitrogen is isoelectronic to carbon monoxide it adsorbs in an on end configuration in which the molecule is bound perpendicular to the metal surface at one nitrogen atom 18 61 3 This has been confirmed by photoelectron spectroscopy 62 Ab initio MO calculations have shown that in addition to the s binding of the free electron pair of nitrogen to the metal there is a p binding from the d orbitals of the metal to the p orbitals of nitrogen which strengthens the iron nitrogen bond The nitrogen in the a state is more strongly bound with 31 kJmol 1 The resulting N N bond weakening could be experimentally confirmed by a reduction of the wave numbers of the N N stretching oscillation to 1490 cm 1 61 Further heating of the Fe 111 area covered by a N2 leads to both desorption and the emergence of a new band at 450 cm 1 This represents a metal nitrogen oscillation the b state A comparison with vibration spectra of complex compounds allows the conclusion that the N2 molecule is bound side on with an N atom in contact with a C7 site This structure is called surface nitride The surface nitride is very strongly bound to the surface 62 Hydrogen atoms Hads which are very mobile on the catalyst surface quickly combine with it Infrared spectroscopically detected surface imides NHad surface amides NH2 ad and surface ammoniacates NH3 ad are formed the latter decay under NH3 release desorption 52 The individual molecules were identified or assigned by X ray photoelectron spectroscopy XPS high resolution electron energy loss spectroscopy HREELS and Ir Spectroscopy Drawn reaction schemeOn the basis of these experimental findings the reaction mechanism is believed to involve the following steps see also figure 63 N2 g N2 adsorbed N2 adsorbed 2 N adsorbed H2 g H2 adsorbed H2 adsorbed 2 H adsorbed N adsorbed 3 H adsorbed NH3 adsorbed NH3 adsorbed NH3 g Reaction 5 occurs in three steps forming NH NH2 and then NH3 Experimental evidence points to reaction 2 as being slow rate determining step This is not unexpected since the bond is broken the nitrogen triple bond is the strongest of the bonds that must be broken As with all Haber Bosch catalysts nitrogen dissociation is the rate determining step for ruthenium activated carbon catalysts The active center for ruthenium is a so called B5 site a 5 fold coordinated position on the Ru 0001 surface where two ruthenium atoms form a step edge with three ruthenium atoms on the Ru 0001 surface 64 The number of B5 sites depends on the size and shape of the ruthenium particles the ruthenium precursor and the amount of ruthenium used 47 The reinforcing effect of the basic carrier used in the ruthenium catalyst is similar to the promoter effect of alkali metals used in the iron catalyst 47 Energy diagram Edit Energy diagram An energy diagram can be created based on the Enthalpy of Reaction of the individual steps The energy diagram can be used to compare homogeneous and heterogeneous reactions Due to the high activation energy of the dissociation of nitrogen the homogeneous gas phase reaction is not realizable The catalyst avoids this problem as the energy gain resulting from the binding of nitrogen atoms to the catalyst surface overcompensates for the necessary dissociation energy so that the reaction is finally exothermic Nevertheless the dissociative adsorption of nitrogen remains the rate determining step not because of the activation energy but mainly because of the unfavorable pre exponential factor of the rate constant Although hydrogenation is endothermic this energy can easily be applied by the reaction temperature about 700 K 3 Economic and environmental aspects EditFurther information Ammonia production Sustainable ammonia production External video How Earth s Population Exploded Bloomberg Quicktake Severnside fertilizer plant northwest of Bristol UK When first invented the Haber process competed against another industrial process the cyanamide process However the cyanamide process consumed large amounts of electrical power and was more labor intensive than the Haber process 7 137 143 As of 2018 the Haber process produces 230 million tonnes of anhydrous ammonia per year 65 The ammonia is used mainly as a nitrogen fertilizer as ammonia itself in the form of ammonium nitrate and as urea The Haber process consumes 3 5 of the world s natural gas production around 1 2 of the world s energy supply 6 66 67 68 In combination with advances in breeding herbicides and pesticides these fertilizers have helped to increase the productivity of agricultural land With average crop yields remaining at the 1900 level the crop harvest in the year 2000 would have required nearly four times more land and the cultivated area would have claimed nearly half of all ice free continents rather than under 15 of the total land area that is required today 69 The energy intensity of the process contributes to climate change and other environmental problems such as the leaching of nitrates into groundwater rivers ponds and lakes expanding dead zones in coastal ocean waters resulting from recurrent eutrophication atmospheric deposition of nitrates and ammonia affecting natural ecosystems higher emissions of nitrous oxide N2O now the third most important greenhouse gas following CO2 and CH4 69 The Haber Bosch process is one of the largest contributors to a buildup of reactive nitrogen in the biosphere causing an anthropogenic disruption to the nitrogen cycle 70 Since nitrogen use efficiency is typically less than 50 71 farm runoff from heavy use of fixed industrial nitrogen disrupts biological habitats 6 72 Nearly 50 of the nitrogen found in human tissues originated from the Haber Bosch process 73 Thus the Haber process serves as the detonator of the population explosion enabling the global population to increase from 1 6 billion in 1900 to 7 7 billion by November 2018 74 Reverse fuel cell technology converts renewable energy water and air into ammonia without a separate hydrogen electrolysis process 75 Research EditResearchers designed and synthesized metal organic frameworks MOF that bind and release ammonia at lower pressures and temperatures 175 C The MOF does not bind the other reactants It uses copper atoms linked by cyclohexanedicarboxylate Ammonia converted the MOF into a bundle copper and ammonia containing polymer strands storing large amounts of ammonia Lower temperatures and pressures make possible the use of smaller local facilities 76 Other researchers are working to produce hydrogen using renewable energy and other catalysts that work at lower temperatures and pressures Still others are working on adapting carbon capture and storage technology for use in ammonia plants It is expected that an early use of quantum computing will be modeling that improves the efficiency of the Haber Bosch process 77 by the mid 2020s 78 although some have predicted it will take longer 79 See also EditHydrogen production Industrial gas Other contemporary nitrogen sources Chilean saltpeter Guano Other nitrogen fixation processes Birkeland Eyde process Cyanamide process Paradas methodNotes Edit Hydrogen required for ammonia synthesis is most often produced through gasification of carbon containing material mostly natural gas but other potential carbon sources include coal petroleum peat biomass or waste As of 2012 the global production of ammonia produced from natural gas using the steam reforming process was 72 21 Hydrogen can also be produced from water and electricity using electrolysis at one time most of Europe s ammonia was produced from the Hydro plant at Vemork Other possibilities include biological hydrogen production or photolysis but at present steam reforming of natural gas is the most economical means of mass producing hydrogen References Edit Habers process chemistry India Arihant publications 2018 p 264 ISBN 978 93 131 6303 9 Appl M 1982 The Haber Bosch Process and the Development of Chemical Engineering A Century of Chemical Engineering New York Plenum Press pp 29 54 ISBN 978 0 306 40895 3 a b c d e f g h i j k l m n o p q r Appl Max 2006 Ammonia Ullmann s Encyclopedia of Industrial Chemistry Weinheim Wiley VCH doi 10 1002 14356007 a02 143 pub2 Clark 2013 The forward reaction the production of ammonia is exothermic According to Le Chatelier s Principle this will be favoured at a lower temperature The system will respond by moving the position of equilibrium to counteract this in other words by producing more heat To obtain as much ammonia as possible in the equilibrium mixture as low a temperature as possible is needed Clark 2013 Notice that there are 4 molecules on the left hand side of the equation but only 2 on the right According to Le Chatelier s Principle by increasing the pressure the system will respond by favouring the reaction which produces fewer molecules That will cause the pressure to fall again To get as much ammonia as possible in the equilibrium mixture as high a pressure as possible is needed 200 atmospheres is a high pressure but not amazingly high a b c Smil Vaclav 2004 Enriching the Earth Fritz Haber Carl Bosch and the Transformation of World Food Production 1st ed Cambridge MA MIT ISBN 978 0 262 69313 4 a b c d e Hager Thomas 2008 The Alchemy of Air A Jewish genius a doomed tycoon and the scientific discovery that fed the world but fueled the rise of Hitler 1st ed New York NY Harmony Books ISBN 978 0 307 35178 4 Sittig Marshall 1979 Fertilizer Industry Processes Pollution Control and Energy Conservation Park Ridge NJ Noyes Data Corp ISBN 978 0 8155 0734 5 Vandermeer John 2011 The Ecology of Agroecosystems Jones amp Bartlett Learning p 149 ISBN 978 0 7637 7153 9 James Laylin K 1993 Nobel Laureates in Chemistry 1901 1992 3rd ed Washington DC American Chemical Society p 118 ISBN 978 0 8412 2690 6 Haber Fritz 1905 Thermodynamik technischer Gasreaktionen in German 1st ed Paderborn Salzwasser Verlag ISBN 978 3 86444 842 3 Robert Le Rossignol 1884 1976 Professional Chemist PDF ChemUCL Newsletter p 8 2009 archived from the original PDF on 13 January 2011 Bosch Carl 2 March 1908 Process of producing ammonia U S Patent 990 191 Philip Phylis Morrison 2001 Fertile Minds Book Review of Enriching the Earth Fritz Haber Carl Bosch and the Transformation of World Food Production American Scientist Archived from the original on 2 July 2012 Nobel Award to Haber PDF The New York Times 3 February 1920 Archived from the original PDF on 24 February 2021 Retrieved 11 October 2010 Bozso F Ertl G Grunze M Weiss M 1977 Interaction of nitrogen with iron surfaces I Fe 100 and Fe 111 Journal of Catalysis 49 1 18 41 doi 10 1016 0021 9517 77 90237 8 Imbihl R Behm R J Ertl G Moritz W 1982 The structure of atomic nitrogen adsorbed on Fe 100 PDF Surface Science 123 1 129 140 Bibcode 1982SurSc 123 129I doi 10 1016 0039 6028 82 90135 2 a b Ertl G Lee S B Weiss M 1982 Kinetics of nitrogen adsorption on Fe 111 Surface Science 114 2 3 515 526 Bibcode 1982SurSc 114 515E doi 10 1016 0039 6028 82 90702 6 Ertl G 1983 Primary steps in catalytic synthesis of ammonia Journal of Vacuum Science and Technology A 1 2 1247 1253 Bibcode 1983JVSTA 1 1247E doi 10 1116 1 572299 100 years of thyssenkrupp Uhde Industrial Solutions in German Retrieved 8 December 2021 Ammonia Industrial Efficiency Technology amp Measures 30 April 2013 Retrieved 6 April 2018 Electrochemically produced ammonia could revolutionize food production 9 July 2018 Retrieved 15 December 2018 Ammonia manufacturing consumes 1 to 2 of total global energy and is responsible for approximately 3 of global carbon dioxide emissions Song Yang Hensley Dale Bonnesen Peter Liang Liango Huang Jingsong Baddorf Arthur Tschaplinski Timothy Engle Nancy Wu Zili Cullen David Meyer Harry III Sumpter Bobby Rondinone Adam 2 May 2018 A physical catalyst for the electrolysis of nitrogen to ammonia Science Advances Oak Ridge National Laboratory 4 4 e1700336 Bibcode 2018SciA 4E0336S doi 10 1126 sciadv 1700336 PMC 5922794 PMID 29719860 Retrieved 15 December 2018 Ammonia synthesis consumes 3 to 5 of the world s natural gas making it a significant contributor to greenhouse gas emissions Kuganathan Navaratnarajah Hosono Hideo Shluger Alexander L Sushko Peter V January 2014 Enhanced N2 Dissociation on Ru Loaded Inorganic Electride Journal of the American Chemical Society 136 6 2216 2219 doi 10 1021 ja410925g PMID 24483141 Hara Michikazu Kitano Masaaki Hosono Hideo Sushko Peter V 2017 Ru Loaded C12A7 e Electride as a Catalyst for Ammonia Synthesi ACS Catalysis 7 4 2313 2324 doi 10 1021 acscatal 6b03357 Ajinomoto Co Inc UMI and Tokyo Institute of Technology Professors Establish New Company to implement the World s First On Site Production of Ammonia Ajinomoto 27 April 2017 Retrieved 9 November 2021 Stephen H Crolius 17 December 2020 Tsubame BHB Launches Joint Evaluation with Mitsubishi Chemical Ammonia Energy Association Retrieved 9 November 2021 Kitano Masaaki Kujirai Jun Ogasawara Kiya Matsuishi Satoru Tada Tomofumi Abe Hitoshi Niwa Yasuhiro Hosono Hideo 2019 Low Temperature Synthesis of Perovskite Oxynitride Hydrides as Ammonia Synthesis Catalysts Journal of the American Chemical Society 141 51 20344 20353 doi 10 1021 jacs 9b10726 PMID 31755269 S2CID 208227325 Wang Ying Meyer Thomas J 14 March 2019 A Route to Renewable Energy Triggered by the Haber Bosch Process Chem 5 3 496 497 doi 10 1016 j chempr 2019 02 021 S2CID 134713643 Schneider Stefan Bajohr Siegfried Graf Frank Kolb Thomas 13 January 2020 State of the Art of Hydrogen Production via Pyrolysis of Natural Gas ChemBioEng Reviews 7 5 150 158 doi 10 1002 cben 202000014 S2CID 221708661 via Wiley Online Library Progress in the Electrochemical Synthesis of Ammonia Request PDF Clark 2013 However 400 450 C isn t a low temperature Rate considerations The lower the temperature you use the slower the reaction becomes A manufacturer is trying to produce as much ammonia as possible per day It makes no sense to try to achieve an equilibrium mixture which contains a very high proportion of ammonia if it takes several years for the reaction to reach that equilibrium Clark 2013 Rate considerations Increasing the pressure brings the molecules closer together In this particular instance it will increase their chances of hitting and sticking to the surface of the catalyst where they can react The higher the pressure the better in terms of the rate of a gas reaction Economic considerations Very high pressures are expensive to produce on two counts Extremely strong pipes and containment vessels are needed to withstand the very high pressure That increases capital costs when the plant is built Chemistry of Nitrogen Compounds Chem LibreTexts org 5 June 2019 Retrieved 7 July 2019 Clark 2013 At each pass of the gases through the reactor only about 15 of the nitrogen and hydrogen converts to ammonia This figure also varies from plant to plant By continual recycling of the unreacted nitrogen and hydrogen the overall conversion is about 98 Brown Theodore L LeMay H Eugene Jr Bursten Bruce E 2006 Table 15 2 Chemistry The Central Science 10th ed Upper Saddle River NJ Pearson ISBN 978 0 13 109686 8 Abild pedersen Frank Bligaard Thomas 1 January 2014 Exploring the limits A low pressure low temperature Haber Bosch process Chemical Physics Letters 598 108 Bibcode 2014CPL 598 108V doi 10 1016 j cplett 2014 03 003 via academia edu Koop Fermin 13 January 2023 Green ammonia and fertilizer may finally be in sight and it would be huge ZME Science Retrieved 21 March 2023 Mittasch Alwin 1926 Bemerkungen zur Katalyse Berichte der Deutschen Chemischen Gesellschaft A and B Series 59 13 36 doi 10 1002 cber 19260590103 3 1 Ammonia synthesis resources schoolscience co uk Archived from the original on 6 July 2020 Rock Peter A 19 June 2013 Chemical Thermodynamics p 317 ISBN 978 1 891389 32 0 a b c Max Appl 2006 Ammonia Ullmann s Encyclopedia of Industrial Chemistry Weinheim Wiley VCH doi 10 1002 14356007 a02 143 pub2 Jozwiak W K Kaczmarek E 2007 Reduction behavior of iron oxides in hydrogen and carbon monoxide atmospheres Applied Catalysis A General 326 17 27 doi 10 1016 j apcata 2007 03 021 Ertl Gerhard 1983 Zum Mechanismus der Ammoniak Synthese Nachrichten aus Chemie Technik und Laboratorium 31 3 178 182 doi 10 1002 nadc 19830310307 Bowker Michael 1993 Chapter 7 In King D A Woodruff D P eds The Chemical Physics of Solid Surfaces Vol 6 Coadsorption promoters and poisons Elsevier pp 225 268 ISBN 978 0 444 81468 5 Tavasoli Ahmad Trepanier Mariane Malek Abbaslou Reza M Dalai Ajay K Abatzoglou Nicolas 1 December 2009 Fischer Tropsch synthesis on mono and bimetallic Co and Fe catalysts supported on carbon nanotubes Fuel Processing Technology 90 12 1486 1494 doi 10 1016 j fuproc 2009 07 007 ISSN 0378 3820 a b c d You Zhixiong Inazu Koji Aika Ken ichi Baba Toshihide October 2007 Electronic and structural promotion of barium hexaaluminate as a ruthenium catalyst support for ammonia synthesis Journal of Catalysis 251 2 321 331 doi 10 1016 j jcat 2007 08 006 Rosowski F Hornung A Hinrichsen O Herein D Muhler M April 1997 Ruthenium catalysts for ammonia synthesis at high pressures Preparation characterization and power law kinetics Applied Catalysis A General 151 2 443 460 doi 10 1016 S0926 860X 96 00304 3 Hojlund Nielsen P E 1995 Nielsen Anders ed Poisoning of Ammonia Synthesis Catalysts Ammonia Catalysis and Manufacture Berlin Heidelberg Springer pp 191 198 doi 10 1007 978 3 642 79197 0 5 ISBN 978 3 642 79197 0 retrieved 30 July 2022 a b Falbe Jurgen 1997 Rompp Lexikon Chemie H L Georg Thieme Verlag pp 1644 1646 ISBN 978 3 13 107830 8 Brown Theodore L LeMay H Eugene Bursten Bruce Edward 2003 Brunauer Linda Sue ed Chemistry the Central Science 9th ed Upper Saddle River NJ Prentice Hall ISBN 978 0 13 038168 2 a b c Holleman Arnold Frederik Wiberg Egon 2001 Wiberg Nils ed Inorganic Chemistry translated by Eagleson Mary Brewer William San Diego Berlin Academic Press De Gruyter pp 662 65 ISBN 0 12 352651 5 Cornils Boy Herrmann Wolfgang A Muhler M Wong C 2007 Catalysis from A to Z A Concise Encyclopedia Verlag Wiley VCH p 31 ISBN 978 3 527 31438 6 Fokus Chemie Oberstufe Einfuhrungsphase Berlin Cornelsen Verlag 2010 p 79 ISBN 978 3 06 013953 8 P Haussinger u a Noble Gases In Ullmann s Encyclopedia of Industrial Chemistry Wiley VCH Weinheim 2006 doi 10 1002 14356007 a17 485 Leibnitz E Koch H Gotze A 1961 Uber die drucklose Aufbereitung von Braunkohlenkokereigas auf Starkgas nach dem Girbotol Verfahren Journal fur Praktische Chemie 13 3 4 215 236 doi 10 1002 prac 19610130315 Steinborn Dirk 2007 Grundlagen der metallorganischen Komplexkatalyse Wiesbaden Teubner pp 319 321 ISBN 978 3 8351 0088 6 a b c Forst Detlef Kolb Maximillian Rosswag Helmut 1993 Chemie fur Ingenieure Springer Verlag pp 234 238 ISBN 978 3 662 00655 9 Ammoniakkonverter Dungemittelanlagen Industrial Solutions in German Retrieved 8 December 2021 Moore Walter J Hummel Dieter O 1983 Physikalische Chemie Berlin Walter de Gruyter p 604 ISBN 978 3 11 008554 9 a b Lee S B Weiss M 1982 Adsorption of nitrogen on potassium promoted Fe 111 and 100 surfaces Surface Science 114 2 3 527 545 Bibcode 1982SurSc 114 527E doi 10 1016 0039 6028 82 90703 8 a b Ertl Gerhard 2010 Reactions at Solid Surfaces John Wiley amp Sons p 123 ISBN 978 0 470 26101 9 Wennerstrom Hakan Lidin Sven Scientific Background on the Nobel Prize in Chemistry 2007 Chemical Processes on Solid Surfaces PDF Nobel Foundation Retrieved 17 September 2015 Gavnholt Jeppe Schiotz Jakob 2008 Structure and reactivity of ruthenium nanoparticles PDF Physical Review B 77 3 035404 Bibcode 2008PhRvB 77c5404G doi 10 1103 PhysRevB 77 035404 S2CID 49236953 Ammonia annual production capacity globally 2030 Statista Retrieved 7 May 2020 International Energy Outlook 2007 eia gov U S Energy Information Administration Fertilizer statistics Raw material reserves www fertilizer org International Fertilizer Industry Association Archived from the original on 24 April 2008 Smith Barry E September 2002 Structure Nitrogenase reveals its inner secrets Science 297 5587 1654 5 doi 10 1126 science 1076659 PMID 12215632 S2CID 82195088 a b Smil Vaclav 2011 Nitrogen cycle and world food production PDF World Agriculture 2 9 13 Kanter David R Bartolini Fabio Kugelberg Susanna Leip Adrian Oenema Oene Uwizeye Aimable 2 December 2019 Nitrogen pollution policy beyond the farm Nature Food 1 27 32 doi 10 1038 s43016 019 0001 5 ISSN 2662 1355 Oenema O Witzke H P Klimont Z Lesschen J P Velthof G L 2009 Integrated assessment of promising measures to decrease nitrogen losses in agriculture in EU 27 Agriculture Ecosystems and Environment 133 3 4 280 288 doi 10 1016 j agee 2009 04 025 Howarth R W 2008 Coastal nitrogen pollution a review of sources and trends globally and regionally Harmful Algae 8 14 20 doi 10 1016 j hal 2008 08 015 Ritter Steven K 18 August 2008 The Haber Bosch Reaction An Early Chemical Impact On Sustainability Chemical amp Engineering News 86 33 Smil Vaclav 1999 Detonator of the population explosion PDF Nature 400 6743 415 Bibcode 1999Natur 400 415S doi 10 1038 22672 S2CID 4301828 Blain Loz 3 September 2021 Green ammonia The rocky pathway to a new clean fuel New Atlas Retrieved 23 March 2023 Koop Fermin 13 January 2023 Green ammonia and fertilizer may finally be in sight and it would be huge ZME Science Retrieved 23 March 2023 Ruane Jonathan McAfee Andrew Oliver William D 1 January 2022 Quantum Computing for Business Leaders Harvard Business Review ISSN 0017 8012 Retrieved 12 April 2023 Budde Florian Volz Daniel 12 July 2019 Quantum computing and the chemical industry McKinsey www mckinsey com McKinsey and Company Retrieved 12 April 2023 Bourzac Katherine 30 October 2017 Chemistry is quantum computing s killer app cen acs org American Chemical Society Retrieved 12 April 2023 Sources EditClark Jim April 2013 2002 The Haber Process Retrieved 15 December 2018 External links Edit Detonator of the population explosion PDF Vaclav Smil Department of Geography University of Manitoba Macmillan Magazines Ltd Review of Between Genius and Genocide The Tragedy of Fritz Haber Father of Chemical Warfare PDF Daniel Charles The Haber Process Chemguide co uk BASF Fertilizer out of thin air Britannica guide to Nobel Prizes Fritz Haber Haber Process for Ammonia Synthesis Haber Bosch process most important invention of the 20th century according to V Smil Nature 29 July 1999 p 415 by Jurgen Schmidhuber Nobel e Museum Biography of Fritz Haber Uses and Production of Ammonia Retrieved from https en wikipedia org w index php title Haber process amp oldid 1149686498, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.