Vous êtes sur la page 1sur 27

Fundamentals of Joining

INTRODUCTION TO CONSOLIDATION PROCESSES


Large-size products, products with a high degree of shape complexity, or products with a wide variation in required properties are often manufactured as joined assemblies of two or more component pieces. These pieces may be smaller and therefore easier to handle, simpler shapes that are easier to manufacture, or segments that have been made from different materials. Assembly is an important part of the manufacturing process, and a wide variety of consolidation processes have been developed to meet the various needs. Each of the methods has its own distinctive characteristics, strengths, and weaknesses. The metallurgical processes of welding, brazing, and soldering are usually used to join metals and often involve the solidification of molten material. The use of discrete fasteners (such as nuts, bolts, screws, and rivets) requires the creation of aligned holes and produces stress localization. While the holes may affect performance, disassembly and reassembly can often be performed with relative ease. Adhesive bonding has grown with new developments in polymeric materials and is being used extensively in automotive and aircraft production. Any material can be joined to any other material, and low temperature joining is particularly attractive for composite materials. Production rates are often low, however, because of the time required for the adhesive to develop full strength. Welding is the permanent joining of two materials, usually metals, by coalescence, which is induced by a combination of temperature, pressure, and metallurgical conditions. The particular combination of these variables can range from high temperature with no pressure to high pressure with no increase in temperature. Because welding can be accomplished under a wide variety of conditions, a number of different processes have been developed. Welding is the dominant method of joining in manufacturing, and a large fraction of metal products would have to be drastically modified if welding were not available. Coalescence between two metals requires sufficient proximity and activity between the atoms of the pieces being joined to cause the formation of common crystals. The ideal metallurgical bond, for which there would be no noticeable or detectable interface, would require (1) perfectly smooth, flat, or matching surfaces; (2) surfaces that are clean and free from oxides, absorbed gases, grease, and other contaminants; (3) metals with no internal impurities. CLASSIFICATION OF WELDING AND THERMAL CUITING PROCESSES Wherever possible, this text will utilize the nomenclature of the American Welding Society (AWS). The various welding processes have been classified in the manner presented in Figure 30-1, and letter symbols have been assigned to facilitate process designation. The variety of processes provides multiple ways of achieving coalescence and makes it possible to produce effective and economical welds in nearly all metals and combinations of metals.

Types of Fusion Welds and Types of Joints

Heat Effects and Welding Metallurgy

Grain structure and various zones in a fusion weld.

Weldability

Gas Flame and Arc Processes


Oxyfuel-Gas Welding (OFW) The combustion of oxygen and acetylene (C2H2) by means of a welding torch of the type shown in figure 31-1 produces temperature of about 3250C in a two-stage reaction.

In the first stage, the supplied oxygen and acetylene react to produce carbon monoxide and hydrogen: C2H2 + O2 ---- 2CO + H2 + heat This reaction occurs near the tip of the torch and generates intense heat. The second stage of the reaction involves the combustion of the CO and H2 and occurs just beyond the first combustion zone. The specific reactions of the second stage are: 2CO + O2 --- 2CO2 + heat H2 + 1/2 O2 ---- H2O + heat

The oxygen for these secondary reactions is generally obtained from the surrounding atmosphere. The two-stage combustion process produces a flame having two distinct regions.

As shown in Figure 31-2, the maximum temperature occurs near the end of the inner cone, where the first stage of combustion is complete. Most welding should be performed with the torch positioned so that this point of maximum temperature is just above the metal being welded. The outer envelope of the flame serves to preheat the metal and, at the same time, provides shielding from oxidation, since oxygen from the surrounding air is consumed in the secondary combustion.

Figure 31-2 Typical oxyacetylene flame and the associated temperature distribution.

Three different types of flames can be obtained by varying the oxygen-acetylene (or oxygen-fuel gas) ratio. If the ratio is between 1:1 and 1.15:1, all reactions are carried to completion and a neutral flame is produced. Most welding is done with a neutral flame, since it will have the least chemical effect on the heated metal.

USES, ADVANTAGES, AND LIMITATIONS


Almost all oxyfuel-gas welding is fusion welding. The metals to be joined are simply where a weld is desired and no pressure is required. Because a slight gap often between the pieces is joined, filler metal can be added in the form of a solid metal rod. Figure 31-3 shows a schematic of oxyfuel-gas welding using a consumable welding rod.

Figure 31-3: Oxyfuel-gas welding using a consumable welding rod.

ARC WELDING
With the development of commercial electricity in the late nineteenth century, it was soon recognized that an arc between two electrodes was a concentrated heat source that could produce temperatures approaching 4000C. Figure 31-8 depicts the basic electrical circuit. If needed, filler metal was provided by a metallic wire or rod that was independently fed into the arc. The metal wire not only carried the welding current but, as it melted in the arc; it also supplied the necessary filler.

Figure 31-8: The basic electrical circuit for arc welding. All arc- welding processes employ the basic circuit depicted in Figure 31-8. Welding currents vary from 1 to 4000 amps, with the range from 100 to 1000 being most typical. Voltages are generally in the range of 20 to 50 volts. If direct current is used' electrode is made negative, the condition is known as straight polarity (DCSP) or DCEN, for direct-current electrode-negative. Electrons are attracted to the positive work piece, while ionized atoms in the arc column are accelerated toward the negative electrode. Since the ions are far more massive than the electrons, the heat of the arc is more concentrated at the electrode. DCEN processes are characterized by fast melting of the electrode (high metal deposition rates) and a shallow molten pool on the work piece (weld penetration).

If the work is made negative and the electrode positive, the condition is known as reverse polarity (DCRP) or DCEP, for direct-current electrode positive. The positive ions impinge on the work piece, breaking up any oxide films and giving deeper penetration. The metal deposition rate is lower, however.

In one group of arc-welding processes, the electrode is consumed (consumable electrode processes) and thus supplies the metal needed to fill the joint. Consumable electrodes have a melting temperature below the temperature of the arc. CONSUMABLE-ELECTRODE ARC WELDING
Four processes make up the bulk of consumable-electrode arc welding: 1. Shielded metal arc welding (SMAW) 2. Flux-cored arc welding (FCAW) 3. Gas metal arc welding (GMAW) 6

4. Submerged arc welding (SAW) These processes all have a medium rate of heat input and produce a fusion zone whose depth is approximately equal to its width. Because

the fusion zone is composed of metal from both of the pieces being joined, plus melted filler (i.e., electrode), the electrode must be of the same material as that being
welded; the ceramics. processes cannot be used to join dissimilar metals or

SHIELDED METAL ARC WELDING


Shielded metal arc welding (SMAW), also called stick welding or covered-electrode welding, is among the most widely used welding processes because of its versatility and because it requires only low-cost equipment. The key to the process is a finite-length electrode that consists of metal wire, usually from 1.5 to 6.5 mm in diameter and 20 to 45 cm in length.

Surrounding the wire is a bonded coating containing chemical components that add a number of desirable characteristics, including all or many of the following: 1. Vaporize to provide a protective atmosphere (a gas shield around the arc and pool of molten metal). 2. Provide ionizing elements to help stabilize the arc, reduce weld metal spatter, and increase efficiency of deposition. 3. Act as a flux to deoxidize and remove impurities from the molten metal. 4. Provide a protective slag coating to accumulate impurities, prevent oxidation, and slow the cooling of the weld metal. 5. Add alloying elements. 6. Add additional filler metal. 7. Affect arc penetration (the depth of melting in the workpiece). 8. Influence the shape of the weld bead.
The coated electrodes are classified by the tensile strength of the deposited weld metal, the welding position in which they may be used, the preferred type of current and polarity (if direct current), and the type of coating. Electrode selection consists of determining the electrode coating, coating thickness, electrode composition, and electrode diameter. The current type and polarity are matched to the electrode.

Flux-Cored Arc Welding [FCAW]

FIGURE 31-13: The flux-cored arc welding (FCAW) process.

GAS METAL ARC WELDING If the supplemental shielding gas flowing through the torch (described above) becomes the primary protection for the arc and molten metal, there is no longer a need for the volatilizing flux. The consumable electrode can now become a continuous, solid, uncoated metal wire. The resulting process, shown in figure 31-14, was formerly called metal inert-gas, or MIG, welding and is now known as gas metal welding (GMAW). The arc is still maintained between the workpiece and the automatically fed bare-wire electrode, which continues to provide the necessary filler metal.

FIGURE 31-14 Schematic diagram of gas metal arc welding (GMAW). Electrode diameters range from 0.6 to 6.4 mm. The welding current, penetration depth, and process cost are all similar to the flux-cored process. Because shielding is provided by the flow of gas, and fluxing and slag-forming agents are no longer required, the gas metal arc process can be applied to all metals. Argon, helium, and mixtures of the two are the primary shielding gases. When welding steel, some O2 or CO2 is usually added to improve the arc stability and reduce weld spatter.

SUBMERGED ARC WELDING


No shielding gas is used in the submerged arc welding (SAW) process, depicted in Figure 31-15. Instead, a thick layer of granular flux is deposited just ahead of a solid bare-wire consumable electrode, and the arc is maintained beneath the blanket of flux with only a few small flames being visible. A portion of the flux melts and acts to remove impurities from the rather large pool of molten metal, while the unmelted excess provides additional shielding. The molten flux solidifies into a glasslike covering over the weld. This layer, along with the flux that is not melted provides good thermal insulation. The slow cooling of the weld metal helps to produce soft, ductile welds.

NONCONSUMABLE-ELECTRODE ARC WELDING


GAS TUNGSTEN ARC WELDING Gas tungsten arc welding (GTAW) was formerly known as tungsten inert-gas (TIG) welding. A nonconsumable tungsten electrode provides the arc but not the filler metal. Inert gas (argon, helium, or a mixture of them) flows through the electrode holder to provide a protective shield around the electrode, the arc, the pool of molten metal, and the adjacent heated areas. (Note: CO2 cannot be used in this process since it provides inadequate protection for the hot tungsten electrode.) While argon is the most widely used gas, and produces a smoother, more stable arc, helium may be added to increase the heat input (higher welding speeds and deeper penetration). Helium alone may be preferred for overhead welding since it is lighter than air and flows upward. In applications where there is a close fit between the pieces being joined metal may be needed. When filler metal is required, it is usually supplied as a separate rod, about a meter long and in various diameters, as illustrated in Figure 31-19. The filler metal is generally selected to match the chemistry and/or tensile strength of the metal being welded. When high deposition rates are desired, a separate resistance circuit can be provided to preheat the filler wire.

RESISTANCE AND SOLID-STATE WELDING PROCESSES


THEORY OF RESISTANCE WELDING In resistance welding, heat and pressure are combined to induce coalescence. Electrodes are placed in contact with the material, and electrical resistance heating is used to raise the temperature of the work pieces and the interface between them. The same electrodes that supply the current also apply the pressure, which is usually varied throughout the weld cycle. A certain amount of pressure is applied initially to hold the workpieces in contact and thereby control the electrical resistance at the interface.
10

When the proper temperature has been attained, the pressure is increased to induce coalescence. Because pressure is utilized, coalescence occurs at a lower temperature than that required for oxyfuel gas or arc welding. In fact, melting of the base metal does not occur in many resistance-welding operations. The required temperature can often be attained, and coalescence can be achieved, in a few seconds or less. Resistance welding, therefore, is a very rapid and economical process, extremely well suited to automated manufacturing. No filler metal is required, and the tight contact maintained between the workpieces excludes air and eliminates the need for fluxes or shielding gases.

Figure 32-2: The desired temperature distribution across the electrodes and the work pieces during resistance welding

Figure 32-3: A typical current and pressure cycle for resistance welding. This cycle includes forging and postheating operations.

11

Figure 32-4: The arrangement of the electrodes and Work pieces in resistance spot welding.

Figure 32-5: A spot welding nugget between two sheets 1.3 mm Aluminum Alloy. The nugget is not symmetrical because the radius of the upper electrode is greater than that of the lower electrode.

12

FRICTION-STIR WELDING
The A relatively new process, first performed by the Welding Institute of Great Britain in 1991, friction-stir welding (FSW) has matured rapidly and currently offers significant benefits compared to conventional methods. As illustrated in Figure 32-18, a non-consumable welding tool (containing a shoulder and protruding cylindrical or tapered or pin) is rotated at several hundred revolutions per minute. It is then lowered into the interface between pieces of rigidly clamped material, and frictional heat is generated along the top surface (under the rotating shoulder) and along the surfaces of the rotating probe. After a period of time for heating and softening, the tool is driven along the material interface. As the probe traverses, a plasticized region is continually created. This softened material is swept along the periphery of the pin, flows to the back advancing probe, and coalesces to form a solid-state bond. The most common application is the formation of butt welds, usually between plates of the lower-melting-point metals (both wrought and cast alloys) or thermoplastic polymers.

Weld quality is excellent. The extensive plastic deformation creates a refined grain structure with no entrapped oxides or gas porosity. As a result, the strength, ductility, fatigue life, and toughness of the resulting weld are all quite good. Welds in aircraft aluminum are 30 to 50% stronger than those formed by arc welding. Since no material is melted, both wrought and cast alloys can be joined, and they can be joined to each other. No filler material or shielding gas is required, and the process is environmentally friendly (no fumes, weld spatter, or arc glare). Because of the high energy efficiency, total heat input and associated distortion and shrinkage are all low. Joint preparation is minimal, and surface oxides need not be removed. Welding can be performed in any position and requires access to only one side of the plate. Gaps up to 10% of the material thickness can be accommodated with no reduction in weld quality or performance.

ELECTRON-BEAM WELDING
In the electron-beam welding (EBW) process, the metal to be welded is heated by the impingement of a beam of high-velocity electrons. Originally developed for 13

obtaining ultra-high-purity welds in reactive and refractory metals, the unique qualities of the process have led to a much wider range of applications.

Figure 33-2 presents the electron optical system. An electric current heats a tungsten filament to about 2200C, causing it to emit a stream of electrons by thermal emission. By means of a control grid, accelerating voltage, and focusing coils, these electrons are collected into a concentrated beam, accelerated, and directed to a focused spot between 0.8 and 3.2 mm in diameter. Since electrons accelerated 150 kV achieve speeds nearly 2/3 the speed of light, the electron beam is concentrated energy, capable of producing temperatures in excess of 1 million degrees Celsius when its kinetic energy is converted to heat. Since the beam is composed of charged particles, it can be positioned and moved by electromagnetic lenses. Unfortunately, the electrons cannot travel well through air. To be effective as a welding heat source, the beam must be generated and focused in a very high vacuum, typically at pressures of 0.01 Pa or less.

Almost any metal can be welded by the electron-beam process, including those are difficult to weld by other methods, such as zirconium, beryllium, and tungsten. Dissimilar metals, including those with extremely different melting points, can also be readily welded, since the intense beam will melt both metals simultaneously. Electronbeam welds typically exhibit a narrow profile and remarkable penetrations. The high power and heat concentrations can produce fusion zones with depth-to-width ratios up to 25:1. This is coupled with low total heat input, low distortion, and a very narrow heataffected zone. Heat-sensitive materials can often be w without damage to the base metal. Deep welds can be made in a single pass. High welding speeds are common; no shielding gas, flux, or filler metal is required. The process can be performed in all positions; and preheat or postheat is generally unnecessary. 14

LASER-BEAM WELDING
Laser beams can be used as a heat source for welding, hole making, cutting, cladding, and heat treating a wide variety of engineering metals. When used for laser-beam welding (LBW), the beam of coherent light can be focused to a diameter of 0.1 to 1.0 mm, providing a power density in excess of 106 watts/mm2. The high-intensity beam can be used to simply melt the material at the joint, but more often, it produces a very narrow column of vaporized metal (a "keyhole") with a surrounding liquid pool. As the beam traverses, the liquid flows into the joint to produce a weld with a depth-towidth ratio generally greater than 5:1. Because of the narrow weld-pool geometry, high travel speed of the beam (typically several meters per minute), and low total heat input, the molten metal solidifies quickly, producing a very thin heat-affected zone and little thermal distortion. Finishing costs are quite low. Since welds require only one-side access, many different joint configurations are possible. As shown in Figure 33-5, laser-beam welding and electron-beam welding both offer some of the highest power densities of the welding processes. The well-collimated beam of intense laser energy can produce deep penetration welds that are similar to electron-beam welds, but the laserbeam technique offers several distinct advantages: 1. The beam can be transmitted through air (i.e., a vacuum environment is not required). There is no physical contact between the welding equipment and the work piece. 2. No harmful X-rays are generated. 3. The laser beam can be easily shaped, directed, and focused with both transmission and reflective optics (lenses and mirrors), and some beams can be transmitted through fiber-optic cables. 4. The only restriction on weld location is optical accessibility. Welds can be made in difficult-to-reach places, and materials can be joined within transparent container such as inside a vacuum tube.

15

A laser-welding system consists of an industrial laser, a means of guiding and focusing the beam, and a means of positioning and manipulating the parts to be welded. The traditional equipment for laser welding is a CO2 laser with a power output range of 1.5 to 10 kW (or even higher). Because CO2 lasers emit light with a far-infrared wavelength, they require mirror systems or special optical materials to focus and position the beam. They can be used to weld steel up to 25 mm thick at speeds ranging from 1 to 20 m/min. Nd:YAG (neodymium: yttrium-aluminum-garnet) lasers are more limited in power and capability but operate in a near-infrared wavelength that can utilize conventional glass lenses or delivery by flexible fiber-optic cable (as much as 3000 W of energy can be transmitted up to 150 m through a 0.6-mm-diameter fiber).

BRAZING
In brazing and soldering, the surfaces to be joined are first cleaned, the components assembled or fixtured, and a low-melting-point nonferrous metal is then melted, drawn into the space between the two solids by capillary action, and allowed to solidify. Brazing is the permanent joining of similar or dissimilar metals or ceramics (or composites based on those two materials) through the use of heat and a filler metal whose melting temperature (actually, liquidus temperature) is above 450C, but below the melting point (or solidus temperature) of the materials being joined.

process is different from welding in a number of ways:

The brazing

1. The composition (or chemistry) of the brazing alloy is significantly different from that of the base metal. 2. The strength of the brazing alloy is usually lower than that of the base metal. 3. The melting point of the brazing alloy is lower than that of the base metal, so none of the base metal is melted. 4. Bonding requires capillary action to distribute the filler metal between the closely fitting surfaces of the joint. The specific flow is dependent upon the viscosity of the liquid, the geometry of the joint, and surface wetting characteristics.

16

Because of these differences, the brazing process has several distinct advantages: 1. A wide range of metallic and nonmetallic materials can be brazed. The process is ideally suited for joining dissimilar materials, such as ferrous metal to nonferrous metal, cast metal to wrought metal, metals with widely different melting points, or even metal to ceramic. 2. Since less heating is required than for welding, the process can be performed quickly and economically. 3. The lower temperatures reduce problems associated with heat-affected zones, warping, and distortion. Thinner and more complex assemblies can be joined successfully. Thin sections can be joined to thick. Metal as thin as 0.01 mm and as thick as 150 mm can be brazed.

17

4. Assembly tolerances are closer than for most welding processes, and joint appearance is usually quite neat. 5. Brazing is highly adaptable to automation and performs well when massproducing complex or delicate assemblies. Complex products can also be brazed in several steps using filler metals with progressively lower melting temperatures. 6. A strong permanent joint is formed.

Successful brazing or soldering requires that the parts have relatively good fit-up (i.e., small joint clearances) to promote capillary flow of the filler metal. The parts must be thoroughly cleaned prior to joining, and many parts will require flux removal after joining. It is also important to remember that any subsequent heating of the assembly can cause inadvertent melting of the braze metal, thereby weakening or destroying the joint.

SOLDERING
Soldering is a brazing-type operation where the filler metal has a melting temperature (or liquidus temperature if the alloy has a freezing range) below 450C. It is typically used for joining thin metals, connecting electronic components, joining metals while avoiding exposure to high elevated temperatures, and filling surface flaws and defects. The process generally involves six important steps: (1) Design of an acceptable joint; (2) Selection of the correct solder for the job; (3) Selection of the proper type of flux;
18

(4) Cleaning of the faces to be joined; (5) Application of flux, solder, and sufficient heat to allow the molten solder to fill the joint by capillary action and solidify; and (6) Removal of the flux residue.

METALS TO BE JOINED
Copper, silver, gold, and tin-plated steels are all easily soldered. Aluminum has a strong, adherent oxide that makes soldering difficult. Special fluxes and modified techniques may be required, but adequate joints are indeed possible, as shown by the number of soldered aluminum radiators currently in automotive use.

PHYSICS OF WELDING
Although several coalescing mechanisms are available for welding, fusion is by far the most common means. To accomplish fusion, a source of high-density heat energy is applied to the faying surfaces, and the resulting temperatures are sufficient to cause localized melting of the base metals. If a filler metal is added, the heat density must be high enough to melt it also. Heat density can be defined as the power transferred to the W per unit surface area, W/mm2. The time to melt the metal is inversely proportional to the power density. At low power densities, a significant amount of time is required to cause melting. If power density is too low, the heat is conducted into work as rapidly as it is added at the surface, and melting never occurs. It has been found that the minimum power density required to melt most metals in welding is about 10 W/mm2. As heat density increases, melting time is reduced. If power density is too high-above around 105 W/mm2the localized temperatures vaporize the metal in the affected region.

19

Thus, there is a practical range of values for power density within which welding can be performed. Differences among welding processes in this range are (1) the rate at which welding can be performed and / or (2) the size of the region that can be welded. Table 1 provides a comparison of power densities for the major groups of fusion welding processes. Oxyfuel gas welding is capable of developing large amounts of heat, but the heat density is relatively low because it is spread over a large area. Oxyacetylene gas, the hottest of the OFW fuels, burns at a top temperature of around 3500C. By comparison, arc welding produces high energy over a smaller area, resulting in local temperatures of 5500 to 6600C. For metallurgical reasons, it is desirable to melt the metal with minimum energy, and high heat densities are generally preferable. TABLE 1 Comparison of several fusion-welding processes on the basis of their power densities.

Approximate Power Density W/mm2 Welding Process Oxyfuel welding 10 Arc welding 50 Resistance welding 1,000 Laser beam welding 9,000 Electron beam welding 10,000

Power density can be computed as the power entering the surface divided by the corresponding surface area: PD = P / A (1)

where PD = power density, W/mm2; P = power entering the surface, W; and A = surface area over which the energy is entering, mm2. The issue is more complicated than indicated by Eq. (1). One complication is that the power source (e.g., the arc) is moving in many welding processes, which results in preheating ahead of the operation and postheating behind it. Another complication is that power density is not uniform throughout the affected surface; it is distributed as a function of area, as demonstrated by the following example. Example: A heat source is capable of transferring 3000 W to the surface of a metal part. The heat impinges the surface in a circular area, with intensities varying inside the circle. The distribution is as follows: 70% of the power is transferred within a circle of diameter = 5 mm, and 90% is transferred within a concentric circle of diameter = 12 mm. What are the power densities in (a) the 5 mm diameter inner circle and (b) the 12 mm diameter ring that lies around the inner circle? Solution: (a) The inner circle has an area A = (5)2 / 4 = 19.63 mm2. The power inside this area P = 0.70 X 3000 = 2100 W Thus the power density PD = 2100 / 19.63 = 107 W/ mm2 (b) The area of the ring outside the inner circle is A = (122 - 52) /4= 93.4 mm2 20

The power in this region P = 0.9 X 3000 - 2100 = 600 W

The power density is therefore PD = 600 / 93.4 = 6.4 W/mm2. Observation: The power density seems high enough for melting in the inner circle, but probably not sufficient in the ring that lies outside this inner circle. The quantity of heat required to melt a given volume of metal is the sum of (1) the heat to raise the temperature of the solid metal to its melting point, which depends on the metals volumetric specific heat, and (2) the heat to transform the metal from solid to liquid phase at the melting point, which depends on the metal's heat of fusion. To a reasonable approximation, this quantity of heat can be estimated by: Um = KTm2 where U
m

(2)

is the unit energy for melting - the quantity of heat required to melt a unit

volume of metal starting from room temperature, J/mm3; T m = melting point of the metal on an absolute temperature scale, K; and K = constant whose value is 3.33 x 10-6 when the Kelvin scale is used. Absolute melting temperatures for selected metals are presented in Table 2.

Not all of the input energy is used to melt the weld metal. There are two heat transfer mechanisms at work, both of which reduce the amount of heat available to the welding process.

The first mechanism

is the transfer of heat between the heat source and the

surface of the work. This process has a certain heat transfer efficiency, f1, defined as the ratio of the actual heat received by the workpiece divided by the total heat generated at the source. The

second mechanism involves the conduction of heat away from the weld area

to be dissipated throughout the work metal, so that only a portion of the heat transferred to the surface is available for melting. This melting efficiency, f2, is the proportion of heat received at the work surface that can be used for melting. The combined effect of these two efficiencies is to reduce the heat energy available for welding as follows: Hw = f 1 f 2 H (3)

where Hw= net heat available for welding, J, f1 = heat transfer efficiency, f2 = the melting efficiency, and H = the total heat generated by the welding process, J.
21

Table 2: Melting temperatures on the absolute temperature scale for selected metals.
Melting Temperature o b Metal Ka R Aluminium alloys 930 (1680) Cast iron 1530 (2760) Copper and alloys Metal Steels Low carbon Medium carbon High carbon Low alloy Ka 1760 1700 1650 1700
o

Rb

(3160) (3060) (2960) (3060)

Pure 1350 (2440) Brass. navy 1160 (2090) Bronze (90 Cu1120 (2010) Stainless steels 10 Sn) Lnconel 1660 (3000) Austenitic Magnesium 940 (1700) Martensitic Nickel 1720 (3110) Titanium Based on values a Kelvin scale = Centigrade (Celsius) temperature + 273. bRankine scale = Fahrenheit temperature + 460.

1670 1700 2070

(3010) (3060) (3730)

It is appropriate to separate f1 and f2 in concept, even though they act in concert during the welding process. Heat transfer efficiency f1 is determined largely by the welding process and the capacity to convert the power source (e.g., electrical energy) into usable heat at the work surface. Oxyfuel gas welding processes are relatively inefficient in this regard, while arc welding processes are relatively efficient. Melting efficiency f2 depends on the welding process, but it is also influenced by the thermal properties of the metal, joint configuration, and work thickness. Metals with high thermal conductivity, such as aluminum and copper, present a problem in welding because of the rapid dissipation of heat away from the heat contact area. The problem is exacerbated by welding heat sources with low energy densities (e.g., oxyfuel welding) because the heat input is spread over a larger area, thus facilitating conduction into the work. In general, a high-intensity welding heat source, combined with a low conductivity work material, results in a high melting efficiency. We can now write a balance equation between the energy input and the energy needed for welding: Hw = Um V (4)

where Hw = net heat energy delivered to the operation, J; Um = unit energy required to melt the metal, J/mm3; and V = the volume of metal melted, mm3. Most welding
22

operations are rate processes; that is, the net heat energy Hw is delivered at a given rate, and the weld bead is made at a certain travel velocity. This is characteristic for example of most arc welding and many oxyfuel gas welding operations. It is therefore appropriate to express Eq. (4) in the form of a rate balance equation: HRw = Um WVR (5)

where HRw = rate of heat energy delivered to the operation, J/s = W; and WVR = volume rate of metal welded, mm3/s. In the welding of a continuous bead, the volume rate of metal welded is the product of weld area Aw and travel velocity v. Substituting these terms into the above equation, the rate balance equation can now be expressed as HRw = f1f2HR = UmAwv (6)

Where f1 and f2 are the heat transfer and melting efficiencies; HR = rate of input energy generated by the welding power source, W; Aw = weld cross-sectional area, mm2; and v = the travel velocity of the welding operation, mm/s.

Example
The power source in a particular welding setup is capable of generating 3500 W that can be transferred to the work surface with an efficiency f1= 0.7. The metal to be welded is low carbon steel, whose melting temperature, from Table 2, is Tm = 1760K. Melting efficiency in the operation is f2 = 0.5. A continuous fillet weld is to be made with a cross sectional area Aw= 20 mm2. Determine the travel speed at which the welding operation can be accomplished. Solution: Let us first find the unit energy required to melt the metal Um from Eq. (2). Um = 3.33(10-6) x 17602 = 10.3 J/mm3 Rearranging Eq. (6) to solve for travel velocity, we have v = f1 f2 HR Um Aw and solving for the conditions of the problem, v = 0.7 (0.5)(3500) = 5.95 mm/s 10.3 (20)

Power Source in Arc Welding Both direct current (DC) and alternating current (AC) are used in arc welding. AC machines are less expensive to purchase and operate, are generally
23

restricted to welding of ferrous metals. DC equipment can be used on metals with good results and is generally noted for better arc control. In all AW processes, power to drive the operation is the product of the current passing through the arc and the voltage E across it. This power is converted into heat, but not all of the heat is transferred to the surface of the work. Convection, conduction, radiation, and spatter account for losses that reduce the amount of usable heat. The effect of the losses is expressed by the heat transfer efficiency f1. Some representative values of f1 for several AW processes are given in Table 3. Heat transfer efficiency is greater for AW processes that use consumable electrodes because most of the heat consumed in melting the electrode is subsequently transferred to the work as molten metal. The process with the lowest f1 value in Table 3 is gas tungsten arc welding, which uses a non-consumable electrode. Melting efficiency f2 further reduces the available heat for welding. The resulting power balance in arc welding is defined by HRw = f1 f2 I E = UmAwv where E = voltage, V; I = current, A; and the other terms were defined earlier. The units of HRw that result from product of amps x voltage are watts, which equal joule/sec. This can be converted to Btu/sec by recalling that 1 Btu = 1055 joule.
TABLE 3: Heat transfer efficiencies for several arc-welding processes. Typical Heat Transfer Arc Welding Process. Shielded metal arc welding Gas metal arc welding Flux-cored arc welding Submerged arc welding Gas tungsten arc welding Example: A gas tungsten AW operation is performed at a current of 300 A and voltage of 20 V. The melting efficiency f2 = 0.5, and the unit melting energy for the metal Um = 10 J/mm3. Determine (a) power in the operation, (b) rate of heat generation at the weld, and (c) volume rate of metal welded. Solution: (a) The power in this arc-welding operation is P = IE = (300 A)(20 V) = 6000 W (b) From, the heat transfer efficiency f1 = 0.7. The rate of heat used for welding is given by 24 Efficiency f1 0.9 0.9 0.9 0.95 0.7

HRw = f1 f2 IE = (0.7)(0.5)(6000) = 2100 W = 2100 J/s (c) The volume rate of metal welded is WVR = (2100 J/s) /(10 J/mm3) = 210 mm3/s

Power Source in Resistance Welding


The heat energy supplied to the welding operation depends on current flow, resistance of the circuit, and length of time the current is applied. This can be expressed by the equation H = I2 R t where H = heat generated, J (to convert to Btu divide by 1055); I = current, A; R = electrical resistance, ohm; and t = time, s. The current used in resistance welding operations is very high (5,000 to 20,000 A, typically), although voltage is relatively low (usually below 10 V). The duration t of the current is short in most processes, perhaps lasting 0.1 to 0.4 s in a typical spot welding operation.

Current is so high in RW because the squared term in Eq. amplifies the effect of current, and resistance is very low (around 0.0001 ohm). Resistance in the welding circuit is the sum of (1) resistance of the electrodes, (2) resistances of the work parts, (3) contact resistances between electrodes and work parts, and (4) contact resistance of the faying surfaces.
The ideal situation is for the faying surfaces to be the largest resistance in the sum, since this is the desired location of the weld. The resistance of the electrodes is minimized by using metals with very low resistivities, such as copper. The resistances of the work parts are a function of the resistivities of the base metals involved and the thicknesses of the parts. The contact resistance between the electrodes and the parts is determined by the contact areas (i.e., size and shape of the electrode) and the condition of the surfaces (e.g., cleanliness of the work surfaces and scale on the electrode). Finally, the resistance at the faying surfaces depends on surface finish, cleanliness, contact area, and pressure. No paint, oil, dirt, or other contaminants should be present to separate the contacting surfaces. 25

EXAMPLE A resistance spot welding operation is performed on two pieces of 1.5 mm thick sheet steel using 12,000 amps for a 0.20 second duration. The electrodes are 6 mm in diameter at the contacting surfaces. Resistance is assumed to be 0.0001 ohms, and the resulting weld nugget is 6 mm in diameter and 2.5 mm thick. The unit melting energy for the metal Um = 12.0 J/mm3. What portion of the heat generated was used to form the weld, and what portion was dissipated into the surrounding metal? Solution: The heat generated in the operation is given by Eq. as H = (12,000)2 (0.0001) (0.2) = 2880 J The volume of the weld nugget (assumed disc-shaped) is V = 2.5 (6)2= 70.7 mm3. The heat required to melt this volume of metal is Hm = 70.7(12.0) = 848 J. The remaining heat, 2880 - 848 = 2032 J (70.6% of the total), is absorbed into the surrounding metal.

Success in resistance welding depends on pressure, as well as heat. The principal functions of pressure in RW are to (1) force contact between the electrodes and the workparts and between the two work surfaces prior to applying current; and (2) press the faying surfaces together to accomplish coalescence when the proper welding temperature has been reached. There are some general advantages of resistance welding: (1) no filler metal is required, (2) high production rates are possible, (3) it lends itself to mechanization and automation, (4) operator skill level is lower than that required for arc welding, and (5) good repeatability and reliability. Drawbacks are that initial equipment cost is high- usually much higher than AW operations, and the types of joints that can be welded are limited to lap joints for most RW processes.

Electron beam welding (EBW) is a fusion-welding process in which the heat for welding is provided by a highly focused, high-intensity stream of electrons impinging against the work surface. The equipment is similar to that used for electron-beam machining. The electron beam gun operates at high voltage to accelerate the electrons (e.g., 10 to 150 kV typical), and
26

beam currents are low (measured in milliamps). High-power density is achieved by focusing the electron beam on a very small area of the work surface, so that the power density PD is based on PD = f1 E I / A where PD = power density, W/mm2; f1 = heat transfer efficiency (typical values for EBW range from 0.8 to 0.95); E = accelerating voltage, V; I = beam current, A; and A = the work surface area on which the electron beam is focused, mm2. Typical weld areas for EBW range from 13 X 10-3 to 2000 x 10-3 mm2.

27

Vous aimerez peut-être aussi