Sie sind auf Seite 1von 14

Articles A/Aluminum/4.

Production

Page 1 of 14

Aluminum
William B. Frank, Aluminum Company of America, Alcoa Center, Pa. 15069, United States Warren E. Haupin, Aluminum Company of America, Alcoa Center, Pa. 15069, United States Ullmann's Encyclopedia of Industrial Chemistry Copyright 2002 by Wiley-VCH Verlag GmbH & Co. KGaA. All rights reserved. DOI: 10.1002/14356007.a01_459 Article Online Posting Date: June 15, 2000

4. Production
The only method now used industrially to produce primary aluminum is the HallHroult process. Production of aluminum before its development was discussed previously in this article. Alternate means of producing aluminum are treated in Section Alternate Processes. 4.1. History of the Electrolytic Reduction of Alumina The technological elements of the process electrolysis of fused salts to produce metals (including aluminum), the use of cryolite as a flux to dissolve alumina, and the use of carbon electrodes had been exploited for some time prior to 1886. The workable electrolytic process was discovered independently, and almost simultaneously, in early 1886 by CHARLES MARTIN HALL in Oberlin, Ohio, and PAUL L. T. HROULT in Gentilly, France. Both these young scientists were familiar with the work of SAINTE-CLAIRE DEVILLE. In less than three years, the invention had been implemented industrially in North America and in Europe. In November 1888 aluminum was first produced commercially by the electrolytic reduction of alumina by HALL and others in a company that later was to become the Aluminum Company of America. At about the same time, HROULT was associated with a company (later to be known as Alusuisse) that operated aluminum electrowinning cells at Neuhausen, Switzerland.
KARL J. BAYER,

an Austrian chemist, was issued a patent, DE 43977, in July 1887 for an improved method of producing alumina from bauxite. Bauxite [1318-16-7], discovered by P. BERTHIER in 1821, is named for Les Baux, the village in the south of France near which it was first found. With the development of a process to produce pure aluminum oxide from this abundant ore, the technology was then complete to spur rapid growth of the aluminum industry in Europe and North America in the last decade of the 19th century. The increase in the production of primary aluminum during the past hundred years is shown in Figure 2 [19].
Figure 2. Annual world production of primary aluminum 1854 1982 [7] [Full View]

4.2. Raw Materials Carbon. In the industrial electrowinning (separation by electrolysis) of aluminum, part of the energy for reducing alumina is supplied as electricity and part comes from consumption of the carbon anode. Carbon is also used as the cathode lining. Because 0.4 0.5 kg of anode is consumed for each kilogram of aluminum produced, this represents the major carbon requirement. Because the ash from the carbon will contaminate either the aluminum produced or the electrolyte, high-purity carbon is desirable. Certain impurities, such as vanadium, are particularly

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 2 of 14

harmful in that they catalyze air burning of the carbon. Other impurities, such as phosphorus, accumulate in the electrolyte and undergo cyclic redox reactions (partial reduction followed by reoxidation), consuming electric current without producing product. The coke residue from petroleum refining is quite pure and, therefore, has been the major source of carbon for anodes. The structure of petroleum coke ( Petroleum Coke) varies depending on the nature of the petroleum feedstocks used at the refinery, the refinery flowstream, and the coking conditions used. This coke produced at about 500 C requires calcining at about 1200 C to remove volatile constituents and increase its density before it is blended into the anode mix. After calcination, the coke is ground and mixed with crushed spent anodes and sufficient coal-tar pitch to allow molding into anode blocks by pressing or by vibrating. They are baked at 1000 1200 C, causing the pitch to carbonize, forming strong carbon blocks. These blocks are made with one or more sockets into each of which is fastened a steel stub by pouring cast iron around it. These stubs both conduct electric current into the anode and support the anodes in the cell. The cost of prebaked carbon anodes in the United States was about $ 0.40/kg ($ 0.15/ pound) in 1983. Anthracite has been the major constituent in the cell cathode blocks, although graphite and metallurgical coke have been used to some extent. The anthracite is calcined at 1200 C or higher, crushed and sized, mixed with coal-tar pitch, molded into blocks, and baked. These blocks, mortared together with a carbonaceous seam mix, form the pot lining, which is the container for both the aluminum and the electrolyte. High purity is not as important for the cathode blocks because leaching of impurities is very slow. Consumption of cathode carbon amounts to 0.02 0.04 kg of carbon per kilogram of aluminum produced. The life of a pot (typically 2 6 years) generally is terminated by failure of the carbon pot lining. Aluminum Oxide. Depending on its purity and losses in handling, 1.90 1.95 kg of alumina are consumed in producing 1 kg of aluminum. The cost of alumina at United States smelting facilities was $ 0.20 0.26/kg ($ 0.09 0.12/ pound) in 1983. The preparation of metallurgical alumina and its required physical and chemical properties are discussed under Aluminum Oxide. Electrolyte Materials. The electrolyte for electrowinning aluminum is basically a solution of aluminum oxide in cryolite [15096-52-3]. The presence of cryolite is essential for dissolution of alumina. Cryolite usually comprises more than 75 % of the electrolyte which typically also contains calcium fluoride (4 8 %), excess aluminum fluoride (5 15 %), alumina (1 6 %), and sometimes lithium fluoride [7789-24-4] (0 5 %) and magnesium fluoride [7783-40-6] (0 5 %). These additives lower operating temperature and increase current efficiency. The mineral cryolite is the double fluoride of sodium and aluminum and has a stoichiometry very near the formula Na3AlF6 and a melting point of about 1010 C. It has been found in substantial quantities only in Greenland, and was mined extensively there in the early 20th century but now is essentially exhausted. Synthetic cryolite can be produced by reacting hydrofluoric acid with an alkaline sodium aluminate solution:

Cryolite also can be recovered from used pot linings. The lining is crushed and treated with dilute sodium hydroxide solution to dissolve fluorides. After being filtered, the solution is neutralized with carbon dioxide to precipitate the cryolite. Cryolite is produced directly in reduction cells by reaction of the soda impurity in the feed alumina with added aluminum fluoride ( Fluorine Compounds, Inorganic):

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 3 of 14

(1)

Electrolyte generated by the above reaction must be tapped from the cells periodically. In modern smelters with dry scrubbing equipment for fume treatment and cell lives greater than 3 years, cryolite is a byproduct rather than a raw material in producing aluminum. Synthetic cryolite could be purchased in the United States for $ 500 600/t in 1983. Aluminum Fluoride. Aluminum fluoride, AlF3 [7784-18-1], may comprise as much as 15 wt % of the electrolyte in excess of the amount represented by the cryolite composition. Aluminum fluoride is consumed during normal operation by three major mechanisms. First, losses of aluminum fluoride by vaporization are appreciable; the most volatile species present in the electrolyte is sodium tetrafluoroaluminate [13821-15-3], NaAlF4, having a partial pressure of 200 600 Pa over the operating melt, depending on composition and temperature ( Fluorine Compounds, Inorganic). Second, aluminum fluoride is depleted by hydrolysis:

And finally, aluminum fluoride is consumed by reaction with the soda present in feed alumina (Eq. 1). Fume capture and scrubbing efficiencies have improved aluminum smelters. Fluoride previously lost by vaporization of NaAlF4 and hydrolysis of bath is now almost completely recycled to the cells. Nevertheless, aluminum fluoride consumption amounts to 0.02 0.04 kg AlF3 per kilogram of aluminum product. In 1983 technical anhydrous aluminum fluoride could be purchased in the United States for about $ 350/t. 4.3. HallHroult Cell for Aluminum Production All commercial production of aluminum today is done in HallHroult cells. Employing prebaked carbon anodes or self-baking Sderberg anodes. The HallHroult cell with prebaked anodes is shown in Figure 3. Essentially pure alumina is fed into the previously discussed cryolite base electrolyte. Electric current deposits aluminum into a pool of molten aluminum held under the electrolyte in the carbon lined cavity of the cell. Oxygen from the alumina deposits electrolytically onto the carbon anode dipping into the electrolyte and reacts with (burns) the anode. Cells typically range from 9 to 12 m long, 3 to 4 m wide, and 1 to 1.2 m high. Thermal insulation surrounds the carbon lining of the cell to control heat losses. Although carbon is the material known to withstand best the combined corrosive action of molten fluorides and molten aluminum, even carbon would have a very limited life in contact with the electrolyte at the sides of the cell were it not protected by a layer of frozen electrolyte. The thermal insulation is adjusted carefully to maintain a protective coating on the walls but not on the bottom, which must remain substantially bare for electrical contact. Steel collector bars in the carbon cathode conduct electric current from the cell. These bars are inserted into holes that have been sized carefully so that thermal expansion forms a tight electrical contact, or cemented in place with a carbonaceous cement containing metal particles, or bonded in place with cast iron.
Figure 3. HallHroult cell with prebaked anodes

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 4 of 14

a) Carbon anode; b) Electrolyte; c) Insulation; d) Carbon lining; e) Current collector bar; f) Thermal insulation; g) Steel shell; h) Carbon block; i) Ledge; j) Crust; k) Alumina cover; l) Removable covers; m) Anode rods; n) Fume collection; o) Air cylinder; p) Feeder; q) Current supply; r) Crust breaker [Full View]

The electrical resistivity of prebaked anodes ranges from 0.005 to 0.006 cm. Current density at the anode face is 0.6 1.3 A/cm2. The Sderberg anode (Fig. 4) uses a premixed paste of petroleum coke and coal-tar pitch. This mixture is added at the top of a rectangular steel casing that is typically 6 8 m long, 2 m wide, and 1 m high. Heat from the electrolyte and heat from the electric current passing through the anode bakes the carbonaceous mix as it progresses through the casing.
Figure 4. Aluminum electrolyzing cell with Sderberg anode a) Manifold gas; b) Steel shell; c) Current collector bars; d) Frozen ledge; e) Molten electrolyte; f) Coke and tar paste; g) Current supplying pins [Full View]

The baked portion extends past the casing and into the molten electrolyte. Baked mix replaces anode being consumed at the bottom surface. Electric current enters the anode through either vertical or sloping steel spikes. These spikes are pulled and reset to a higher level as they approach the lower surface. Sderberg anodes have an electrical resistivity about 30 % higher than that of prebaked anodes. Because of the resulting lower power efficiency and the greater difficulty in collecting and disposing of baking fumes, Sderberg anodes are being replaced by prebaked anodes, even though the former save the capital, labor, and energy required to manufacture the latter. 4.3.1. Electrolyte Pure cryolite melts at about 1010 C. Alumina and other additives lower the melting point, allowing operation at 940 980 C. The cryolite-aluminum fluoride-alumina system (Fig. 5) [20] has binary eutectic points at 961 C and 694 C and a ternary one at 684 C. Calcium fluoride and lithium fluoride further reduce the liquidus temperature (Figs. 6 [21] and 7 [22]). Calcium fluoride is seldom added intentionally. Because of a small amount of calcium oxide impurity in the alumina, it attains a steady-state concentration of 3 8 % in the melt. At this level calcium is codeposited into the aluminum and emitted in the off-gas at a rate equal to its introduction. Magnesium fluoride accumulates to 0.1 0.3 %, in the electrolyte by the same mechanism as calcium fluoride. Some operators add up to 5 % MgF2 because it expels carbon dust from the electrolyte by decreasing the electrolyte's ability to wet carbon.
Figure 5. The Na3AlF6AlF3Al2O3 system [20] [Full View]

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 5 of 14

Figure 6. The Na3AlF6Al2O3CaF2 system [21] [Full View]

Figure 7. The Na3AlF6Li3AlF6Al2O3 system [22] [Full View]

Calcium fluoride, in addition to lowering the liquidus temperature, decreases the vapor pressure and solubility of reduced species in the electrolyte for better current efficiency. Detrimentally, it lowers alumina solubility and electrical conductivity and increases density, viscosity, and surface tension of the electrolyte. Lithium fluoride, in addition to lowering the melting point, decreases the vapor pressure, density, reduced species solubility, and viscosity; it also increases electrical conductivity. The only negative effect appears to be lowered alumina solubility. Its high cost, however, requires that its benefits be weighed against the price. Aluminum fluoride decreases solubility of reduced species and lowers surface tension, viscosity, and density. It has the undesirable effects of decreasing alumina solubility and electrical conductivity and increasing vapor pressure. Aluminum fluoride acts as a Lewis acid with sodium fluoride acting as a Lewis base. Neutrality has been defined arbitrarily as a molar ratio of sodium fluoride to aluminum fluoride of 3 : 1. Control of electrolyte acidity or the NaF : AlF3 molar ratio, referred to as the cryolite ratio (Rc), is of importance to cell operation. Lithium fluoride is a slightly weaker Lewis base than sodium fluoride. Magnesium fluoride and calcium fluoride are weak Lewis acids. Ionic Structure of the Melt. There is general agreement that molten cryolite is completely ionized to sodium ions and hexafluoroaluminate ions:

Also it is well established that the hexafluoroaluminate ion dissociates further:

At Rc = 3, [AlF6]3 is about 30 % dissociated [23]. Raman spectroscopy has showed that the dissociation increases with decreasing cryolite ratio to complete dissociation at Rc = 1 [24]. The nature of the species formed when alumina is added is not so well established. Combination of the results of cryoscopic measurements, Raman spectrographic data, and equilibrium studies

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 6 of 14

and vapor pressure measurements (reviewed in [23]) leads to the conclusion that [Al2O2F4]2 and [Al2OF6]2 are the two major oxygen-containing ions in the melt. Possible reactions for their formation are:

4.3.2. Electrode Reactions Cathode Reaction. Even though Na+ is the principal current carrier, it does not discharge at the cathode. The reversible electromotive force for the formation of liquid aluminum is about 0.24 V lower than that for the formation of sodium gas at 101.3 kPa (1 atm) for the range of compositions and temperatures used industrially. BOWMAN [25], using cyclic voltammetry, stationary electrode polarography, differential pulse polarography, and chronopotentiometry, found evidence only for a reversible three-electron transfer process. There was no evidence for a chemical reaction, either preceding or following the electron-transfer process. This ruled out discharge of sodium at low activity followed by a chemical reaction to form aluminum, and also eliminated dissociation of the [AlF6]3 or [AlF4] to form Al3+ and F ions in the double layer preceding charge transfer. ROLIN, et al. [26] believed this latter process does take place. If so, dissociation must be too rapid to be detected by BOWMAN's techniques. The cathode overvoltage can be represented by [27]: (2) where R is the gas constant, T is the temperature (K), Rc is the mole ratio NaF : AlF3 , F is the Faraday constant, and i is the electrode's current density (in A/cm2). Although this relationship mathematically looks like activation overvoltage, actually the cathode overvoltage is caused by an increase in the NaF : AlF3 ratio at the aluminum surface [28]. Sodium ions carry the current while complex aluminum anions discharge. This requires a diffusional flux of [AlF6]3 and [AlF4] ions to the cathode interface with a similar diffusional flux of Na+ and F ions away from the interface. Using Fick's first law, the resulting concentration gradients were calculated and good agreement was found between the measured overvoltages and the electromotive forces between the two aluminum half-cells, one containing electrolyte of the bulk composition and the other, electrolyte corresponding to the calculated interfacial composition. Anode Reactions. The primary anode reaction can be written:

However, O2 ion is not present in the bulk electrolyte; instead, oxygen is present as structurally large complexes, i.e., [Al2OF6]2 and [Al2O2F4]2. Thermodynamically, oxygen depositing onto

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 7 of 14

carbon at 950 1000 C should equilibrate to about 99 % CO and 1 % CO2. However, based on either net carbon consumption or the volume of gas produced, the primary anode product is essentially all CO2. The high anodic overvoltage implies that reaction kinetics cause this surprising displacement from thermodynamic equilibrium. Rotating disk [29] and impedance [30] measurements indicate that there is a small diffusional overvoltage, probably caused by reaction within the pores of the electrode. Using the general treatment for heterogeneous reaction control [31], overvoltage data can be expressed by the relationship: (3) where v is the number of executions of rate-controlling steps to produce one overall step, p is the reaction order, n is the number of electrons transferred in one overall step, and i0 is the reaction limiting current density. The reaction order, p = 0.57, was found from measurements of overvoltage versus current density [32]. In industrial practice, the reaction order ranges from 0.4 to 0.6, varying with carbon reactivity and porosity. The reaction-limiting current density, i0, goes from 0.0039 to 0.0085 A/cm2 as alumina concentration varies from 2 to 8 wt %. The reaction order also has been determined [32] from the rate of change of the limiting current with reactive species concentration and a similar value obtained: (4)

Measurements [33] of the ordinary combustion of graphite have shown that when oxygen reacts both in pores and on the surface, a chemical reaction of approximately half-order results. Linear sweep voltammograms showed voltage peaks with increasing current density corresponding to discharge of CO2 , COF2 , and CF4 [34]. The following anode reaction mechanism is consistent with these observations. Oxyfluoride ions dissociate within the double layer to oxygen ions:

Oxygen ions discharge upon the carbon surface (surf) forming an activated complex, CO:

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 8 of 14

The activated complex converts very slowly to CO (g) through an adsorbed (ads) intermediate:

Carbon burning in pure oxygen at 940 970 C has a combustion rate equivalent to between 0.1 and 0.2 mA/cm2. One would expect this reaction to proceed at a similarly slow rate. As available surface sites become covered with C*O, oxygen is deposed at higher energy (overvoltage) onto a carbon site already bonded to oxygen, producing unstable [C*O2], which breaks carbon to carbon bonds almost immediately as it is formed:

The adsorbed CO2 quickly desorbs as CO2 (g). This mechanism explains both the high anodic overvoltage and the primary production of CO2 instead of the thermodynamically favored CO. When there is insufficient alumina in the electrolyte, the cell experiences a phenomenon called anode effect. Bubbles grow larger and larger on the anode until the electrolyte no longer wets the anode. With a constant potential the current falls to a low value; but with the constant current source used industrially, the cell potential rises to 30 V. Current then penetrates the gas film by a multitude of small electric arcs or sparks. The gas produced at the anode changes from CO2 to CO, with significant quantities (3 25 %) of CF4 and minor amounts of C2F6 generated. Fluorocarbon compounds most likely deposit on the surface of the anode and trigger the anode effect. As alumina is depleted, anode overvoltage increases. At about 1.2 V anode overpotential, sufficient thermodynamic activity of fluorine is produced to cause fluorine to bond to the carbon. Even though these low-surface-energy carbon-fluorine compounds decompose on the surface to CF4 and C2F6 at cell temperature, their rate of formation can exceed the rate of thermal decomposition, producing high coverage. Once the cell is on anode effect, restoring the alumina concentration is not sufficient to return it to normal operation. The gas film must be broken by splashing aluminum, by interrupting the current momentarily, or by lowering the anodes to expose new areas not contaminated with fluorine. 4.3.3. Current Efficiency According to Faraday's law, 1 kA h of electric current should produce 0.3356 kg of aluminum, but only 85 95 % of this amount is obtained. The principal loss mechanism is recombination of anodic and cathodic products. Reduced species go into solution in the electrolyte at the aluminum-electrolyte interface (Fig. 8). Disagreement exists over whether the dissolved species is metallic sodium, monovalent aluminum, subvalent sodium, colloidal aluminum, or a combination

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 9 of 14

of these species. Equation (4) produces a thermodynamic activity of sodium in the melt, whereas Equation (5) produces an activity of aluminum monofluoride. (4)

(5)

Figure 8. Current efficiency loss by metal reoxidation [Full View]

Metal going into solution must first diffuse through the metal-electrolyte boundary layer (Fig. 8). The metal is then transported by convection to the vicinity of the anode. Here it reacts with carbon dioxide. Chemical reactions appear to be fast compared to mass transport. The rate-controlling step was previously assumed to be diffusion of dissolved metal through the boundary layer at the metal-electrolyte interface. However, ref. [35] indicates mixed control with diffusion at both the aluminum-electrolyte interface and the bubble-electrolyte interface being important. There are several mechanisms accounting for additional minor losses in current efficiency. New cell linings absorb sodium, with Equation (4) maintaining an equilibrium activity of sodium. Fortunately, the lining saturates early in the cell's life but until this occurs, current efficiency is low. When a metal dissolves in a molten salt it usually imparts electronic conductivity to the melt, thereby lowering current efficiency. Some investigators have found a small electronic conductance for cryolite-base melts but others have not. Such elements as phosphorus and vanadium, which can be reduced partially at the cathode and then reoxidized at the anode, lower efficiency. 4.3.4. Cell Voltage The voltage of a HallHroult cell is made up of a number of components:

where E0 is the thermodynamic equilibrium voltage described under Thermodynamic Considerations (Section Thermodynamic Considerations); CA is the concentration overpotential at the anode; SA is the surface overpotential at the anode described by Equation (3); CC is the concentration overpotential at the cathode described by Equation (2); SC is the surface overpotential at the cathode, generally negligible; I is the total cell current; RA is the electrical resistance of the anodes or anode; RB is the effective resistance of the bath, allowing for fanning out of current as it flows from anode to cathode and the increased bath resistivity caused by gas bubbles (for details, see 27); RC is the cathode resistance; and RX is the resistance external to the cell but included in calculating power consumption. The anode concentration overpotential can be calculated by:

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 10 of 14

(5) The quantity ic is the concentration-limited current density and it varies between 1.00.5 A/cm2 at 1 wt % Al2O3 and 15 A/cm2 at 8 wt % Al2O3. The specific power consumption in kilowatt-hours per kilogram of aluminum can be calculated:

where % CE is the percentage current efficiency or Faraday efficiency and Ecell is the cell voltage. 4.3.5. Heat Balance Heat balance considerations are important in designing and operating a cell at the optimum temperature. Calculation is complicated by simultaneous heat flow and heat generation in conductors that carry electric current into and out of the cell. More than 50 % of the heat loss may occur through the anodes and top crust. The sidewalls must be protected from erosion by a ledge of frozen electrolyte maintained by extracting the exact amount of heat required for the desired ledge thickness. These calculations have been described [36], [37]. The complex geometry and the interaction between heat flow and electrical flow generally require a computer using either a finite element or finite difference technique for heat balances. Figure 9 [38] shows a typical heat loss distribution for a center-fed cell and Figure 10 [38], the typical temperature distribution.
Figure 9. Cell heat loss distribution [38] [Full View]

Figure 10. Cell lining isotherms [38] Dashed line: measured profile [Full View]

4.3.6. Fluid Dynamics The large electric currents used in modern cells generate relatively strong magnetic fields within the cell. These magnetic fields interact with electric currents and exert Lorentz forces, producing movement of liquid conductors. Three types of magnetic disturbances have been observed: a vertical or static displacement of the metal pad, a circulating flow within the metal pad with velocities as high as 20 cm/s, and finally a wave motion in the metal pad. The last is probably the most harmful because waves may lead to electrical short circuiting. Calculation of the magnetic fields and electric current flow patterns is complicated. Computer programs have been designed to make these calculations and describe the fluid-dynamic consequences. These calculations have

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 11 of 14

been refined to the point where cells of over 2.5105 A have been designed and operated with high efficiency. Natural convection caused by temperature gradients or composition differences in the electrolyte are insignificant compared with the movement induced by magnetic forces and gas bubbles. Gas bubbles produce significant stirring in the electrolyte and are the dominant force in bath movement, whereas electromagnetic forces predominate in metal pad movement. 4.4. Thermodynamic Considerations The thermochemistry of constituents of the electrolyte and of the electrolyte-aluminum system is treated in [23]. Thermodynamic data used in the following section are from the JANAF (Joint Army-Navy-Air Force) Thermochemical Tables [39] and the 1978 supplement to the tables [40]. For pure -alumina, reduction to aluminum can be represented:

In the industrial production of aluminum, a portion of the energy is supplied by the combustion of carbon anodes to carbon dioxide. The overall cell reaction can be represented: (6)

Additional energy is required to heat alumina and carbon from room temperature to operating temperature. The alumina fed to the cell usually is not pure -alumina. For -alumina, the overall process can be represented:

The enthalpy requirement for transformation of the reactants at 25 C (298 K) to products at 960 C (1233 K) is:

This value corresponds to a theoretical energy requirement of 6.25 kW h per kilogram of aluminum produced. The reversible cell potential (decomposition voltage) for Reaction (6) can be calculated from the relation:

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 12 of 14

(7)

where is the Gibbs free energy change for Equation (6), 3.450105 J; n is the number of electrons per unit cell reaction, i.e., 3; F is the Faraday constant, 96.487 kJ V1 mol1; and is the standard electrode potential, in volts, at 960 C for the reaction with all reactants and products at unit activity. Substituting the appropriate values into Equation (7) and solving for the cell potential:

The above decomposition potential applies to electrolyte saturated with alumina at 950 C. The decomposition potential from melts with alumina activity less than unity can be calculated [27] from the equation:

The activity of alumina in cryolite-base melts can be approximated [27] from the equation:

where is the concentration of alumina in the electrolyte, in wt %, and saturation concentration of alumina, in wt %.

is the

4.5. Alternate Processes Although the HallHroult process has gained industrial dominance, it has several inherent disadvantages. The most serious are the large capital investment required and the high consumption of costly electrical power. There are also the costs of the Bayer alumina refining plant and of the carbon anode plant. Many of the aluminum-producing countries must import alumina or bauxite. The supply of petroleum coke is limited. These deficiencies have spurred research to find alternate processes. The earliest commercial process for producing aluminum was sodiothermic reduction of aluminum halides. Reduction of aluminum chloride by manganese has been investigated [41]. However, neither can compete with the HallHroult process. Many attempts have been made at direct carbothermic reduction of alumina butthese have resulted in very low yields, owing to the formation of solid aluminum carbide, aluminum suboxide vapor, and aluminum vapor that reacts with carbon monoxide as the temperature drops on the way from the furnace. Yields as high as 67 % can be obtained by staging the reactions to produce aluminum carbide at 1930 2030 C, which then reacts with alumina to produce aluminum and carbon monoxide at 2030 2130 C [42]. Better yields result from adding to the furnace a metal (or a metal oxide that is subsequently reduced to a metal), such as iron, silicon, or copper, to alloy with the aluminum and lower its vapor pressure. Of course it is then necessary to extract the aluminum from the alloy. In principle this can be accomplished by electrolytic refining, by fractional crystallization, or by monohalide

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 13 of 14

distillation. In the latter process the aluminum extraction takes place at 1000 1400 C:

The AlCl gas is transported to a cooler zone, 600 800 C, where pure aluminum is formed:

These combined processes to date have proved noncompetitive. Selective solution of aluminum from the alloy by using a volatile metal, such as mercury, lead, bismuth, cadmium, magnesium, or zinc, has been investigated. Following extraction, the volatile metal is distilled, leaving pure aluminum. If FeAl3 , TiAl3 , or Al4C3 is formed, neither electrolysis nor volatile metal extraction will remove the aluminum from the compound. The aluminum vapor pressure can be lowered also by alloying the aluminum with aluminum carbide. Over 40 wt % aluminum carbide is soluble in aluminum at 2200 C [43]. This observation led to a process in which an alloy of aluminum and aluminum carbide was produced at 2400 C. When this alloy was tapped from the furnace and allowed to cool slowly, the aluminum carbide crystallized into an open lattice, the interstices of which were filled with pure aluminum. Pure aluminum was then removed either by leaching with molten chloride fluxes or by vacuum distillation. The aluminum carbide residue was recycled to the arc furnace. Alternately, the aluminum carbide could be distilled destructively above 2200 C to produce aluminum and a residue of pure graphite. In a joint project with the U.S. Department of Energy concluded in 1983, Alcoa investigated producing aluminum-silicon alloy carbothermally in a blast furnace. The required high temperature was obtained by using a pure oxygen blast. Using a low-pressure blast furnace to produce an aluminum-silicon alloy was found to be infeasible because of severe bridging and interruption of the burden movement caused by total reflux of Al, Al2O, and SiO vapors [44], [45]. Addition of iron improved operation but required an improved technique for extracting the aluminum from the dilute aluminum alloy. An alloy having higher aluminum content could be obtained if a significant portion of the process energy was supplied by electrical power [46]. Japanese researchers are continuing to explore the aluminum blast furnace concept aiming for a low-silicon, high-iron alloy [47]. This improves the blast furnace efficiency but complicates extraction of aluminum from the alloy. They propose to overcome this difficulty by extracting the aluminum with lead [48]. In 1976 Alcoa described a smelting process wherein aluminum chloride, dissolved in a molten sodium chloride-lithium chloride electrolyte, was electrolyzed in a bipolar electrode cell to produce aluminum and chlorine. The chlorine was recycled to a fluid-bed chemical reactor, where it reacted with alumina, pyrolytically coated with carbon from fuel oil, to produce aluminum chloride, carbon dioxide, and carbon monoxide. This reaction was highly exothermic. The aluminum chloride was desublimed to separate it from the gas and recycled to the cell to produce more aluminum and chlorine. This process required 30 % less electrical power than their best HallHroult cells. Because the entire process is a closed system, it is also environmentally more attractive than the HallHroult process. Problems with the chemical plant and low demand for aluminum caused the pilot plant to be shut down temporarily in 1982. Research on the plant was continuing in 1984.

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Articles A/Aluminum/4. Production

Page 14 of 14

[Top of page] [A to Z] [Authors] [Subjects] [Search] [HitList]


2004 by Wiley-VCH Verlag GmbH & Co. KGaA All rights reserved.

http://127.0.0.1:49152/Wiley/lpext.dll/Ullmann-1/a01_459/sect4.html

23/12/2010

Das könnte Ihnen auch gefallen