Sei sulla pagina 1di 20

Aluminium Casting Techniques - Sand Casting and Die Casting Processes

Aluminium casting processes are classified as Ingot casting or Mould casting. During the first process, primary or secondary aluminium is cast into rolling ingot (slab), extrusion ingot (billet) and wire bar ingot which are subsequently transformed in semi- and finished products. The second process is used in the foundries for producing cast products. This is the oldest and simplest (in theory but not in practice) means of manufacturing shaped components. This section describes exclusively Mould casting which can be divided into two main groups : Sand casting Die casting

Other techniques such as "lost foam" or "wax pattern" processes are also used but their economical importance is considerably lower than both listed techniques.

Sand Casting
In sand casting, re-usable, permanent patterns are used to make the sand moulds. The preparation and the bonding of this sand mould are the critical step and very often are the rate-controlling step of this process. Two main routes are used for bonding the sand moulds: The "green sand" consists of mixtures of sand, clay and moisture. The "dry sand" consists of sand and synthetic binders cured thermally or chemically.

The sand cores used for forming the inside shape of hollow parts of the casting are made using dry sand components. This versatile technique is generally used for high-volume production. An example of half sand mould is given in Figure 1.

Figure 1. Half mould with cores and an example of a cast air intake for a turbocharger. Normally, such moulds are filled by pouring the melted metal in the filling system. Mould designing is a particularly complex art and is based on the same principle as gravity die casting illustrated in Figure 4. 1

In the "low pressure" sand casting technique, the melted metal is forced to enter the mould by low pressure difference. This more complicated process allows the production of cast products with thinner wall thickness.

Die Casting
In this technique, the mould is generally not destroyed at each cast but is permanent, being made of a metal such as cast iron or steel. There are a number of die casting processes, as summarised in Figure 2. High pressure die casting is the most widely used, representing about 50% of all light alloy casting production. Low pressure die casting currently accounts for about 20% of production and its use is increasing. Gravity die casting accounts for the rest, with the exception of a small but growing contribution from the recently introduced vacuum die casting and squeeze casting process.

Figure 2. Classifications of die casting processes.

Gravity Casting
A schematic view in Figure 3 shows the main parts constituting a classical mould for gravity die casting. Cores (inner parts of the mould) are generally made of bonded sand.

Figure 3. Schematic view of the components of a casting mould (gravity die casting). 2

Gravity die casting is suitable for mass production and for fully mechanised casting.

High Pressure Die Casting


In this process, the liquid metal is injected at high speed and high pressure into a metal mould. A schematic view of high pressure die casting is given in Figure 4.

Figure 4. Schematic view of a high pressure die casting machine. This equipment consists of two vertical platens on which bolsters are located which hold the die halves. One platen is fixed and the other can move so that the die can be opened and closed. A measured amount of metal is poured into the shot sleeve and then introduced into the mould cavity using a hydraulically-driven piston. Once the metal has solidified, the die is opened and the casting removed. In this process, special precautions must be taken to avoid too many gas inclusions which cause blistering during subsequent heat-treatment or welding of the casting product. Both the machine and its dies are very expensive, and for this reason pressure die casting is economical only for high-volume production.

Low Pressure Die Casting


As schematised in Figure 5, the die is filled from a pressurised crucible below, and pressures of up to 0.7 bar are usual. Low-pressure die casting is especially suited to the production of components that are symmetric about an axis of rotation. Light automotive wheels are normally manufactured by this technique.

Figure 5. Schematic view of a low pressure die casting machine.

Vacuum Die Casting


The principle is the same as low-pressure die casting. The pressure inside the die is decreased by a vacuum pump and the difference of pressure forces the liquid metal to enter the die. This transfer is less turbulent than by other casting techniques so that gas inclusions can be very limited. As a consequence, this new technique is specially aimed to components which can subsequently be heat-treated.

Squeeze Casting or Squeeze Forming


As shown in Figure 6, liquid metal is introduced into an open die, just as in a closed die forging process. The dies are then closed. During the final stages of closure, the liquid is displaced into the further parts of the die. No great fluidity requirements are demanded of the liquid, since the displacements are small. Thus forging alloys, which generally have poor fluidities which normally precludes the casting route, can be cast by this process.

Figure 6. The squeeze casting principle. This technique is especially suited for making fibre-reinforced castings from fibre cake preform. Squeeze casting forces liquid aluminium to infiltrate the preform. In comparison with non4

reinforced aluminium alloy, aluminium alloy matrix composites manufactured by this technique can double the fatigue strength at 300C. Hence, such reinforcements are commonly used at the edges of the piston head of a diesel engine where solicitations are particularly high.

Conclusions
Aluminium castings are very powerful and versatile techniques for manufacturing semi- or finished products with intricate shapes. Those techniques are continuously improved and developed to satisfy the user needs and to penetrate new markets. Innovations are mainly oriented to the automobile sector which is the most important market for castings. This continual improvement and development will ensure that aluminium castings continue to play a vital role in this field. Source: European Aluminium Association. For more information on this source please visit European Aluminium Association.

Riser (casting)
A riser or feeder is a reservoir built into a metal-casting mold to prevent cavities due to shrinkage. Because metals are less dense as liquids than as solids, castings shrink as they cool. This can leave a void, generally at the last point to solidify. Risers prevent this by providing molten metal at the point of likely shrinkage, so that the cavity forms in the riser, not the casting. This only works if the riser cools after the rest of the casting. Chvorinov's rule states that the solidification time t of molten metal is related to the constant C (which depends on the thermal properties of the mold and the material) and the local volume (V) and surface area (A) of the material, according to the relationship

Therefore, to ensure that the casting solidifies before the riser, the ratio of the volume to the surface area of the riser should be greater than that of the casting. Because risers exist only to ensure the integrity of the casting, they are removed after the part has cooled, and their metal scrapped. As a result, riser size, number, and placement should be carefully planned to reduce waste while filling all the shrinkage in the casting.

Bronze casting showing sprue and risers (side view)

Improve profitability by reducing aluminum melting costs.


By zeroing in on the often overlooked areas of energy and melt loss, aluminum foundries can increase their furnace efficiency for bottom line results. As profit margins shrink in response to low-cost global competition and the growing cost of regulatory compliance, cost control is priority one. Today's aluminum foundry produces daily reports on production costs, scrap rates, delivery performance, sand control and disposal, and maintenance. One area often overlooked by foundry managers in their quest for cost reduction, however, is the energy cost in aluminum melting and the melt loss. The total energy cost of melting and the total cost of melt loss are simple figures to determine. Melting cost is the energy used per day (kW) vs. the amount of metal melted (lb). Melt loss is calculated by the metal purchased (lb) minus the castings shipped (lb) (over a month or year to account for the remelt of gates, risers and scrapped castings). By monitoring these costs and controlling them when possible, a significant savings and increased melt performance can be achieved. Energy Costs As foundries upgrade their equipment and increase capacities, they often replace old furnaces with newer energy-efficient models to control melting energy costs and increase efficiency (Table 1). This step alone, however, isn't enough. Regardless of the type or age of melting equipment, foundries must be aware of their furnace melt rates, energy usage and waste to control melt energy costs. Any increases in melt rate or reduction in energy use is a direct profit margin enhancement. The best way to increase the melt rate of a furnace is to make sure that the furnace is operating at full efficiency with minimal heat loss. Figure 1 shows the primary sources of heat loss during melting. It is important to note that only a percentage of the heat generated by the furnace is available heat for melting, and that this percentage can be increased through control measures.

Walls - The heat loss through the walls is a direct result of the refractory lining efficiency, maintenance and installation. Regular repair and maintenance of this refractory will help minimize heat loss. Refractory walls of melting furnaces must be checked regularly for cracks and spalling, and repaired as necessary, and the furnace lining must be replaced as needed. Also, check with your refractory supplier about options for lining the furnace. Newer refractory materials continually are being developed in an attempt to increase refractory life. A change in lining or a new design also can increase efficiency and decrease this heat loss factor. Opening - Whenever the furnace dip well is open, the furnace is losing heat. If the furnace does not have a cover, this heat loss is constant. Reverberatory furnace dip wells and crucible furnaces should be covered to reduce heat loss. In one test, a foundry decreased its melt time by more than 25% by adding a refractory cover to an uncovered crucible furnace during melting. The time a furnace is open during degassing, fluxing and melting also contributes to the heat loss. This heat loss can be minimized by performing these processes under a cover. Flue loss - Although some heat loss through the exhaust flue is inevitable, it can be minimized by keeping the burners properly "tuned" for maximum burner efficiency. The proper fuel-air ratio can produce greater heat (a hotter flame) with less fuel, which allows the furnace to melt more quickly and decrease the melt time. The local gas utility can be a resource bringing burners to maximum efficiency, and many offer incentives and cost-sharing for repairs or new equipment that save 7

energy. Other options are new furnace designs and aftermarket equipment that capture flue loss and reuse it as an additional source of heat for melting. Pyrometer care - Although the pyrometer is not part of the furnace, it is the only means of melt temperature measurement. The pyrometer should be kept clean and calibrated as a mistake of 50F results in a foundry using at least 2% more energy per melt. Burner Control - High energy bills may warrant the consideration of new furnace controls or burners. The newest burners are designed for maximum fuel economy and offer high-efficiency combustion methods, low-flame cycling options and other controls that older equipment may not have. Melt Loss Melt loss is another cost factor that is rarely quantified by foundries, and, as a result, is rarely targeted for reduction. Aluminum foundry melt loss averages 8% per year, with some shops experiencing up to 15%. An opportunity to improve foundry profit margins exists if the sources for those losses are identified and reduced or eliminated (Table 2). Five major areas of melt loss exist in a foundry: dross and furnace skimmings; metal spills and mixed metal; saw kerfs or cuttings; dust collector fines; and spilled sand and core butts that contain metal. These "waste materials" often are disposed of through smelters, metal suppliers and scrap brokers. This "hassle-free" convenience of having the materials hauled away often clouds the basic issue that a large potential source for profit improvement is disappearing out the door with little or no management control. Furnaces vary widely in the amount of melt loss and waste they generate. For reverberatory furnaces, oxide films caused by direct flame impingement on the molten aluminum and charge materials generate high metal loss rates. This oxidation is minimized through proper burner maintenance. Crucible furnaces, meanwhile, should run with an oxidizing flame and with exhaustion of combustion gases away from the molten metal surface. The condition of the charge material also is a source of metal loss. Foundries typically melt 30-lb ingot bars and "returns" of gates, risers and scraped castings. Home returns, however, are frequently contaminated with residual green sand adhering to the surface and lubricants from the cutoff and cleaning operations. In addition, they may contain glass screens or ceramic filters that were utilized in the gating systems. While clean and dry charge materials sustain a minimal loss of material when remelted, wet or contaminated constituents charged into the crucibles or open wells of reverberatory furnaces result in excessive aluminum-rich dross from the reaction of the moisture and organic compounds with molten aluminum. In addition, ingot and returns that have oxidized due to contact with moisture and humidity also react, resulting in dross formed from the oxides. Foundries can combat this excess oxide by keeping all ingot and scrap indoors and dry and by preheating returns. Table 1. Efficiencies of Aluminum Melting Furnaces
Furnace Gas reverberatory Electric reverberatory Stack melter Crucible (gas) Electric resistance Melt Loss 2-12% 1% 1-3% 2-3% less than 2% Efficiency 12-28% 70-75% 40-60% 7-17% 70-90% Btu/lb 1500-5000 820 850-1250 1800-3300 700

Note - This information was collected from various furnace manufacturers.

Title: Improve profitability by reducing aluminum melting costs. Author(s): Groteke, Daniel E.; Robison, Stephen T. Publication: Modern Casting Issue: March, 1999 Related Topics: Aluminum casting (Process) (Production management), Foundries (Production management), Metal castings (Production management) Product References: Aluminum Die Castings, Iron & Steel Foundries SIC: Aluminum die-castings, Iron and Steel Foundries Geographic References: United States

The largest percentage of waste (melt loss) is in the removal of dross. Most aluminum melters generate thousands of pounds of dross each year and dispose of it to the highest bidder. Because most dross lots are small shipments (incurring high freight and handling charges), foundries are typically compensated for dross at only 20% of its true value. Foundries accept this cost penalty because they incorrectly believe they have little control over it.
Table 2. High Cost of Aluminum Melt Loss Based on Metal Usage Melt loss 1% 2% 4% 6% 8% Metal melted (lb) 7215 7289 7441 7599 7764 Metal loss (lb) 72 146 298 456 621 Aluminum cost ($) 0.90 0.90 0.90 0.90 0.90 Daily loss ($) 65 131 268 410 559 Annual loss ($) 16,900 34,060 69,680 106,600 145,340

The amount of metal remaining in the dross is a direct result of fluxing efficiency. When properly performed, the dross should be loose, granular material with minimal metal. The treatment of the dross varies from operation to operation as some perform a fluxing and cleansing process on the melt while others simply skim the untreated melt. This dross results in a 1-2% metal loss for clean and dry charge materials, but can reach up to 10% for contaminated and wet charge materials of light section thickness. If machine turnings are part of the charge, melt losses can exceed 15% because of oxidized material Efficient foundries or diecasting shops typically ship out dross for reprocessing with metal contents up to 50%, while shops doing a poor job of in-house fluxing and treatment may be shipping out dross with metal contents approaching 85%. If the dross contains 70% metal and the alloy cost is $0.90/lb, then a foundry is losing the equivalent of $63 per 100 lb of dross (minus the payment received from the collection/dross processor). Dross recovery units designed to remove the metal from the furnace skimmings offer a means of reducing metal loss for smaller foundries. These units vary in style and method, but are designed to minimize the amount of metal in the final dross that leaves the foundry for recycling.

Simulation reduces overflow costs.


Simulation is saving money in aluminium diecasting by making it possible for engineers to design efficient overflow positions on cast parts, reducing the volume of material that must be removed and recycled. The trial and error design methods historically used to position overflow reservoirs were primarily based on experience and, all too often, guess work. Because these features are critical to the casting quality, it is imperative that they be properly placed before any production quantities are made. The overflow volumes are used to capture oxides and other non-metallic inclusions, which are formed as the molten metal fills the cavity of the die. Ford has been able to reduce overflow volume by using computational fluid dynamics software (CFD) to simulate how the metal flows inside of the die. The key for a successful simulation based design for runner and overflow positioning is the ability of the software to track inclusions throughout the filling process so that overflow pads can be positioned to capture them at the very final points where the metal fills the mould cavity. In addition to the importance of the quality of the casting, the long term savings comes from minimising the amount of metal cast into these overflows. High pressure diecasting has emerged as one of the automobile industry's most economical methods for the production of light weight components. A key advantage of this technology is its ability to cast parts with times of less than a minute. These cycle times are achieved by rapidly forcing the molten metal into the die under tremendous pressure so that mould filling typically occurs in a period ranging from 20 to 500 milliseconds. This tremendous filling speed also provides a problem in many applications. In other types of moulding operations it is possible to perform physical experiments to monitor the filling process and solve problems. In high pressure diecasting, mould filling occurs so quickly and the filling patterns are so complex that physical experiments, such as making 'short shots', provide very limited information.

Title: Simulation reduces overflow costs. (News). Publication: Foundry Trade Journal Issue: March, 2002 Related Topics: Metal castings industry (Technology application), Aluminum industry (Technology application), Computer simulation (Usage) Company References: Flow Science Inc. (Product information) Product References: Aluminum Die Castings, Computer Software SIC: Aluminum die-castings, Prepackaged software

As a result, automotive diecasters have long had difficulty with positioning overflow reservoirs on the first try, requiring the addition of better placed overflows. Overflow reservoirs are required in most aluminium diecasting applications because when the molten metal enters the relatively cold die cavity during the filling cycle, oxides and other non-metallic inclusions are inevitably formed. Also, the air that is resident in the cavity must be vented away as the metal fills the volume. The air and inclusions negatively impact the structural properties of the finished part, so they must be moved from the cavity volume by transfer to the overflow. The universal method for removing air and inclusions is to incorporate pad shaped overflow reservoirs connected to the cavity through thin sections called gates. The idea is that the air and inclusions, typically located on the leading edge of the liquid metal front, will flow into the reservoirs where the air is vented out to the atmosphere and the inclusions are trapped in metal.

10

The difficulty is in accurately locating and sizing these overflow vents. The overflow pad must be large enough to capture all of the inclusions. Yet, every gram of metal that flows into the overflow must be re-melted, filtered and recast in another part. For a diecaster like Ford that produces tens of thousands of diecast parts every day, the expense of recasting overflow material is huge. Yet in the past, Ford engineers had to rely on experience and guesswork to design the overflow reservoirs, so they had to make the reservoirs much larger than necessary. Once the part was in production, the engineers gained information by examining the overflow material. Changes made to the die at this point, however, are expensive because they involve repair costs, downtime and almost always require welding the cavity which reduces its life by an average of 50%. CFD simulation That is why Ford has made a major effort to simulate the diecasting process using CFD software. CFD involves the solution of the governing equations for fluid flow and heat transfer at thousands of discrete points on a computational grid in the flow domain. When properly validated, a CFD analysis allows engineers to look inside the die and determine the exact position of the flow front at any point in time as well as the temperature and pressure of the metal at any point in the die. This is far more information than can be obtained from plant trial and usually allows engineers to understand where problems might arise. The geometry of the model representing the die can be changed quickly on the computer and reanalysed to determine the effect of the change. As a result, engineers can eliminate porosity problems in days as opposed to the weeks or months required with the trial approach. Ford engineers have used this technology to eliminate porosity problems for in house diecasting operations and also for suppliers. Simulations have been conducted with several software packages. One software package particularly well suited to simulating the filing of a die is FLOW3D CFD software from Flow Science, Inc, Santa Fe, New Mexico. FLOW-3D is a general purpose, three dimensional. CFD program that has the ability to predict flows with free surfaces during die filling. It uses the volume of fluid (VOF) method to predict free surface fluid motions, surface tension and other flow complexities. In particular, this package provides algorithms that track sharp liquid interfaces through arbitrary deformations and applies the correct normal and tangential stress boundary conditions -- an accuracy feature that distinguishes it from other CFD programs. Surface defect tracking Accurate simulation of the mould filling process gives engineers the tools to make dramatic process improvements. First of all, viewing the filling process makes it possible to determine whether the molten metal is taking a direct route through the cavity or is instead swirling around the cavity. The disadvantage of swirling is that the re-circulating metal surfaces come into contact multiple times and begin to solidify while flowing, increasing the formation and trapping of inclusions. Another role of simulation in reducing overflow volume is the potential for identifying the location of inclusions at each stage of the filling process so that overflow reservoirs can be positioned and sized to capture the inclusions.

Title: Simulation reduces overflow costs. (News). Publication: Foundry Trade Journal Issue: March, 2002 Related Topics: Metal castings industry (Technology application), Aluminum industry (Technology application), Computer simulation (Usage) Company References: Flow Science Inc. (Product information) Product References: Aluminum Die Castings, Computer Software SIC: Aluminum die-castings, Prepackaged software

11

FLOW 3-D provides a unique feature called surface defect tracking that is particularly useful in this regard. It predicts the formation of oxides and then as each oxide particle is created it is followed until freeze-off. This feature greatly improves the ability of the simulation to optimise reservoir and gate positioning and sizing. Reducing porosity Ford engineers have also used CFD to improve casting quality by reducing porosity. Porosity is formed when a flow front breaks apart and traps a volume of air, or other gases, when the flow front converges back together. Sometimes a single flow front can curls over itself forming a wave eddy. Porosity formed as a result of poor filling can be large enough to result in blisters within the 2-4mm thick wall or provide a leak path when it is machined into. An ideal runner and gate design generates a uniform filling pattern. The metal flow does not form branches, but sweeps uniformly through the part pushing all existing and evolving gases in front of it. This sweeping motion also pushes gas and the formed oxidised layer on the surface of the metal flow front into the overflows at the very end of the filling sequence. FLOW 3-D contains a special model called the adiabatic bubble model that can be used to predict the formation of porosity bubbles. This model takes advantage of the fact that the inertia of air can be neglected because its density is small compared to the molten metal. The model treats each region of air as a bubble whose pressure satisfies a pressure volume relationship. This makes it possible to track air bubbles through the filling process just as the surface defect model tracks inclusions. Engineers can then modify the model by making adjustments to the gating, vents and overflows in order to eliminate the porosity. Ford engineers have modelled a number of new diecast components in order to ensure that porosity will not cause quality problems, thus minimising the need for expensive die modifications.

12

Modification and Refinement of Aluminum-Silicon Alloys


Abstract: Hypoeutectic aluminum-silicon alloys can be improved by inducing structural modification of the normally occurring eutectic. In general, the greatest benefits are achieved in alloys containing from 5% Si to the eutectic concentration; this range includes most common gravity cast compositions. The elimination of large, coarse primary silicon crystals that are harmful in the casting and machining of hypereutectic silicon alloy compositions is a function of primary silicon refinement. Hypoeutectic aluminum-silicon alloys can be improved by inducing structural modification of the normally occurring eutectic. In general, the greatest benefits are achieved in alloys containing from 5% Si to the eutectic concentration; this range includes most common gravity cast compositions.

Chemical Modifier
The addition of certain elements, such as calcium, sodium, strontium, and antimony, to hypoeutectic aluminum-silicon alloys results in a finer lamellar or fibrous eutectic network. It is also understood that increased solidification rates are useful in providing similar structures. There is, however, no agreement on the mechanisms involved. The most popular explanations suggest that modifying additions suppress the growth of silicon crystals within the eutectic, providing a finer distribution of lamellae relative to the growth of the eutectic. The results of modification by strontium, sodium, and calcium are similar. Sodium has been shown to be the superior modifier, followed by strontium and calcium, respectively. Each of these elements is mutually compatible so that combinations of modification additions can be made without adverse effects. Eutectic modification is, however, transient when artificially promoted by additions of these elements. Antimony has been advocated as a permanent means of achieving structural modification. In this case, the modified structure differs; a more acicular refined eutectic is obtained compared to the uniform lacelike dispersed structures of sodium-, calcium-, or strontium-modified metal. As a result, the improvements in castability and mechanical properties offered by this group of elements are not completely achieved. Structural refinement is obtained that is time independent when two conditions are satisfied. First, the metal to be treated must be essentially phosphorus free, and second, the velocity of the solidification front must exceed a minimum value approximately equal to that obtained in conventional permanent mold casting. Antimony is not compatible with other modifying elements. In cases in which antimony and other modifiers are present, coarse antimony-containing intermetallics are formed that preclude the attainment of an effectively modified structure and adversely affect casting results. Modifier additions are usually accompanied by an increase in hydrogen content. In the case of sodium and calcium, the reactions involved in element solution are invariably turbulent or are accompanied by compound reactions that by their nature increase dissolved hydrogen levels. In the case of strontium, master alloys may be highly contaminated with hydrogen, and there are numerous indications that hydrogen solubility is increased after alloying. For sodium, calcium, and strontium modifiers, the removal of hydrogen by reactive gases also results in the removal of the modifying element. Recommended practices are to obtain modification through additions of modifying elements added to well-processed melts, followed by inert gas fluxing to acceptable hydrogen levels. No such disadvantages accompany antimony use. 13

Calcium and sodium can be added to molten aluminum in metallic or salt form. Vacuumprepackaged sodium metal is commonly used. Strontium is currently available in many forms, including aluminum-strontium master alloys ranging from approximately 10 to 90% Sr and Al-SiSr master alloys of varying strontium content. Very low sodium concentrations (approximately 0.001%) are required for effective modification. More typically, additions are made to obtain a sodium content in the melt of 0.005 to 0.015%. Remodification is performed as required to maintain the desired modification level. A much wider range of strontium concentrations is in use. In general, addition rates far exceed those required for effective sodium modification. A range of 0.015 to 0.050% is standard industry practice. Normally, good modification is achievable in the range of 0.008 to 0.015% Sr. Remodification through strontium additions may be required, although retreatment is less frequent than for sodium. To be effective in modification, antimony must be alloyed to approximately 0.06%. In practice, antimony is employed in the much higher range of 0.10 to 0.50%. The Importance of Phosphorus. It has been well established that phosphorus interferes with the modification mechanism. Phosphorus reacts with sodium and probably with strontium and calcium to form phosphides that nullify the intended modification additions. It is therefore desirable to use low-phosphorus metal when modification is a process objective and to make larger modifier additions to compensate for phosphorus-related losses. Primary producers may control phosphorus contents in smelting and processing to provide less than 5 ppm of phosphorus in alloyed ingot. At these levels, normal additions of modification agents are effective in achieving modified structures. However, phosphorus contamination may occur in the foundry through contamination by phosphate-bonded refractories and mortars and by phosphorus contained in other melt additions, such as master alloys and alloying elements including silicon. Effects of Modification. Typically, modified structures display somewhat higher tensile properties and appreciably improved ductility when compared to similar but unmodified structures. Improved performance in casting is characterized by improved flow and feeding as well as by superior resistance to elevated-temperature cracking.

Refinement of Hypereutectic Aluminum-Silicon Alloys


The elimination of large, coarse primary silicon crystals that are harmful in the casting and machining of hypereutectic silicon alloy compositions is a function of primary silicon refinement. Phosphorus added to molten alloys containing more than the eutectic concentration of silicon, made in the form of metallic phosphorus or phosphorus-containing compounds such as phosphorcopper and phosphorus pentachloride, has a marked effect on the distribution and form of the primary silicon phase. Investigations have shown that retained trace concentrations as low as 0.0015 through 0.03% P are effective in achieving the refined structure. Disagreements on recommended phosphorus ranges and addition rates have been caused by the extreme difficulty of accurately sampling and analyzing for phosphorus. More recent developments employing vacuum stage spectrographic or quantometric analysis now provide rapid and accurate phosphorus measurements. Following melt treatment by phosphorus-containing compounds, refinement can be expected to be less transient than the effects of conventional modifiers on hypoeutectic modification. Furthermore, the solidification of phosphorus-treated melts, cooling to room temperature, reheating, remelting, and resampling in repetitive tests have shown that refinement is not lost; 14

however, primary silicon particle size increases gradually, responding to a loss in phosphorus concentration. Common degassing methods accelerate phosphorus loss, especially when chlorine or freon is used. In fact, brief inert gas fluxing is frequently employed to reactivate aluminum phosphide nuclei, presumably by resuspension. Practices that are recommended for melt refinement are as follows:

Melting and holding temperature should be held to a minimum The alloy should be thoroughly chlorine or freon fluxed before refining to remove phosphorus-scavenging impurities such as calcium and sodium Brief fluxing after the addition of phosphorus is recommended to remove the hydrogen introduced during the addition and to distribute the aluminum phosphide nuclei uniformly in the melt

Refinement substantially improves mechanical properties and castability. In some cases, especially at higher silicon concentrations, refinement forms the basis for acceptable foundry results. Modification and Refinement. No elements are known that beneficially affect both eutectic and hypereutectic phases. The potential negative consequences of employing modifying and refining additions in the melt are characterized by the interaction of phosphorus with calcium, sodium, and strontium. Strontium has been claimed to benefit hypoeutectic and hypereutectic structures, but this claim has not been substantiated.

Metal Preparation
Regardless of the types of melting and holding furnaces and the particular gravity casting process used, there is great concern for reducing or eliminating dissolved hydrogen and entrained oxides. These procedures are less frequently employed for pressure die casting, in which concerns are focused on the dominant process-related causes of casting unsoundness, namely, entrapped gas and pouring injection-associated inclusions. Sensitivity to melt quality varies with the casting process and part design and necessitates special consideration of relevant criteria for each application. In general, the melt is processed to achieve hydrogen reductions and the removal of oxides to meet specific casting requirements. Modification and grain-refiner additions are made as appropriate to the given alloy and end product. Different melt preparation practices are employed in die casting operations because process related conditions are more dominant in the control of product quality than those controlled by melt treatment. For this reason, degassing for the removal of hydrogen, grain refinement, and modification or silicon refinement in the case of hypereutectic silicon alloys are often intentionally neglected. The movement toward higher integrity die castings has brought into focus the importance of the same melt quality parameters established and used in the gravity casting of aluminum alloys. In high-production die casting operations, the consumption of internal and external scrap is of primary importance in reducing base metal costs for the predominantly secondary alloy compositions that are consumed. Scrap crushing, shredding, and pre-treatment of various types precede melting, often in efficient induction systems. Oxides entrained in the melt as a result of this sequence of operations are dealt with through the use of salts and/or reactive gas fluxing. A concern in die casting is the formation of complex intermetallics that are insoluble at melt- holding temperatures and/or precipitate under holding conditions or during transfer to and injection from the hot or cold chamber. These intermetallics (sludge) affect furnaces, transfer systems, and, by inclusion, the quality of the castings produced. 15

Die casters are familiar with composition limits that prevent sludge formation. A common rule is that iron content plus two times manganese content plus three times chromium content should not exceed the sum of 1.7%. This limit is arbitrary and inexact, it is often assigned values from 1.5 through 1.9%, and it is subject to the specific composition and actual minimum process temperature.

Aluminum-Silicon Alloys
Abstract: Castings are the main use of aluminum-silicon alloys, although some sheet or wire is made for 16

welding and brazing, and some of the piston alloys are extruded for forging stock. Often the brazing sheet has only a cladding of aluminum-silicon alloy and the core consists of some other high melting alloy. The copper-free alloys are used for low- to medium-strength castings with good corrosion resistance; the copper-bearing for medium- to high-strength castings, where corrosion resistance is not critical. Because of their excellent castability, it is possible to produce reliable castings, even in complex shapes, in which the minimum mechanical properties obtained in poorly fed sections are higher than in castings made from higher-strength but lower-castability alloys. Castings are the main use of aluminum-silicon alloys, although some sheet or wire is made for welding and brazing, and some of the piston alloys are extruded for forging stock. Often the brazing sheet has only a cladding of aluminum-silicon alloy and the core consists of some other high melting alloy. The copper-free alloys are used for low- to medium-strength castings with good corrosion resistance; the copper-bearing for medium- to high-strength castings, where corrosion resistance is not critical. Because of their excellent castability, it is possible to produce reliable castings, even in complex shapes, in which the minimum mechanical properties obtained in poorly fed sections are higher than in castings made from higher-strength but lower-castability alloys. The alloys of this group fall within the composition limits: Si Cu Mg Zn 5-25% 0-5% 0-2% 0-3% Mn, Cr, Co, Mo Ni, Be, Zr Fe Na, Sr P up to 3% up to 3% < 0.02% < 0.01%

Silicon is the main alloying element; it imparts high fluidity and low shrinkage, which result in good castability and weldability. The low thermal expansion coefficient is exploited for pistons, the high hardness of the silicon particles for wear resistance. The maximum amount of silicon in cast alloys is of the order of 22-24% Si, but alloys made by powder metallurgy may go as high as 40-50% Si. Sodium or strontium produces the modification and phosphorus nucleates the silicon to permit of a fine distribution of the primary crystals. Iron is the main impurity and in most alloys efforts are made to keep it as low as economically possible, because of its deleterious effects on ductility and corrosion resistance. In sand castings and permanent mold castings the upper limit is usually 0.60.7% Fe. In some piston alloys iron may be added deliberately and in die-castings up to 3% Fe may be tolerated. Cobalt, chromium, manganese, molybdenum and nickel are sometimes added as correctives for iron; their addition also improves strength at high temperature. Copper is added to increase the strength and fatigue resistance without loss of castability, but at the expense of corrosion resistance. Magnesium, especially after heat treatment, increases substantially the strength, but at the expense of ductility. Zinc is a tolerated impurity in many alloys, often up to 1.5-2% Zn, because it has no substantial effect on room-temperature properties. Titanium and boron are sometimes added as grain refiners, although grain size in these alloys is not too important, because the properties are mainly controlled by the amount and structure of the silicon, as affected by modification produced by sodium additions or by phosphorus additions.

17

A distinction between dissolved and graphitic silicon is sometimes made by dissolving the alloy in acids, in which the dissolved silicon transforms in SiO2 whereas the graphitic remains uncombined. Prolonged or repeated heating tends to spheroidise the silicon. This spheroidising is faster in modified alloys and results in a coarsening of the silicon to a size very close to that of non modified material. In the absence of copper the iron is usually in the Al-FeSiAl5-Si eutectic as thin platelets interspread with the silicon needles or rods. If there is more than 0.8% Fe, primary FeSiAl5, crystals appear. Titanium and boron are usually added in amounts well within their solid solubility and do not form any separate phase. Iron reduces their solubility, so that less is needed for grain refinement; 0.10.2% V is reported to refine the FeMn compounds. Tin and lead, if present together with magnesium, tend to enter the Mg2Si phase. All the phases formed tend to concentrate at the grain boundaries, in the form of complex eutectics, more or less coupled. The lattice parameter is decreased slightly by silicon in solution and somewhat more by copper; none of the other elements affects it appreciably. Thus, the parameter of the alloys is between a = 4.045 x 10-10m and a = 4.05 x 10-10m, depending on composition and treatment. Thermal expansion is reduced substantially by silicon and much less pronouncedly by all other additions except magnesium, which tends to increase it slightly. Expansion coefficients at subzero temperatures also are some 10-20% lower than those for pure aluminum. A reduction of expansion coefficient by titanium and zirconium additions is reported, but it is very doubtful that it can be appreciable. Alloys produced by powder metallurgy containing up to 50% Si have even lower expansion coefficients. Permanent expansion accompanies precipitation out of solution of silicon, magnesium and copper; the amount varies but maybe as high as 0.15%. Thermal conductivity is of the order of 1.2-1.6 x 10-2W/m/K, the lower values being for the alloys cast in metallic molds or heat treated to retain silicon, copper or magnesium in solution. Electric conductivity depends mostly on the amount of silicon in solution; copper and magnesium also affect it. Values of the order of 35-40% IACS for annealed materials and of 22-35% IACS for solution treated alloys are reported. In the liquid state resistivity is some 10-15 times the resistivity at room temperature. Manganese, chromium, titanium, zirconium also reduce conductivity, and so does modification. Magnetic susceptibility is only slightly decreased by silicon, copper and magnesium, but depends mostly on manganese content. Mechanical properties. Alloys prepared from powders exhibit somewhat higher strengths, especially at elevated temperatures. In wrought products ultimate tensile strengths of 200-400 MPa, with elongation correspondingly from 20 to 2-3% are obtained. Poor casting technique may reduce the properties, although the aluminum-silicon alloys are among the least sensitive to such variables as gas content, design of castings, rate of cooling and feeding. High purity find special treatments can produce properties some 10-20% better than average, and, conversely, secondary alloys tend to have lower ductility than do primary ones. Casting under pressure improves properties toward those of forgings. Increasing silicon content increases strength at the expense of ductility, but this effect is not very marked. Modification by sodium produces a limited increase of strength, but the increase of ductility is substantial, especially in sand castings. At the higher cooling rates, normal with metal mold castings, the silicon is already somewhat refined without modification and the improvement from modification is reduced. The effect of cell size and dendritic arm spacing on mechanical properties of alloys with Si > 8% is not very marked, but in lower-silicon alloys, in which the aluminum dendrites predominate, the effect is normal. 18

Iron may slightly increase the strength, but drastically decreases the ductility, especially if above 0.7% Fe and not corrected by manganese, cobalt, etc. Beryllium, manganese, chromium, molybdenum, nickel, cobalt and zirconium all slightly increase the strength; manganese, cobalt, nickel and molybdenum, if needed to correct for the iron, can also increase the ductility; otherwise all of them reduce it. Beryllium is also reported to correct the iron effect. Copper and zinc increase the strength at the expense of ductility, but the most effective strengthener is magnesium, especially after heat treatment, provided that the amount and distribution of the magnesium are correct. Grain refinement by titanium, boron and zirconium additions has only a limited effect on mechanical properties. Silver additions are reported to increase the elongation. Antimony, tin, lead and cadmium decrease all properties, and antimony, by combining with magnesium, may reduce response to heat treatment. Calcium may increase strength and decrease elongation in straight aluminum-silicon alloys, but it has a deleterious effect on piston alloys. Compressive strength is higher than tensile by some 10-15%. Shear strength is approximately 70% of the tensile strength. Impact resistance is low, but so is notch sensitivity, as is to be expected in alloys that contain a large amount of hard, brittle second phase, often with sharp angles. Impact resistance is improved by spheroidising the silicon. The modulus of elasticity is of the order of 85-95 GPa, changing with temperature, as does tensile strength. A decrease in damping capacity with aging is reported. Properties at cryogenic temperatures are higher than at room temperature; there is little or no increase down to 170 K, but at 70 K the strength has become some 20% higher than at room temperature, with little or no decline in ductility. Notch strength does not change substantially at cryogenic temperatures. The effect of alloying elements on cryogenic properties is not too well established, but probably it is negligible. At high temperature the strength declines and the ductility increases. The decline is regular and more rapid than for other aluminum alloys except the aluminum-zinc-magnesium group. The slight increase in strength shown by heat treatable alloys, especially if only naturally aged, is only temporary, once the overaging stage is reached, there is a sharp drop and then the decline of strength with temperature becomes regular. Impact resistance increases with increasing temperature. At the higher temperatures elements with high melting points (copper, iron, manganese, nickel, cobalt, chromium, tungsten) reduce to some extent the decline in strength, although their effect is not substantial. Beryllium, too, is reported to improve the high-temperature strength. In spite of their poor high-temperature strength and fatigue resistance, aluminum-silicon alloys are used extensively for pistons because of their low expansion coefficient, good wear resistance and good castability. Hypereutectic alloys with up to 2-3% additions of copper, nickel, iron, manganese, chromium or magnesium are preferred, although good performance has been obtained also with hypoeutectic alloys and alloys low in heavy metals. Zinc, lead and tin decrease the high-temperature strength. Modified alloys have slightly lower high-temperature strength. Creep resistance is not particularly good. Silicon increases the creep resistance of aluminum much less than do most other alloying elements. Copper, iron, manganese, nickel, cobalt, chromium, etc., increase it, as is to be expected, and so do magnesium and rare earths. Fatigue resistance is relatively low, especially if the silicon is not modified or is spheroidised by heat treatment. Cobalt and manganese may improve the fatigue resistance. Pressure during freezing increases the fatigue strength and wear resistance; surface defects and complex loads reduce it, especially at high temperature. Fatigue strength drops gradually with temperature in straight 19

aluminum-silicon, but there is no drop up to 500 K in aluminum-copper-silicon alloys. The alloys are susceptible to thermal fatigue because of the substantial difference in expansion coefficient of the matrix and silicon particles. Wear resistance is very good, especially in hypereutectic alloys in which the hard silicon particles are well distributed either by phosphorus nucleation or by powder metallurgy fabrication, or in alloys to which bismuth has been added. Wear resistance of high-silicon alloys (20-25% Si) is 10 times better than that of plain steel and comparable with that of surface hardened steel. Friction in couples of steel against aluminum-silicon alloys decreases with surface perfection and hardness of the steel; however, aluminum-silicon alloys for bearings have not been successful unless they contain substantial tin. Corrosion resistance. Aluminum-silicon alloys without copper have good corrosion resistance in most reagents; only in alkaline solutions which attack silicon as well as aluminum their performance is poor. Copper reduces appreciably the corrosion resistance and so does iron, unless corrected with manganese or chromium. Zinc up to 2-3% has no effect. Tin and calcium also have a deleterious effect on corrosion resistance. Porosity decreases corrosion resistance. Corrosion by flowing water is more rapid than in still water, but of the same type. Aluminum-silicon alloys with iron and nickel have particularly good resistance to high-temperature water or steam. In secondary alloys, where many elements are present in small amounts, zinc and manganese compensate for copper and nickel, and corrosion resistance is reported as very close to that of primary alloys. Contact corrosion is especially poor in aluminum-silicon-copper alloys, but even copper-free alloys are worse in this respect than aluminum 99.8%. Machinability is poor, because the extreme hardness of the silicon combined with the relative softness of the matrix tends to wear the tools very rapidly. In hypereutectic alloys phosphorus additions that improve the silicon distribution improve machinability; but in hypoeutectic alloys phosphorus tends to reduce it, whereas sodium improves it. Copper reduces further the machinability for the same silicon content, especially after heat treatment, but same of the coppersilicon alloys with low silicon may have machinability equal to or better of high that of highsilicon, copper-free alloys. Iron, manganese, nickel, zinc, titanium, etc., do not decrease machinability.

20

Potrebbero piacerti anche