You are on page 1of 31

Brazing, Soldering and Adhesive Bonding

1.0

Introduction

Brazing, soldering, and adhesive bonding are somewhat similar to welding. Both brazing and soldering use a filler metal to joint two or more metal parts (like welding). However, a soldered or brazed joint can be disassembled after being formed, though it may be difficult. The strength of a brazed joint and soldered joint lies between joints produced by fusion and solid state welding (fusion welding being stronger). Brazing and soldering both always use a filler metal. No melting of the base metal occurs in either of these two operations. Both these processes are considered distinct from welding. Brazing and soldering are particularly attractive when The metals have poor weldability (difficult to melt, for example) Dissimilar metals are to be joined (one may melt before the other in welding) The intense heat of welding may damage the components being joined The geometry of the joint does not lend itself to any welding methods High strength is not a requirement

Adhesive bonding has certain similarities to brazing and soldering. It uses the attachment between a filler and two closely placed surfaces to bond the parts. The main differences are that adhesive bonding is not metallic, the joined components need not be metallic, and joining process is carried out at room temperature or only modestly above.

2.0

Brazing

In this process, the filler is molten and distributed between the faying surfaces by capillary action. No melting of the base metals occur only the filler melts. In brazing, the filler (also called brazing metal) has a melting temperature that is about 450 oC (U.S standard above 800 oF or 427 oC) , but below the melting temperature of the metals to be joined. If the joint is properly designed and the brazing operation has been performed properly, the resulting joint will be stronger than the filler used. This is due to the small part clearances in brazing, the metallurgical bonding that occurs between the base and filler, and geometric constriction that are imposed on the joint by the base parts. Brazing has several advantage over welding: Any metal can be joined, including dissimilar metals Certain brazing methods can be performed quickly and consistently, permitting high cycle rates and automated production Some methods allow multiple joints to be brazed simultaneously Brazing can be used to join thin-welded parts that cannot be welded

In general, less heat and power are required than fusion welding Problems with HAZ in the base metal near the joint is minimised Joints inaccessible by many welding operations can be brazed since capillary action draws the molten filler metal into the joint.

The disadvantages include: Joint strength is generally less than that of a welded joint Although strength of a good brazed joint is greater than that of the filler metal, it is likely to be less than that of the base metals High service temperatures may weaken a brazed joint The colour of the metal in a brazed joint may not match the colour of the base metal parts, a possible aesthetic disadvantage

Brazing is used in a wide variety of industries, including automotive (joining tubes and pipes), electrical equipment (joining wires and cables), cutting tools (brazing cemented carbide inserts into shanks), and jewellery making. In addition, the chemical processing industry and plumbing and heating contractors join metal pipes and tubes by welding. The process is used extensively for repair and maintenance work in nearly all industries.

2.1

Brazed joints

Brazed joints are commonly of two types butt and lap. However, these two are adapted for brazing in various ways. The conventional butt joint provides a very limited area for brazing, jeopardizing the strength of the joint. To increase the area of the faying joint, the mating parts are usually scarfed or stepped or otherwise altered. This requires additional processing. One of the particular difficulties associated with these joints is problem of maintaining its alignment before and after brazing. Lap joints are more widely used in brazing, as they provide a relatively large interface between the parts. An overlap of at least three times the thickness of the thinner part is considered good practice. An advantage of brazing over welding for lap joints is that the filler is bonded to the base part throughout the entire area between the parts, rather than only at the edges or at discrete spots.

Figure 01: (a) conventional butt joint. Adaptations of butt joint: (b) scarf joint, (c) stepped butt joint, (d) increased cross-section of the part at the joint

Figure 02: (a) Conventional lap joint. Adaptations for brazing: (b) cylindrical parts, (c) sandwiched parts, (d) use of sleeve to convert butt into lap joint. Clearance between the mating parts is important in brazing. The clearance must be large enough not to restrict the flow of molten filler and prevent it from covering the entire surface. However, if the joint clearance is too high, capillary action will be reduce and areas will be formed where no filler metal is present. Joint strength is affected by clearance. There is an optimum clearance value at which joint strength is maximized. The issue is complicated by the fact that the optimum depends on base and filler metals, joint configuration and processing conditions. Typical brazing clearances in practice are 0.025 and 0.25 mm. However, in some brazing operations it is not uncommon to have joint clearances around 0.6 mm. These values represent clearances at brazing temperature which may differ from the clearances at room temperature, depending in the thermal expansion of the base metals.

Figure 03: Joint strength as a function of joint clearance Cleanliness of the joint surfaces prior to brazing is also important. The surface must be free of oxides, oils, and other contaminants to promote wetting and capillary attraction during the process as well as bonding across the entire interface. Chemical treatment like solvent cleaning and mechanical means like wire brushing and sand blasting are used to clean the surface. After cleaning and during brazing, fluxes are used to maintain the cleanliness and promote wetting for capillary action between the faying surfaces. The two main methods for cleaning parts, prior to brazing are chemical cleaning, and abrasive or mechanical cleaning. In the case of mechanical cleaning, it is of vital importance to maintain the

proper surface roughness as wetting on a rough surface occurs much more readily than on a smooth surface of the same geometry. Another important factor is the effect of temperature and time on the quality of brazed joints. As the temperature of the braze alloy is increased, the alloying and wetting action of the filler metal increases as well. In general, the brazing temperature selected must be above the melting point of the filler metal. However, there are several factors that influence the joint designer's temperature selection. The best temperature is usually selected so as to: be the lowest possible braze temperature minimize any heat effects on the assembly keep filler metal/base metal interactions to a minimum maximize the life of any fixtures or jigs used.

In some cases, a higher temperature may be selected to allow for other factors in the design (e.g. to allow use of a different filler metal, or to control metallurgical effects, or to sufficiently remove surface contamination). The effect of time on the brazed joint primarily affects the extent to which the aforementioned effects are present; however, in general most production processes are selected to minimize brazing time and the associated costs. This is not always the case, however, since in some non-production settings, time and cost are secondary to other joint attributes (e.g. strength, appearance).

2.2

Filler metals

The common filler metals used and the base metals for which they are used are given in table 01. A brazing metal should have the following characteristics: Melting temperature must be compatible with the base metals Surface tension is the liquid phase must be low for good wettability Fluidity of the molten metal must be high enough for penetration into the interface The metal must be capable of being brazed into a joint of adequate strength for the application Chemical and physical interactions with the base metal must be avoided

Filler metals for brazing materials are applied in various ways, including wire, rod, sheets, strips, powder, pastes, cream, preformed parts made of braze metals designed to fit a particular configuration, and cladding on one of the surfaces to be brazed. Braze metal pastes consist filler metals powders fixed with fluid fluxes and binders. In general, braze alloys are made up of 3 or more metals to form an alloy with the desired properties. The filler metal for a particular application is chosen based on its ability to: wet the base metals, withstand the service conditions required, and melt at a lower temperature than the base metals or at a very specific temperature. Some brazes come in the form of trifoils, laminated foils of a carrier metal clad with a layer of braze at each side. The centre metal is often copper; its role is to act as a carrier for the alloy, to absorb

mechanical stresses due to e.g. differential thermal expansion of dissimilar materials (e.g. a carbide tip and a steel holder), and to act as a diffusion barrier (e.g. to stop diffusion of aluminium from aluminium bronze to steel when brazing these two). 2.2.1 Braze families

Brazing alloys form several distinct groups; the alloys in the same group have similar properties and uses. Pure metals: Unalloyed. Often noble metals silver, gold, palladium. Ag-Cu: Good melting properties. Silver enhances flow. Eutectic alloy used for furnace brazing. Copper-rich alloys prone to stress cracking by ammonia. Ag-Zn: Similar to Cu-Zn, used in jewellery due to high silver content to be compliant with hallmarking. Colour matches silver. Resistant to ammonia-containing silver-cleaning fluids. Cu-Zn (brass): General purpose, used for joining steel and cast iron. Corrosion resistance usually inadequate for copper, silicon bronze, copper-nickel, and stainless steel. Reasonably ductile. High vapour pressure due to volatile zinc, unsuitable for furnace brazing. Copper-rich alloys prone to stress cracking by ammonia. Ag-Cu-Zn: Lower melting point than Ag-Cu for same Ag content. Combines advantages of Ag-Cu and Cu-Zn. At above 40% Zn the ductility and strength drop, so only lower-zinc alloys of this type are used. At above 25% zinc less ductile copper-zinc and silver-zinc phases appear. Copper content above 60% yields reduced strength and liquidus above 900 C. Silver content above 85% yields reduced strength, high liquidus and high cost. Copper-rich alloys prone to stress cracking by ammonia. Silver-rich brazes (above 67.5% Ag) are hallmarkable and used in jewellery; alloys with lower silver content are used for engineering purposes. Alloys with copper-zinc ratio of about 60:40 contain the same phases as brass and match its colour; they are used for joining brass. Small amount of nickel improves strength and corrosion resistance and promotes wetting of carbides. Addition of manganese together with nickel increases fracture toughness. Addition of cadmium yields Ag-Cu-Zn-Cd alloys with improved fluidity and wetting and lower melting point; however cadmium is toxic. Addition of tin can play mostly the same role. Cu-P: Widely used for copper and copper alloys. Does not require flux for copper. Can be also used with silver, tungsten, and molybdenum. Copper-rich alloys prone to stress cracking by ammonia. Ag-Cu-P: Like Cu-P, with improved flow. Better for larger gaps. More ductile, better electrical conductivity. Copper-rich alloys prone to stress cracking by ammonia. Au-Ag: Noble metals. Used in jewellery. Au-Cu: Continuous series of solid solutions. Readily wet many metals, including refractory ones. Narrow melting ranges, good fluidity. Frequently used in jewellery. Narrow melting range, excellent fluidity. Alloys with 4090% of gold harden on cooling but stay ductile. Nickel improves ductility. Silver lowers melting point but worsens corrosion resistance; to maintain corrosion resistance gold has to be kept above 60%. High-temperature strength and corrosion resistance can be improved by further alloying, e.g. with chromium, palladium, manganese and molybdenum. Addition of vanadium allows wetting ceramics. Low vapour pressure. Au-Ni: Continuous series of solid solutions. Wider melting range than Au-Cu alloys but better corrosion resistance and improved wetting. Frequently alloyed with other metals to reduce proportion of gold while maintaining properties. Copper may be added to lower gold proportion,

chromium to compensate for loss of corrosion resistance, and boron for improving wetting impaired by the chromium. Generally no more than 35% Ni is used, as higher Ni/Au ratios have too wide melting range. Low vapour pressure. Au-Pd: Improved corrosion resistance over Au-Cu and Au-Ni alloys. Used for joining superalloys and refractory metals for high-temperature applications, e.g. jet engines. Expensive. May be substituted for by cobalt-based brazes. Low vapour pressure. Pd: Good high-temperature performance, high corrosion resistance (less than gold), high strength (more than gold). Usually alloyed with nickel, copper, or silver. Forms solid solutions with most metals, does not form brittle intermetallics. Low vapour pressure. Ni: Nickel alloys, even more numerous than silver alloys. High strength. Lower cost than silver alloys. Good high-temperature performance, good corrosion resistance in moderately aggressive environments. Often used for stainless steels and heat-resistant alloys. Embrittled with sulphur and some lower-melting point metals, e.g. zinc. Boron, phosphorus, silicon and carbon lower melting point and rapidly diffuse to base metals; this allows diffusion brazing and allows the joint to be used above the brazing temperature. Borides and phosphides form brittle phases; amorphous preforms can be made by rapid solidification. Co: Cobalt alloys. Good high-temperature corrosion resistance, possible alternative to Au-Pd brazes. Low workability at low temperatures, preforms prepared by rapid solidification. Al-Si: for brazing aluminium. Active alloys: Containing active metals, e.g. titanium or vanadium. Used for brazing nonmetallic materials, e.g. graphite or ceramics.

2.2.2

Role of elements

Silver: Enhances capillary flow, improves corrosion resistance of less-noble alloys, worsens corrosion resistance of gold and palladium. Relatively expensive. High vapour pressure, problematic in vacuum brazing. Wets copper. Does not wet nickel and iron. Reduces melting point of many alloys, including gold-copper. Copper: Good mechanical properties. Often used with silver. Dissolves and wets nickel. Somewhat dissolves and wets iron. Copper-rich alloys sensitive to stress cracking in presence of ammonia. Zinc: Lowers melting point. Often used with copper. Susceptible to corrosion. Improves wetting on ferrous metals and on nickel alloys. Compatible with aluminium. High vapour tension, produces somewhat toxic fumes, requires ventilation; highly volatile above 500 C. At high temperatures may boil and create voids. Prone to selective leaching in some environments, which may cause joint failure. Traces of bismuth and beryllium together with tin or zinc in aluminium-based braze destabilize oxide film on aluminium, facilitating its wetting. High affinity to oxygen, promotes wetting of copper in air by reduction of the cuprous oxide surface film. Less such benefit in furnace brazing with controlled atmosphere. Embrittles nickel. High levels of zinc may result in a brittle alloy. Aluminium: Usual base for brazing aluminium and its alloys. Embrittles ferrous alloys. Gold: Excellent corrosion resistance. Very expensive. Wets most metals. Palladium: Excellent corrosion resistance, though less than gold. Higher mechanical strength than gold. Good high-temperature strength. Very expensive, though less than gold. Makes the joint less prone to fail due to intergranular penetration when brazing alloys of nickel,

molybdenum, or tungsten. Increases high-temperature strength of gold-based alloys. Improves high-temperature strength and corrosion resistance of gold-copper alloys. Forms solid solutions with most engineering metals, does not form brittle intermetallics. High oxidation resistance at high temperatures, especially Pd-Ni alloys. Cadmium: Lowers melting point, improves fluidity. Toxic. Produces toxic fumes, requires ventilation. High affinity to oxygen, promotes wetting of copper in air by reduction of the cuprous oxide surface film. Less such benefit in furnace brazing with controlled atmosphere. Allows reducing silver content of Ag-Cu-Zn alloys. Replaced by tin in more modern alloys. Lead: Lowers melting point. Toxic. Produces toxic fumes, requires ventilation. Tin: Lowers melting point, improves fluidity. Broadens melting range. Can be used with copper, with which it forms bronze. Improves wetting of many difficult-to-wet metals, e.g. stainless steels and tungsten carbide. Traces of bismuth and beryllium together with tin or zinc in aluminium-based braze destabilize oxide film on aluminium, facilitating its wetting. Low solubility in zinc, which limits its content in zinc-bearing alloys. Bismuth: Lowers melting point. May disrupt surface oxides. Traces of bismuth and beryllium together with tin or zinc in aluminium-based braze destabilize oxide film on aluminium, facilitating its wetting. Beryllium: Traces of bismuth and beryllium together with tin or zinc in aluminium-based braze destabilize oxide film on aluminium, facilitating its wetting. Nickel: Strong, corrosion-resistant. Impedes flow of the melt. Addition to gold-copper alloys improves ductility and resistance to creep at high temperatures. Addition to silver allows wetting of silver-tungsten alloys and improves bond strength. Improves wetting of copper-based brazes. Improves ductility of gold-copper brazes. Improves mechanical properties and corrosion resistance of silver-copper-zinc brazes. Nickel content offsets brittleness induced by diffusion of aluminium when brazing aluminium-containing alloys, e.g. aluminium bronzes. In some alloys increases mechanical properties and corrosion resistance, by a combination of solid solution strengthening, grain refinement, and segregation on fillet surface and in grain boundaries, where it forms a corrosion-resistant layer. Extensive intersolubility with iron, chromium, manganese, and others; can severely erode such alloys. Embrittled by zinc, many other low melting point metals, and sulphur. Chromium: Corrosion-resistant. Increases high-temperature corrosion and strength of goldbased alloys. Added to copper and nickel to increase corrosion resistance of them and their alloys. Wets oxides, carbides, and graphite; frequently a major alloy component for hightemperature brazing of such materials. Impairs wetting by gold-nickel alloys, which can be compensated for by addition of boron. Manganese: High vapour pressure, unsuitable for vacuum brazing. In gold-based alloys increases ductility. Increases corrosion resistance of copper and nickel alloys. Improves hightemperature strength and corrosion resistance of gold-copper alloys. Higher manganese content may aggravate tendency to liquation. Manganese in some alloys may tend to cause porosity in fillets. Tends to react with graphite moulds and jigs. Oxidizes easily, requires flux. Lowers melting point of high-copper brazes. Improves mechanical properties and corrosion resistance of silver-copper-zinc brazes. Cheap, even less expensive than zinc. Part of the Cu-Zn-Mn system is brittle, some ratios cannot be used. In some alloys increases mechanical properties and corrosion resistance, by a combination of solid solution strengthening, grain refinement, and segregation on fillet surface and in grain boundaries, where it forms a corrosion-resistant layer. Facilitates wetting of cast iron due to its ability to dissolve carbon.

Molybdenum: Increases high-temperature corrosion and strength of gold-based alloys. Increased ductility of gold-based alloys, promotes their wetting of refractory materials, namely carbides and graphite. When present in alloys being joined, may destabilize the surface oxide layer (by oxidizing and then volatilizing) and facilitate wetting. Cobalt: Good high-temperature properties and corrosion resistance. In nuclear applications can absorb neutrons and build up cobalt-60, a potent gamma radiation emitter. Magnesium: Addition to aluminium makes the alloy suitable for vacuum brazing. Volatile, though less than zinc. Vaporization promotes wetting by removing oxides from the surface, vapours act as getter for oxygen in the furnace atmosphere. Indium: Lowers melting point. Improves wetting of ferrous alloys by copper-silver alloys. Carbon: Lowers melting point. Can form carbides. Can diffuse to the base metal, resulting in higher remelt temperature, potentially allowing step-brazing with the same alloy. At above 0.1% worsens corrosion resistance of nickel alloys. Trace amounts present in stainless steel may facilitate reduction of surface chromium(III) oxide in vacuum and allow fluxless brazing. Diffusion away from the braze increases its remelt temperature; exploited in diffusion brazing. Silicon: Lowers melting point. Can form silicides. Improves wetting of copper-based brazes. Promotes flow. Causes intergranular embrittlement of nickel alloys. Rapidly diffuses into the base metals. Diffusion away from the braze increases its remelt temperature; exploited in diffusion brazing. Germanium: Lowers melting point. Expensive. For special applications. May create brittle phases. Boron: Lowers melting point. Can form hard and brittle borides. Unsuitable for nuclear reactors. Fast diffusion to the base metals. Can diffuse to the base metal, resulting in higher remelt temperature, potentially allowing step-brazing with the same alloy. Can erode some base materials or penetrate between grain boundaries of many heat-resistant structural alloys, degrading their mechanical properties. Has to be avoided in nuclear applications due to its interaction with neutrons. Causes intergranular embrittlement of nickel alloys. Improves wetting of/by some alloys, can be added to Au-Ni-Cr alloy to compensate for wetting loss by chromium addition. In low concentrations improves wetting and lowers melting point of nickel brazes. Rapidly diffuses to base materials, may lower their melting point; especially a concern when brazing thin materials. Diffusion away from the braze increases its remelt temperature; exploited in diffusion brazing. Mischmetal, in amount of about 0.08%, can be used to substitute boron where boron would have detrimental effects. Cerium, in trace quantities, improves fluidity of brazes. Particularly useful for alloys of four or more components, where the other additives compromise flow and spreading. Strontium, in trace quantities, refines the grain structure of aluminium-based alloys. 2.2.3 Deoxidizers Phosphorus: Lowers melting point. Deoxidizer, decomposes copper oxide; phosphorusbearing alloys can be used on copper without flux. Does not decompose zinc oxide, so flux is needed for brass. Forms brittle phosphides with some metals, e.g. nickel (Ni3P) and iron, phosphorus alloys unsuitable for brazing alloys bearing iron, nickel or cobalt in amount above 3%. The phosphides segregate at grain boundaries and cause intergranular embrittlement. (Sometimes the brittle joint is actually desired, though. Fragmentation grenades can be brazed with phosporus bearing alloy to produce joints that shatter easily at detonation.) Avoid in environments with presence of sulfur dioxide (e.g. paper mills) and hydrogen sulfide (e.g.

sewers, or close to volcanoes); the phosphorus-rich phase rapidly corrodes in presence of sulfur and the joint fails. Phosphorus can be also present as an impurity introduced from e.g. electroplating baths. In low concentrations improves wetting and lowers melting point of nickel brazes. Diffusion away from the braze increases its remelt temperature; exploited in diffusion brazing. Lithium: Deoxidizer. Eliminates the need for flux with some materials. Lithium oxide formed by reaction with the surface oxides is easily displaced by molten braze alloy.[17] 2.2.4 Active metals Titanium: Most commonly used active metal. Few percents added to Ag-Cu alloys facilitate wetting of ceramics, e.g. silicon nitride. Most metals, except few (namely silver, copper and gold), form brittle phases with titanium. When brazing ceramics, like other active metals, titanium reacts with them and forms a complex layer on their surface, which in turn is wettable by the silver-copper braze. Wets oxides, carbides, and graphite; frequently a major alloy component for high-temperature brazing of such materials. Zirconium: Wets oxides, carbides, and graphite; frequently a major alloy component for high-temperature brazing of such materials. Hafnium Vanadium: Promotes wetting of alumina ceramics by gold-based alloys. Aluminium: Base component of most brazes for aluminium. Embrittles ferrous metals. 2.2.5 Impurities Sulfur: Compromises integrity of nickel alloys. Can enter the joints from residues of lubricants, grease or paint. Forms brittle nickel sulfide (Ni3S2) that segregates at grain boundaries and cause intergranular failure. Some additives and impurities act at very low levels. Both positive and negative effects can be observed. Strontium at levels of 0.01% refines grain structure of aluminium. Beryllium and bismuth at similar levels help disrupt the passivation layer of aluminium oxide and promote wetting. Carbon at 0.1% impairs corrosion resistance of nickel alloys. Aluminium can embrittle mild steel at 0.001%, phosphorus at 0.01%. In some cases, especially for vacuum brazing, high-purity metals and alloys are used. 99.99% and 99.999% purity levels are available commercially. Care has to be taken to not introduce deletrious impurities from joint contaminations or by dissolution of the base metals during brazing.

2.3

Fluxes

The use of brazing fluxes is somewhat similar to that of welding fluxes. They dissolve, combine with and inhibit the formation of oxides and other unwanted by-products in the brazing process. However, it does not substitute for the cleaning process. Characteristics of a god flux are: Low melting temperature Low viscosity so that it can be displaced by the filler Facilitates wetting Protects joint till the solidification of filler.

The flux should be easily removed after brazing. Common ingredients in brazing fluxes are borax, borates, fluorides, and chlorides. Wetting agents are included to reduce surface tension of the molten filler and to improve wettability. Forms of flux include powders, pastes, and slurries. A vacuum or reducing atmosphere that inhibits the formation of oxides can also be used instead of flux. The flux also serves the purpose of cleaning any contamination left on the brazing surfaces. Flux can be applied in any number of forms including flux paste, liquid, powder or pre-made brazing pastes that combine flux with filler metal powder. Flux can also be applied using brazing rods with a coating of flux, or a flux core. In either case, the flux flows into the joint when applied to the heated joint and is displaced by the molten filler metal entering the joint. Excess flux should be removed when the cycle is completed because flux left in the joint can lead to corrosion, impede joint inspection, and prevent further surface finishing operations. Phosphorus-containing brazing alloys can be self-fluxing when joining copper to copper. Fluxes are generally selected based on their performance on particular base metals. To be effective, the flux must be chemically compatible with both the base metal and the filler metal being used. Self-fluxing phosphorus filler alloys produce brittle phosphides if used on iron or nickel. As a general rule, longer brazing cycles should use less active fluxes than short brazing operations. Table 01: Common filler metals used in brazing base metals on which they can be used Filler metal Aluminium and Silicon Copper Copper and phosphorous Copper and Zinc Gold and Silver Nickel alloys Silver alloys Typical composition 90 Al, 10 Si 99.9 Cu 95 Cu. 5 P 60 Cu. 40 Zn 80 Au, 20 Ag Ni, Cr, others Ag, Cu, Zn, Cd Approximate brazing temperature (oC) 600 1120 850 925 950 1120 730 Base metals Aluminium Nickel, copper Copper Steels, cast irons, nickel Stainless steels, nickel alloys Stainless steels, nickel alloys Titanium, Monel, Inconel, tool steel, nickel

Figure 04: Several techniques for applying filler metal in brazing: (a) torch and filler rod, (b) ring of filler metal at entrance of gap, (c) foil of filler metal between flat part surfaces. Sequence: (1) before, (2) after.

2.4

Atmosphere in brazing

As the brazing work requires high temperatures, oxidation of the metal surface occurs in oxygen-containing atmosphere. This may necessitate use of other environments than air. The commonly used atmospheres are:
Air: Simple and economical. Many materials susceptible to oxidation and buildup of scale. Acid cleaning bath or mechanical cleaning can be used to remove the oxidation after work. Flux tends to be employed to counteract the oxidation, but it may weaken the joint. Combusted fuel gas (low hydrogen, AWS type 1, exothermic generated atmospheres): 87% N2, 1112% CO2, 5-1% CO, 5-1% H2. For silver, copper-phosphorus and copper-zinc filler metals. For brazing copper and brass. Combusted fuel gas (decarburizing, AWS type 2, "endothermic generated atmospheres"): 7071% N2, 56% CO2, 910% CO, 1415% H2. For copper, silver, copper-phosphorus and copper-zinc filler metals. For brazing copper, brass, nickel alloys, Monel, medium carbon steels. Combusted fuel gas (dried, AWS type 3, "endothermic generated atmospheres"): 7375% N2, 1011% CO, 1516% H2. For copper, silver, copper-phosphorus and copper-zinc filler metals. For brazing copper, brass, low-nickel alloys, Monel, medium and high carbon steels.

Combusted fuel gas (dried, decarburizing, AWS type 4): 4145% N2, 1719% CO, 3840% H2. For copper, silver, copper-phosphorus and copper-zinc filler metals. For brazing copper, brass, low-nickel alloys, medium and high carbon steels. Ammonia (AWS type 5): Dissociated ammonia (75% hydrogen, 25% nitrogen) can be used for many types of brazing and annealing. Inexpensive. For copper, silver, nickel, copperphosphorus and copper-zinc filler metals. For brazing copper, brass, nickel alloys, Monel, medium and high carbon steels and chromium alloys. Nitrogen+hydrogen, cryogenic or purified (AWS type 6A): 7099% N2, 130% H2. For copper, silver, nickel, copper-phosphorus and copper-zinc filler metals. Nitrogen+hydrogen+carbon monoxide, cryogenic or purified (AWS type 6B): 7099% N2, 220% H2, 110% CO. For copper, silver, nickel, copper-phosphorus and copper-zinc filler metals. For brazing copper, brass, low-nickel alloys, medium and high carbon steels. Nitrogen, cryogenic or purified (AWS type 6C): Non-oxidizing, economical. At high temperatures can react with some metals, e.g. certain steels, forming nitrides. For copper, silver, nickel, copper-phosphorus and copper-zinc filler metals. For brazing copper, brass, low-nickel alloys, Monel, medium and high carbon steels. Hydrogen (AWS type 7): Strong deoxidizer, highly thermally conductive. Can be used for copper brazing and annealing steel. May cause hydrogen embrittlement to some alloys. For copper, silver, nickel, copper-phosphorus and copper-zinc filler metals. For brazing copper, brass, nickel alloys, Monel, medium and high carbon steels and chromium alloys, cobalt alloys, tungsten alloys, and carbides. Inorganic vapors (various volatile fluorides, AWS type 8): Special purpose. Can be mixed with atmospheres AWS 15 to replace flux. Used for silver-brazing of brasses. Noble gas (usually argon, AWS type 9): Non-oxidizing, more expensive than nitrogen. Inert. Parts must be very clean, gas must be pure. For copper, silver, nickel, copper-phosphorus and copper-zinc filler metals. For brazing copper, brass, nickel alloys, Monel, medium and high carbon steels chromium alloys, titanium, zirconium, hafnium. Noble gas+hydrogen (AWS type 9A) Vacuum: Requires evacuating the work chamber. Expensive. Unsuitable (or requires special care) for metals with high vapor pressure, e.g. silver, zinc, phosphorus, cadmium, and manganese. Used for highest-quality joints, for e.g. aerospace applications.

2.5

Brazing methods

There are various methods used in brazing. The brazing processes are differentiated by the heat sources used. 2.5.1 Torch brazing

In torch brazing, flux is applied to the part surfaces and a torch is used to direct a flame against the work in vicinity of the joint. A reducing flame is typically used to inhibit oxidation. A reducing flame is usually used to inhibit oxidation. After the workparts areas are heated to a suitable temperature, a

filler is added, usually in rod or wire form. Fuels used include acetylene, propane, and other gases, with air or oxygen. The selection of the mixture depends on the heating requirements of the job. Torch brazing is usually done manually. Skilled worker must be employed to control the flame, manipulate hand-held torches, and properly judge the temperatures. Repair work is a common application. It can also be used in mechanised production operations, in which parts and brazing metal are loaded onto a conveyor or indexing table and passed under one or more torches. It is by far the most common method of mechanized brazing in use. It is best used in small production volumes or in specialized operations, and in some countries, it accounts for a majority of the brazing taking place. There are three main categories of torch brazing in use: manual, machine, and automatic torch brazing.

Manual torch brazing: this is a procedure where the heat is applied using a gas flame placed on or near the joint being brazed. The torch can either be hand held or held in a fixed position depending on if the operation is completely manual or has some level of automation. Manual brazing is most commonly used on small production volumes or in applications where the part size or configuration makes other brazing methods impossible. The main drawback is the high labor cost associated with the method as well as the operator skill required to obtain quality brazed joints. The use of flux or self-fluxing material is required to prevent oxidation. Machine torch brazing: This is commonly used where a repetitive braze operation is being carried out. This method is a mix of both automated and manual operations with an operator often placing brazes material, flux and jigging parts while the machine mechanism carries out the actual braze. The advantage of this method is that it reduces the high labor and skill requirement of manual brazing. The use of flux is also required for this method as there is no protective atmosphere, and it is best suited to small to medium production volumes. Automatic torch brazing: This is a method that almost eliminates the need for manual labor in the brazing operation, except for loading and unloading of the machine. The main advantages of this method are: a high production rate, uniform braze quality, and reduced operating cost. The equipment used is essentially the same as that used for Machine torch brazing, with the main difference being that the machinery replaces the operator in the part preparation. 2.5.2 Furnace brazing

This method uses a furnace to supply the necessary heat and is best suited to medium and high production. In medium production, the component parts and brazing metal are loaded to the furnace in batches. Then it is heated to brazing temperature, cooled, and removed. High production uses flowthrough furnaces, in which parts are placed on a conveyor and are transported through the various cooling and heating sections. Temperature and atmospheric control are important the atmosphere must be neutral or reducing. Sometimes, vacuum furnaces are used. Depending on the atmosphere, flux may be eliminated.

Figure 05: Furnace brazing schematic

Furnace brazing is a semi-automatic process used widely in industrial brazing operations due to its adaptability to mass production and use of unskilled labor. There are many advantages of furnace brazing over other heating methods that make it ideal for mass production. One main advantage is the ease with which it can produce large numbers of small parts that are easily jigged or self-locating. The process also offers the benefits of a controlled heat cycle (allowing use of parts that might distort under localized heating) and no need for post braze cleaning. Common atmospheres used include: inert, reducing or vacuum atmospheres all of which protect the part from oxidation. Some other advantages include the low unit cost when used in mass production, close temperature control, and the ability to braze multiple joints at once. Furnaces are typically heated using either electric, gas or oil depending on the type of furnace and application. However, some of the disadvantages of this method include high capital equipment cost, more difficult design considerations and high power consumption. There are four main types of furnaces used in brazing operations: batch type; continuous; retort with controlled atmosphere; and vacuum. Batch type: These furnaces have relatively low initial equipment costs and heat each part load separately. It is capable of being turned on and off at will which reduces operating expenses when not in use. These furnaces are well suited to medium to large volume production and offer a large degree of flexibility in type of parts that can be brazed. Either controlled atmospheres or flux can be used to control oxidation and cleanliness of parts. Continuous type: furnaces are best suited to a steady flow of similar-sized parts through the furnace. These furnaces are often conveyor fed, allowing parts to be moved through the hot zone at a controlled speed. It is common to use either controlled atmosphere or pre-applied flux in continuous furnaces. In particular, these furnaces offer the benefit of very low manual labor requirements and so are best suited to large scale production operations. Retort-type: furnaces differ from other batch-type furnaces in that they make use of a sealed lining called a "retort". The retort is generally sealed with either a gasket or is welded shut and filled completely with the desired atmosphere and then heated externally by conventional heating elements. Due to the high temperatures involved, the retort usually made of heat resistant alloys that resist oxidation. Retort furnaces are often either used in a batch or semi-continuous versions. Vacuum furnaces: This is a relatively economical method of oxide prevention and is most often used to braze materials with very stable oxides (aluminum, titanium and zirconium) that cannot be brazed in atmosphere furnaces. Vacuum brazing is also used heavily with refractory materials and other exotic alloy combinations unsuited to atmosphere furnaces. Due to the absence of flux or a reducing atmosphere, the part cleanliness is critical when brazing in a vacuum. The three main types of vacuum furnace are: single-wall hot retort, double-walled hot retort, and cold-wall retort. Typical vacuum levels for brazing range from pressures of 1.3 to 0.13 Pascals to 0.00013 Pa or lower. Vacuum furnaces are most commonly batch-type, and they are suited to medium and high production volumes. 2.5.3 Induction brazing

This uses the heat generated due to the resistance to a high frequency current induced in the work. The parts are preloaded with filler and placed in a high frequency AC field, with no contact with the induction coil. The frequencies range from 5 5000 kHz. High frequency power source tend to

produce surface heating while lower frequencies cause deeper heat penetration. Low frequencies are more appropriate for heavier sections. Induction brazing can be used for low to high production. 2.5.4 Resistance brazing

The required heat is obtained by the resistance to the flow of an electric current through the part. The parts are directly connected to the current in this case. The equipment is similar to that used for induction brazing. A lower power level is required. The parts with filler metal are preplaced and held between electrodes while pressure and current are applied. Both induction and resistance brazing produce rapid heating and are used for relatively small parts. Induction brazing is more widely used. 2.5.5 Dip brazing

Heating is done using a molten metal bath or a molten salt bath. In both methods, the assembled parts are immersed in baths contained in a heating pot. Solidification occurs when parts are removed from the bath. In the salt bath method, the molten mixture contains the flux and filler is preloaded into the assembly. In the metal bath method, the molten filler metal is the heating medium. It is drawn into the joint by capillary action during submersion. Dip brazing achieves fast cycles and can be used to braze many joints simultaneously. 2.5.6 Infrared brazing

This uses heat from a high-intensity infrared lamp. Some lamps can generate up to 5000 W of radiant energy, which is directed at the workpart for brazing. The process is slower than most of the other processes mentioned and is generally limited to thin sections. 2.5.7 Braze welding

This process differs by the type of joint to which it is applied. Braze welding is often used for a more conventional weld joint (e.g., V-joint). A greater quantity of filler is deposited when compared to other types of brazing. No capillary action occurs. The joint consists entirely of filler metal. The base metal does not melt, and therefore it is not fused into the joint as in a conventional fusion welding process. The main application is repair work. A bronze or brass filler rod coated with flux is generally used to join steel workpieces. The equipment needed for braze welding is basically identical to the equipment used in brazing. Since braze welding usually requires more heat than brazing, acetylene or methylacetylene-propadiene (MPS) gas fuel is commonly used. The American Welding Society states that the name comes from the fact that no capillary action is used. Braze welding has many advantages over fusion welding. It allows the joining of dissimilar metals, minimization of heat distortion, and can reduce the need for extensive pre-heating. Additionally, since the metals joined are not melted in the process, the components retain their original shape; edges and contours are not eroded or changed by the formation of a fillet. Another side effect of braze welding is the elimination of stored-up stresses that are often present in fusion welding. This is extremely important in the repair of large castings. The disadvantages are the loss of strength when subjected to high temperatures and the inability to withstand high stresses. Carbide, cermet and ceramic tips are plated and then joined to steel to make tipped band saws. The plating acts as a braze alloy.

Figure 06: A braze welded T-joint

Figure 07: Braze welding. The joint consists of braze (filler) metal, no base metal is fused in the joint.

3.0

Soldering

This is similar to brazing and can be defined as a joining process in which the filler metal has a melting point not exceeding 450 oC is melted and distributed by capillary action between the faying surfaces of the metal parts being joined. No melting of the base metal occurs, but the filler metal wets and combines with the base metal to form a metallurgical bond. It is very similar to brazing and the heating methods used are similar. Surfaces to be soldered must be precleaned so that they are free of oxides, oils, etc. An appropriate flux is applied to the faying surfaces and the surfaces are heated. The filler metal, called solder, is added to the joint, which distributes itself between the parts. In some applications, solder is precoated to one or both surfaces. This is called tinning (though the solder may not contain tin). Typical clearances range from 0.75 to 0.125 mm except when surfaces are tinned. Then the clearance is about 0.025 mm. After solidification, the flux residue is removed. As an industrial process, soldering is most closely associated with electronics assembly. It is also used for mechanical joints that are not subjected to elevated stresses or temperatures. Advantages of soldering are: Low energy input relative to brazing and fusion welding Variety of heating methods available Good electrical and thermal conductivity of the joint Capability to make air-tight and liquid-tight seams for containers Easy to repair and rework

The main disadvantages are Low joint strength unless reinforced by mechanical means Possible weakening and melting of joint in elevated temperature service

3.1

Joint designs in soldering

The joints types in soldering are also limited to butt and lap joints. Butt joints should not be used for load bearing applications. Some brazing applications of these joints apply to soldering. In addition, there are some variations peculiar to soldering to cope with special part geometries that arise in electrical connections. In soldered mechanical joints of sheet metal parts, the edges of the sheets are often bent over and interlocked before soldering to increase joint strength. For electronics applications, the main use of soldering is to provide an electrically conductive path between the two parts being joined. Other design considerations in these joints include heat generation (from the resistance of the joint) and vibration. Mechanical strength of a soldered electrical connection is achieved by deforming one or both metal parts to accomplish a mechanical joint between them or by making the surface area larger to provide maximum support by the solder.

Figure 08: Mechanical interlocking in soldered joints for increased strength (a) flat lock seam, (b) bolted or riveted joint, (c) copper pipe fittings lap cylindrical joint, (d) crimping (forming) of cylindrical tap joint

Figure 09: Techniques for securing the joint by mechanical means prior to soldering in electrical connections: (a) crimped lead wire on PC board, (b) plated through-hole on PC board to maximise solder contact surface, (c) Hooked wire on flat terminal, (d) twisted wires

3.2

Solders and fluxes

Solders and fluxes are both used in soldering. Both are critically important in the joining process. 3.2.1 Solders

Most solders are alloys of lead and tin, as both have low melting temperatures. The alloys posses a wide range of liquidus and solidus temperatures to achieve a good control of the soldering process for a variety of applications. Lead is poisonous and its percentage is reduced as much as possible in solders. Tin is chemically active at soldering temperatures and promotes wetting required for successful joining. In soldering copper, intermetallic compounds of copper and tin are formed that strengthen the bond. Silver and antimony are also used sometimes. Lead free solders are becoming more important. Table 02 lists various solders.

Figure 10: Lead-tin phase diagram An eutectic formulation has several advantages for soldering; chief among these is the coincidence of the liquidus and solidus temperatures, i.e. the absence of a plastic phase. This allows for quicker wetting as the solder heats up, and quicker setup as the solder cools. A non-eutectic formulation must remain still as the temperature drops through the liquidus and solidus temperatures. Any differential movement during the plastic phase may result in cracks, giving an unreliable joint. Additionally, a eutectic formulation has the lowest possible melting point, which minimizes heat stress on electronic components during soldering. Lead-free solders are suggested anywhere young children may come into contact with (since young children are likely to place things into their mouths), or for outdoor use where rain and other

precipitation may wash the lead into the groundwater. Lead-free solder alloys melt around 250 C (482 F), depending on their composition. For environmental reasons, 'no-lead' solders are becoming more widely used. Unfortunately most 'no-lead' solders are not eutectic formulations, making it more difficult to create reliable joints with them. See complete discussion below; see also RoHS. Other common solders include low-temperature formulations (often containing bismuth), which are often used to join previously-soldered assemblies without un-soldering earlier connections, and hightemperature formulations (usually containing silver) which are used for high-temperature operation or for first assembly of items which must not become unsoldered during subsequent operations. Alloying silver with other metals changes the melting point, adhesion and wetting characteristics, and tensile strength. Of all the brazing alloys, the silver solders have the greatest strength and the broadest applications. Specialty alloys are available with properties such as higher strength, better electrical conductivity and higher corrosion resistance. Table 02: Some common solder alloy compositions with their melting temperatures and applications Filler metal Lead-Silver Tin-Antimony Tin-Lead Approximate composition 96 Pb, 4 Ag 95 Sn, 5 Sb 63 Sn, 37 Pb 60 Sn, 40 Pb 50 Sn, 50 Pb 40 Sn, 60 Pb 96 Sn, 4 Ag 91 Sn, 9 Zn 95.5 S, 3.9 Ag, 0.6 Cu Approximate melting temperature (oC) 305 238 183 188 199 207 221 199 217 Principal Application Elevated temperature joints Plumbing and heating Electrical/ electronics Electrical, electronics General purpose Automobile radiators Food containers Aluminium joining Electronics surface mount technology

Tin-Silver Tin-Zinc Tin-Silver-Copper

3.2.2

Soldering fluxes

Soldering fluxes should be: Molten at soldering temperatures Removes oxide films and tarnish from the base part surfaces Prevent oxidation during heating Promote wetting of faying surfaces Be readily displaces by molten solder during the process Leave a residue that is noncorrosive and non-conductivity.

There is no single flux which meets all these conditions for all combinations of solder and base metals. The flux formulation must be selected for a given application. Soldering fluxes can be classified as organic and inorganic. Organic fluxes are made of either rosin (i.e., natural rosin such as gum wood, which is not water soluble) or water soluble ingredients (e.g., alcohols, organic acids, and halogenated salts). The water soluble type facilitates cleanup after

soldering. Organic fluxes are commonly used for electrical and electronic connections. They are generally chemically reactive at elevated temperatures but relatively non-corrosive at room temperatures. Inorganic fluxes consist of inorganic acids (e.g., muriatic acid) and salt (e.g., combination of zinc and aluminium oxides). They are used to achieve rapid and active fluxing where oxide films are a problem. The salts become active when melted, but are less corrosive than acids. When solder is purchased with an acid core it is in this category. Both types of fluxes must be removed after soldering, but it is more important in the case of inorganic acids to prevent continued corrosion of the surfaces. Flux removal is usually accomplished using water solution except in the case of rosins where chemical solvents are required.Recently, preference has been given to water soluble solvents as chemical solvents used for cleaning are harmful to the environment and to humans.

3.3

Soldering methods

Many soldering methods are similar to the methods used in brazing, except that less heat and lower temperatures are required for soldering. These methods include torch soldering, furnace soldering, resistance soldering, dip soldering, and infrared soldering. There are also methods specific to soldering (i.e., not used in brazing) hand soldering, wave soldering, and reflow soldering. 3.3.1 Hand soldering

Hot soldering is performed manually using a hot soldering iron. A bit, made of copper, is used as the working end of the soldering iron. Its functions are Deliver heat to the parts being soldered To melt the solder To convey molten solder to the joint To withdraw excess solder

Most modern soldering irons use resistance heating. Some are designed as fast heating soldering guns which are used in electronic assembly for intermittent operation actuated by a trigger. They are capable of making a solder joint in about a second. 3.3.2 Wave soldering

This is a mechanised technique that allows soldering multiple leads into a PCB as it passes over a wave of molten solder. Typically components are placed on the PCB with their leads sticking out through the holes in the board. This is loaded onto a conveyor for transport through the wavesoldering equipment. The conveyor supports the PCB on its sides so that the underside is exposed to the processing steps. These are: Flux is applied using any of the several methods, including forming, spraying, or brushing Preheating (using light bulbs, heating coils, and infrared devices) to evaporate solvents, activate the flux and raise the temperature of the assembly Wave soldering, in which liquid solder is pumped from a molten bath through a slit onto the bottom of the board to make the soldering connections between the lead wires and metal

circuit on the board. The board is often slightly tilted and a thinning oil is added to the solder to lower its surface tension. Both these processes prevent the deposition of excess solder and the formation of icicles on the bottom of the board. Wave soldering is widely applied in electronics to produce printed circuit board assemblies.

Figure 11: Wave soldering, in which molten solder is delivered up through a narrow slot onto the underside of the PCB to connect the component lead wires.

3.3.3

Reflow soldering

This is also widely used to assemble surface mounted components onto PCBs. In this case, s solder paste consisting of solder powders and a flux binder is applied to spots on the board where electrical contacts are to be made between the surface mount components, and the copper circuit. The components are then placed on the paste spots. And the board is heated to melt the solder. This forms electrical and mechanical bonds between the components and the copper circuit board. Heating methods used include vapour phase reflow and infrared reflow. In vapour phase reflow, an inert fluorinated hydrocarbon is vaporised by heating in an oven. This vapour condenses on the board, transferring the latent heat to the board. In infrared reflow soldering, heat from an infra-red lamp is used to melt the paste to form joints between components leads and circuit areas on the board. Additional heating methods to reflow the solder paste include the use of hot plates, hot air, and lasers. 3.3.4 Induction soldering

Induction soldering is a process which is similar to brazing. The source of heat in induction soldering is induction heating by high-frequency AC current. Generally copper coils are used for the induction heating. This induces currents in the part being soldered. The coils are usually made of copper or a copper base alloy. The copper rings can be made to fit the part needed to be soldered for precision in the work piece. Induction soldering is a process in which a filler metal (solder) is placed between the faying surfaces of (to be joined) metals. The filler metal in this process is melted at a fairly low temperature. Fluxes are commonly used in induction soldering. This is a process which is particularly suitable for soldering continuously. The process is usually done with coils that wrap around a cylinder/pipe that needs to be soldered.

Figure 12: A tube of multicore electronics solder used for manual soldering

Figure 13: soldered copper pipes

3.4

Soldering defects

Various problems may arise in the soldering process which lead to joints which are non functional either immediately or after a period of use. The most common defect when hand-soldering results from the parts being joined not exceeding the solder's liquidus temperature, resulting in a "cold solder" joint. This is usually the result of the soldering iron being used to heat the solder directly, rather than the parts themselves. Properly done, the iron heats the parts to be connected, which in turn melt the solder, guaranteeing adequate heat in the joined parts for thorough wetting. In 'electronic' hand soldering solder the flux is embedded in the solder. Therefore heating the solder first may cause the flux to evaporate before it cleans the surfaces (pcb pad and component connection) being soldered.

Figure 14: An improperly soldered 'cold' joint An improperly selected or applied flux can cause joint failure, or if not properly cleaned off the joint, may corrode the metals in the joint over time and cause eventual joint failure. Without flux the joint may not be clean, or may be oxidized, resulting in an unsound joint. In electronics non-corrosive fluxes are often used. Therefore cleaning flux off may merely be a matter of aesthetics or to make visual inspection of joints easier in specialised 'mission critical' applications such as medical devices, military and aerospace i.e. satellites. For satellites also to reduce weight slightly but usefully. In some conditions i.e. high humidity, even non-corrosive flux might remain slightly active, therefore the flux may be removed to absolutely negate the possibility of corrosion

over time. In some applications, the PCB might also be coated in some form of protective material such as a lacquer to protect it and/or exposed solder joints from the environment. Movement of metals being soldered before the solder has cooled will cause a highly unreliable cracked joint. In electronics' soldering terminology this is known as a 'dry' joint. It has a characteristically dull or grainy appearance immediately after the joint is made, rather than being smooth, bright and shiny. This appearance is caused by crystallization of the liquid solder. A dry joint is weak mechanically and a poor conductor electrically. In general a good looking soldered joint is a good joint. As mentioned it should be smooth, bright and shiny. If not smooth i.e. lumps or balls of otherwise shiny solder the metal has not 'wetted' properly. Not being bright and shiny suggests a weak 'dry' joint. In electronics a 'concave' fillet is ideal. This indicates good wetting and minimal use of solder (therefore minimal heating of heat sensitive components). A joint may be good, but if a large amount of unnecessary solder is used then more heating is obviously required. Excessive heating of a PCB may result in 'delamination', the copper track may actually lift off the board, particularly on single sided PCBs without 'through hole' plating. In the joining of copper tube, failure to properly heat and fill a joint may lead to a 'void' being formed. This is usually a result of improper placement of the flame. If the heat of the flame is not directed at the back of the fitting cup, and the solder wire applied 180 degrees opposite the flame, then solder will quickly fill the opening of the fitting, trapping some flux inside the joint. This bubble of trapped flux is the void; an area inside a soldered joint where solder is unable to completely fill the fittings' cup, because flux has become sealed inside the joint, preventing solder from occupying that space.

Figure 15: Broken solder joints on a Flyback transformer

4.0

Adhesive bonding

Adhesive bonding is a joining process in which a filler metal is used to hold two or more closely placed parts together by surface attachment. The filler metal that bind the components together is the adhesive. It is non-metallic generally polymeric. The parts being joined are generally known as adherents. The adhesives of the greatest interest in engineering is structural adhesives. These are capable of forming strong, permanent joints between strong, rigid adherents. A large number of commercially available adhesives are cured by a variety of mechanisms best suited to the bonding of various materials. Curing refers to the process by which the adhesive is converted from liquid to a solid, usually by chemical reaction to accomplish surface attachment of the parts. The chemical reaction may involve polymerisation, condensation or vulcanisation. Curing is often motivated by heat and/or a catalyst. Pressure is applied to activate the process. If heat is required, the curing

temperatures are relatively low, so the materials being joined are usually unaffected. This is an advantage. Curing or hardening of the adhesive takes time, usually called curing time or setting time. In some cases, this is significant, which is usually a disadvantage in manufacturing. The joint strength in adhesive bonding is determined by the strength of the adhesive and the strength of the attachment between the adhesive and each of the adherents. One of the criteria often used to define a satisfactory adhesive joint is that if a failure occurs, it should occur in one of the adherents instead of happening at the interface or in the adhesive. The strength results from several mechanisms, all of which are dependent on the adhesive and the adherents: Chemical bonding, in which the adhesive unites with the adherents and forms a primary chemical bond upon hardening Physical interactions, in which secondary bonding forces result between the atoms of opposing surfaces Mechanical interlocking, in which the surface roughness of the adherents causes the hardened adhesive to become entangled or trapped in its microscopic surface asperities.

For these mechanisms to operate with the best results, the following conditions must exist: The surfaces of the adherents must be clean (free of dirt, oil, and oxide films that would interfere with achieving intimate contact between adhesive and adherent). Special preparation is usually required. The adhesive in its initial liquid form must achieve through wetting of the adherent surface It is usually helpful for the surfaces to be other than perfectly smooth. A slightly roughened surface increases the effective contact area and promotes mechanical interlocking.

Also, the joint must be designed to exploit the strengths of adhesive bonding and to avoid its drawbacks.

4.1

History

Adhesives have been used since the ancient times and were probably one of the first permanent joining methods. The oldest known adhesive, dated to approximately 200,000 BC, is from spear stone flakes glued to a wood with birch-bark-tar, which was found in central Italy. The use of compound glues to haft stone spears into wood dates back to round 70,000 BC. Evidence for this has been found in Sibudu Cave, South Africa and the compound glues used were made from plant gum and red ochre. The Tyrolean Iceman had weapons fixed together with the aid of glue. 6000-year-old ceramics show evidence of adhesives based upon animal glues made by rendering animal products such as horse teeth. During the times of Babylonia, tar-like glue was used for gluing statues. The Egyptians made much use of animal glues to adhere furniture, ivory, and papyrus. The Mongols also used adhesives to make their short bows, and the Native Americans of the eastern United States used a mixture of spruce gum and fat as adhesives to fashion waterproof seams in their birchbark canoes. Carvings 3300 years old show a glue pot and brush for gluing veneer to wood planks. Ancient Egyptians used the gum from the acacia tree for various assembly and sealing purposes. Bitumen

(from asphalt) was used since ancient times as a cement and mortar for construction in Asia Minor. Pine wood tar and beeswax was used by Romans caulk their ships. Glues derived from fish, stag horns, and cheese were used in early centuries after Christ for assembling components of wood. Recently, adhesives have become an important joining process. Plywood (which relies of adhesives to bond multiple layers of wood) was developed around 1900 Phenol formaldehyde was the first synthetic adhesive developed (around 1910) and its primary use was in bonding wood products such as plywood. During world war II, phenolic resins were developed for adhesive bonding of certain aircraft components, In the 1950s, epoxies were first formulated. Since the 1950s, a variety of adhesives have been produced, including anaerobics, various new polymers, and second-generation acrylics. Today, adhesives are used in a wide range of bonding and sealing applications for joining similar and dissimilar metals, plastics, ceramics, wood, paper, and cardboard. Although it is a well-established joining technique, it is considered a growth area among assembly technologies because of its tremendous opportunities for increased applications.

4.2

Joint strength

The strength of the joint is generally not as high as in welding, brazing, or soldering. Therefore, consideration must be given to the design of the joints that are adhesively bonded. The following design principles apply: Joint contact area should be maximised. Adhesive joints are strongest in shear and tension. Joints should be designed so that the applied stresses belong to one of the two types mentioned above. Adhesive bonds are weakest in cleavage and peeling and the joint should be designed to avoid these.

F igure 10: Types of stresses that must be considered in adhesive bonded joints: (a) tension, (b) shear, (c) cleavage, (d) peeling Typical joint designs adhesive bonding that illustrate these are shown in figure 12. Some joint designs also combine adhesive bonding with another joining method to increase strength and/or to seal the joint. For example, the combination of adhesive bonding and spot welding is called weld bonding. In addition to the mechanical configuration of a joint, the application must be selected so that the physical and chemical properties of the adherents are compatible under the service conditions to which the assembly will be subjected. Adherents include metals, ceramics, glass, plastics, wood, rubber, leather, cloth, paper, and cardboard. This list includes materials that are rigid and flexible,

porous and non-porous, metallic and non-metallic, and that similar or dissimilar substances can be bonded together.

Fig ure 13: Some joint designs for adhesive bonding: (a) through (d): butt joints, (e) and (f): T-joints, (g) through (j): corner joints.

Figure 14: Adhesive bonding combined with other joining methods: (a) weld bonding spot welded and adhesive bonded, (b) riveted (or bolted) and adhesive bonded, (c) formed plus adhesive bonded

4.3

Failure of adhesive joints

There are several factors that could contribute to the failure of two adhered surfaces. Sunlight and heat may weaken the adhesive. Solvents can deteriorate or dissolve adhesive. Physical stresses may also cause the separation of surfaces. When subjected to loading, debonding may occur at different locations in the adhesive joint. The major fracture types are the following: 4.3.1 Cohesive fracture

Cohesive fracture is obtained if a crack propagates in the bulk polymer which constitutes the adhesive. In this case the surfaces of both adherents after debonding will be covered by fractured adhesive. The crack may propagate in the centre of the layer or near an interface. For this last case, the cohesive fracture can be said to be cohesive near the interface. Most quality control standards consider a good adhesive bond to be cohesive. 4.3.2 Interfacial fracture

The fracture is adhesive or interfacial when debonding occurs between the adhesive and the adherent. In most cases, the occurrence of interfacial fracture for a given adhesive goes along with a smaller fracture toughness. The interfacial character of a fracture surface is usually to identify the precise location of the crack path in the interphase.

4.3.3

Other types of fracture

Other types of fracture include: The mixed type, which occurs if the crack propagates at some spots in a cohesive and in others in an interfacial manner. Mixed fracture surfaces can be characterised by a certain percentage of adhesive and cohesive areas. The alternating crack path type which occurs if the cracks jumps from one interface to the other. This type of fracture appears in the presence of tensile pre-stresses in the adhesive layer. Fracture can also occur in the adherent if the adhesive is tougher than the adherent. In this case the adhesive remains intact and is still bonded to one substrate and remnants of the other. For example, when one removes a price label, adhesive usually remains on the label and the surface. This is cohesive failure. If, however, a layer of paper remains stuck to the surface, the adhesive has not failed. Another example is when someone tries to pull apart Oreo cookies and all the filling remains on one side; this is an adhesive failure, rather than a cohesive failure

Figure 15: Failure of the adhesive joint can occur in different locations

4.4

Adhesive types

There is a large number of commercial adhesives available. They fall into three main groups natural, inorganic, and synthetic. 4.4.1 Natural adhesives

These are derived from natural sources (e.g., plants and animals), including gums, starch, dextrin, soy flour, and collagen. This type of adhesive is typically limited to low-stress applications, such as cardboard cartons, furniture, and bookbinding. Also, it is used where large surface areas are present, like plywood. 4.4.2 Inorganic adhesives

These are based mains on sodium silicate and magnesium oxyxhloride.They are relatively low in cost and, but also low in strength, which is a serious limitation as a structural adhesive.

4.4.3

Synthetic adhesives

These are the most important group in manufacturing, They include a variety of thermosetting and thermoplastic polymers. They are cured by various mechanisms: Mixing a catalyst or reactive ingredient with the polymer immediately prior to application Heating to initiate the chemical reaction Radiation curing, such as ultraviolet light Curing by evapouration of water from the liquid or paste adhesive

Some synthetic adhesives are also applied as films or as pressure-sensitive coatings on the surface of surface of the adherents.

Table 03: Important synthetic adhesives Adhesive Anaerobic Modified acrylics Cyanoacrylate Epoxy Description and applications Single-component, thermosetting, acrylic-based. Cures by free radical mechanism at room temperature. Applications: sealant, structural assembly. Two component thermoset, consisting of acrylic-based resin and initiator/hardener. Cures at room temperature after mixing. Applications: fibreglass in boats, sheet metal in cars and aircraft Single-component, thermosetting acrylic based that cures at room temperature on alkaline surfaces. Applications: rubber to plastic, electronic components on circuit boards, plastic and metal cosmetic cases. Includes a variety of widely used adhesives formulated from epoxy resins, curing agents, and filler/modifiers that harden upon mixing. Some are cured when heated. Applications: Aluminium bonding application and honeycomb panels for aircraft, sheet metal reinforcements for cars, lamination of wooden beams, seals in electronics. Single-component, thermoplastic adhesive which hardens from molten state after cooling from elevated temperatures. Formulated from thermoplastic polymers including ethylene vinyl acetate, polyethylene, styrene block copolymer, butyl rubber, polyamide, polyurethane, and polyester. Applications: packaging (e.g., cartons, labels), furniture, footwear, bookbinding, carpeting, and assemblies in appliances and cars. Usually one component in solid form that possesses high tackiness resulting in bonding when pressure is applied. Formed from various polymers of high molecular weight. Can be single sided or double sided. Applications: solar panels, electronic assemblies, plastics to wood and metals One or two components, thermosetting liquid, based on silicone polymers. Curing by room temperature vulcanisation to rubbery solid. Applications: seals in cars (e.g., windshields), electronic seals and insulation, gaskets, bonding of plastics. One or two components, thermosetting, based on urethane polymers. Applications: boning of fibreglass and plastics

Hot melt

Pressure sensitive tapes and films Silicone Urethane

Figure 16: A glue gun, an example of a hot melt adhesive

4.5

Adhesive application technology

Industrial applications of adhesive bonding are widespread and growing. Major uses are automotive, aircraft, building products, and packaging industries. Other industries include footwear, furniture, bookbinding, electrical, and shipbuilding. Table 03 shows some specific applications of synthetic adhesives. 4.5.1 Surface preparation

For adhesive bonding to succeed, part surfaces must be extremely clean. The strength of the bond depends on the degree of adhesion between the adhesive and the adherent. This depends of the cleanliness of the surface. In most cases, additional processing steps are required for cleaning and surface preparation, the methods varying with different adherent materials. For metals, solvent wiping is often used for cleaning, and abrading the surface by sand blasting or other processes usually improves adhesion. For non-metallic parts, solvent cleaning is generally used. The surfaces are sometimes mechanically abraded or chemically etched to increase roughness. It is desirable to accomplish the adhesive bonding process as soon as possible after treatments, since surface oxidation and dirt accumulation increase with time. 4.5.2 Application methods

The actual application of the adhesive to the part surfaces can be achieved in a number of ways: Brushing: performed manually. Uses a stiff-bristled brush. Coatings are often uneven. Flowing: using manually operated pressure-fed flow guns. Has more consistent control than brushing. Manual rollers: similar to paint roller, are used to apply adhesive from a flat container. Silk screening: involves brushing the adhesive through the open areas of the screen onto the part surface, so only selected areas are coated. Spraying: uses an air driven (or airless) spray gun for application over large or difficult to reach areas. Automatic applications: include various automatic dispensers and nozzles for use on medium to high speed production applications. Roll coating: is a mechanised technique in which a rotating roller is partially submersed in a pan of liquid adhesive and picks up a coating of the adhesive, which is then transferred to the work surface. Figure 18 shows one possible application, in which the work is a thin, flexible

material (e.g., paper, cloth, leather, and plastic). Variations of the method are used for adhesive coating into wood, wood composites cardboard, and similar materials with large surface areas.

Figure 17: adhesive is dispensed by a manually controlled dispenser to bond parts during assembly.

Figure 18: Roll coating adhesive into thin, flexible material such as paper, cloth, or flexible polymer.

4.6

Advantages and limitations

The advantages of adhesive bonding are: The process is applicable to a wide variety of materials Parts of different sized can be joined fragile parts can be joined by adhesive bonding Bonding occurs over the entire surface area of the joint, rather than in discrete spot or seams as in fusion welding, thereby distributing stresses over the entire area Some adhesives are flexible after bonding and are thus tolerant of cylindrical loading and difference of thermal expansion of adherents. Low temperature curing avoids damage to parts being joined Sealing as well as bonding can be achieved Joint designs are often simplified (e.g., two flat surfaces can be joined without providing special part features such as screw holes).

The principal disadvantages are: Joints are generally not as strong as other joining methods. Adhesive must be compatible with materials being joined. Service temperatures are limited Cleanliness and surface preparation prior to application of adhesive are important Curing times can impose a limit on production rates Inspection of the bonded joint is difficult

You might also like