You are on page 1of 245

http://users6.nofeehost.

com/mestijaya/cmm/

Inspector Knowledge Series 04-0


Effect of Alloying Elements on Steels

Mok Chek Min

Descriptive approach- Steel Alloys

This Ebook are meant to be read with internet connection hook-on. Online interactive material, videos and animations will assist you in the understanding of corrosion basic. Video contents are highlighted by icons

BookPlan In house training ADDITIONAL CODE REMARKS MAIN TAG NUMBER CLIENT PO NUMBER DISCIPLINE Metallurgy SDRL CODE TOTAL PGS

116+

CLIENT DOCUMENT NUMBER 20081018-Rev001

CMM NDT Services

Menu

CM MOK cmmok128@yahoo.com

Chapter One:
Theory of Strengthening
The Mechanism of Strengthening. Elastic and Plastic Deformation. The Nature of Dislocation. Crystal Defects. Point Defects Linear Defects Planar Defects. Bulk Defects. Pinning Overview. Works or Strain Hardening. Grain Boundary Strengthening. Dispersion Strengthening. Solid Solution Strengthening. Precipitation Hardening. Understanding Dislocation.

Methods of Strengthening.

Chapter Two:
Effects of Alloying Elements on Iron Carbon Alloy
Introduction Characteristics of alloying element

Brief Summary. Carbon. Manganese. Nickel. Chromium. Nickel & Chromium Molybdenum. Vanadium Tungsten Silicon Copper Phosphorous Sulphur
Lead Hydrogen Selenium Tantalum Tin Calcium Cerium Nitrogen Boron Aluminium Zirconium Niobium Titanium Cobalt

Appendix D. Page 1 of 116.

Menu

CM MOK cmmok128@yahoo.com Carbon contents, steel classification and alloy steels. Carbon steels Stainless steels Strength in steels Alloying and its effects on the critical temperature and tensile strength. Control of HSLA properties. Influence of alloying elements on steel microstructures Martensite in austenitic stainless steel welds. Alloying effects on martensite, pearlite and bainite formation. Steel alloys Carbon steels to austenitic steels. Selection of age-hardenable superalloys. Microplasticity More reading.

http://www.arab-eng.org/vb/t93373.html Introduction to Dislocations, Fourth Edition by Derek Hull, D J Bacon http://mihd.net/lmec67y

Appendix D. Page 2 of 116.

Menu

CM MOK cmmok128@yahoo.com

Chapter One: The Mechanism of strengthening.

Elastic/Plastic Deformation When a sufficient load is applied to a metal or other structural material, it will cause the material to change shape. This change in shape is called deformation. A temporary shape change that is self-reversing after the force is removed, so that the object returns to its original shape, is called elastic deformation. In other words, elastic deformation is a change in shape of a material at low stress that is recoverable after the stress is removed. This type of deformation involves stretching of the bonds, but the atoms do not slip past each other. When the stress is sufficient to permanently deform the metal, it is called plastic deformation. As discussed in the section on crystal defects, plastic deformation involves the breaking of a limited number of atomic bonds by the movement of dislocations. Recall that the force needed to break the bonds of all the atoms in a crystal plane all at once is very great. However, the movement of dislocations allows atoms in crystal planes to slip past one another at a much lower stress levels. Since the energy required to move is lowest along the densest planes of atoms, dislocations have a preferred direction of travel within a grain of the material. This results in slip that occurs along parallel planes within the grain. These parallel slip planes group together to form slip bands, which can be seen with an optical microscope. A slip band appears as a single line under the microscope, but it is in fact made up of closely spaced parallel slip planes as shown in the image.

Plastic deformation occurs when large numbers of dislocations move and multiply so as to result in macroscopic deformation. In other words, it is the movement of dislocations in the material which allows for deformation. If we want to enhance a material's mechanical properties (i.e. increase the yield and tensile strength), we simply need to introduce a mechanism which prohibits the mobility of these dislocations. Whatever the mechanism may be, (work hardening, grain size reduction, etc) they all hinder dislocation motion and render the material stronger than previously.

Appendix D. Page 3 of 116.

Menu

CM MOK cmmok128@yahoo.com
The stress required to cause dislocation motion is orders of magnitude lower than the theoretical stress required to shift an entire plane of atoms, so this mode of stress relief is energetically favorable. Hence, the hardness and strength (both yield and tensile) critically depend on the ease with which dislocations move. Pinning points, or locations in the crystal that oppose the motion of dislocations, can be introduced into the lattice to reduce dislocation mobility, thereby increasing mechanical strength. Dislocations may be pinned due to stress field interactions with other dislocations and solute particles, or physical barriers from grain boundaries and second phase precipitates. There are several strengthening mechanisms for metals, however the key concept to remember about strengthening of metallic materials is that it is all about preventing dislocation motion and propagation; you are making it energetically unfavorable for the dislocation to move or propagate. For a material that has been strengthened, by some processing method, the amount of force required to start irreversible (plastic) deformation is greater than it was for the original material. In amorphous materials such as polymers, amorphous ceramics (glass), and amorphous metals, the lack of long range order leads to yielding via mechanisms such as brittle fracture, crazing, and shear band formation. In these systems, strengthening mechanisms do not involve dislocations, but rather consist of modifications to the chemical structure and processing of the constituent material. Unfortunately, strength of materials cannot infinitely increase. Each of the mechanisms elaborated below involves some trade off by which other material properties are compromised in the process of strengthening. Steel can be strengthened by several basic mechanisms, the most important of which are: 1. 2. 3. 4. 5. 6. 7. 8. Work hardening or strain hardening. Solid solution strengthening by interstitial atoms. Solid solution strengthening by substitutional atoms. Refinement of grain size. Effects of heat treatment on microstructures. Precipitation strengthening. Grain boundary strengthening. Dispersion strengthening, including lamellar and random dispersed structures.

The most distinctive aspect of strengthening of iron and steel is the role of the interstitial solutes carbon and nitrogen. These elements also play a vital part in interacting with dislocations, and in combining preferentially with some of the metallic alloying elements used in steels.

Appendix D. Page 4 of 116.

Menu

CM MOK cmmok128@yahoo.com

Understanding dislocations
Introduction to dislocations http://www.msm.cam.ac.uk/doitpoms/tlplib/dislocations/printall.php

The Nature of Dislocations

Plastid deformation is a measure of material strength, plastic deformation is irreversible. Therefore, the configuration of the atoms must be changed during plastic deformation, for otherwise they would return to their original position on unloading. If we consider shearing a single crystal as an example, it can be deformed plastically by sliding whole layers of atoms against each other as shown in figure above for this sliding to happen, the bonds between the atoms have to be stretched elastically until they can switch to the next atom. The stress required for this process can be estimated and is of the order of one fifth of the shear modulus of the crystal. The yield strength predicted this way for metallic single crystals is thus between 1GPa and 25GPa. If we measure the strength of single crystals of pure metals, the values found are several orders of magnitudes below this theoretical value and even lie below that of engineering alloys. Typical values are in the range of a few mega Pascal. As single crystals always contain lattice defects, one possible explanation could be that these are responsible for the reduced strength. If, however, the number of defects is reduced further, for instance by a heat treatment, the yield strength becomes even smaller. Only an absolutely perfect single crystal without any defects would possess a yield strength agreeing with the theoretical prediction. This can only be nearly realized in so-called whiskers, which, however, are extremely small. The reason for this spectacular failure of the theoretical prediction is that plastic deformation does not occur by sliding of complete layers of atoms. Instead, it proceeds by a mechanism that is based on a special type of lattice defect, the dislocations. To understand plastic deformation of metals thus requires an understanding of dislocations.

Appendix D. Page 5 of 116.

Menu

CM MOK cmmok128@yahoo.com
Dislocation densities Dislocation is a lattice imperfection in a crystal structure which exerts a profound effect on a structure sensitive properties such as strength, hardness, ductility and toughness. There are two types, edge and screw or combination of both, all of which are characterized by a Burgers vector which represents the amount and direction of slip when the dislocation moves. Click on the web links provided to read further.
Interaction of Dislocations

Transmission Electron Micrograph of Dislocations

Transmission Electron Micrograph of Dislocations

More reading:

http://en.wikipedia.org/wiki/Dislocation http://www-sgrgroup.materials.ox.ac.uk/lectures/microplasticity.html Microplasticity: dislocations and strengthening mechanisms

"Atomistics" of edge dislocation motion and Asymmetry of screw and edge dislocation motion in Mo

Appendix D. Page 6 of 116.

Menu

CM MOK cmmok128@yahoo.com

Crystal Defects
A perfect crystal, with every atom of the same type in the correct position, does not exist. All crystals have some defects. Defects contribute to the mechanical properties of metals. In fact, using the term defect is sort of a misnomer since these features are commonly intentionally used to manipulate the mechanical properties of a material. Adding alloying elements to a metal is one way of introducing a crystal defect. Nevertheless, the term defect will be used, just keep in mind that crystalline defects are not always bad. There are basic classes of crystal defects: point defects, which are places where an atom is missing or irregularly placed in the lattice structure. Point defects include lattice vacancies, self-interstitial atoms, substitution impurity atoms, and interstitial impurity atoms linear defects, which are groups of atoms in irregular positions. Linear defects are commonly called dislocations. planar defects, which are interfaces between homogeneous regions of the material. Planar defects include grain boundaries, stacking faults and external surfaces. It is important to note at this point that plastic deformation in a material occurs due to the movement of dislocations (linear defects). Millions of dislocations result for plastic forming operations such as rolling and extruding. It is also important to note that any defect in the regular lattice structure disrupts the motion of dislocation, which makes slip or plastic deformation more difficult. These defects not only include the point and planer defects mentioned above, and also other dislocations. Dislocation movement produces additional dislocations, and when dislocations run into each other it often impedes movement of the dislocations. This drives up the force needed to move the dislocation or, in other words, strengthens the material. Each of the crystal defects will be discussed in more detail in the following pages.

Point Defects
Point defects are where an atom is missing or is in an irregular place in the lattice structure. Point defects include self interstitial atoms, interstitial impurity atoms, substitutional atoms and vacancies. A self interstitial atom is an extra atom that has crowded its way into an interstitial void in the crystal structure. Self interstitial atoms occur only in low concentrations in metals because they distort and highly stress the tightly packed lattice structure. A substitutional impurity atom is an atom of a different type than the bulk atoms, which has replaced one of the bulk atoms in the lattice. Substitutional impurity atoms are usually close in size (within approximately 15%) to the bulk atom. An example of substitutional impurity atoms is the zinc atoms in brass. In brass, zinc atoms with a radius of 0.133 nm have replaced some of the copper atoms, which have a radius of 0.128 nm. Interstitial impurity atoms are much smaller than the atoms in the bulk matrix. Interstitial impurity atoms fit into the open space between the bulk atoms of the lattice structure. An example of interstitial impurity atoms is the carbon atoms that are added to iron to make steel. Carbon atoms, with a radius of 0.071 nm, fit nicely in the open spaces between the larger (0.124 nm) iron atoms.

Appendix D. Page 7 of 116.

Menu

CM MOK cmmok128@yahoo.com
Vacancies are empty spaces where an atom should be, but is missing. They are common, especially at high temperatures when atoms are frequently and randomly change their positions leaving behind empty lattice sites. In most cases diffusion (mass transport by atomic motion) can only occur because of vacancies.

Linear Defects - Dislocations


Dislocations are another type of defect in crystals. Dislocations are areas were the atoms are out of position in the crystal structure. Dislocations are generated and move when a stress is applied. The motion of dislocations allows slip plastic deformation to occur. Before the discovery of the dislocation by Taylor, Orowan and Polyani in 1934, no one could figure out how the plastic deformation properties of a metal could be greatly changed by solely by forming (without changing the chemical composition). This became even bigger mystery when in the early 1900s scientists estimated that metals undergo plastic deformation at forces much smaller than the theoretical strength of the forces that are holding the metal atoms together. Many metallurgists remained skeptical of the dislocation theory until the development of the transmission electron microscope in the late 1950s. The TEM allowed experimental evidence to be collected that showed that the strength and ductility of metals are controlled by dislocations. There are two basic types of dislocations, the edge dislocation and the screw dislocation. Actually, edge and screw dislocations are just extreme forms of the possible dislocation structures that can occur. Most dislocations are probably a hybrid of the edge and screw forms but this discussion will be limited to these two types.
Appendix D. Page 8 of 116.

Menu

CM MOK cmmok128@yahoo.com
Edge Dislocations The edge defect can be easily visualized as an extra half-plane of atoms in a lattice. The dislocation is called a line defect because the locus of defective points produced in the lattice by the dislocation lie along a line. This line runs along the top of the extra half-plane. The inter-atomic bonds are significantly distorted only in the immediate vicinity of the dislocation line. Understanding the movement of a dislocation is key to understanding why dislocations allow deformation to occur at much lower stress than in a perfect crystal. Dislocation motion is analogous to movement of a caterpillar. The caterpillar would have to exert a large force to move its entire body at once. Instead it moves the rear portion of its body forward a small amount and creates a hump. The hump then moves forward and eventual moves all of the body forward by a small amount.

Appendix D. Page 9 of 116.

Menu

CM MOK cmmok128@yahoo.com
As shown in the set of images above, the dislocation moves similarly moves a small amount at a time. The dislocation in the top half of the crystal is slipping one plane at a time as it moves to the right from its position in image (a) to its position in image (b) and finally image (c). In the process of slipping one plane at a time the dislocation propagates across the crystal. The movement of the dislocation across the plane eventually causes the top half of the crystal to move with respect to the bottom half. However, only a small fraction of the bonds are broken at any given time. Movement in this manner requires a much smaller force than breaking all the bonds across the middle plane simultaneously. Screw Dislocations There is a second basic type of dislocation, called screw dislocation. The screw dislocation is slightly more difficult to visualize. The motion of a screw dislocation is also a result of shear stress, but the defect line movement is perpendicular to direction of the stress and the atom displacement, rather than parallel. To visualize a screw dislocation, imagine a block of metal with a shear stress applied across one end so that the metal begins to rip. This is shown in the upper right image. The lower right image shows the plane of atoms just above the rip. The atoms represented by the blue circles have not yet moved from their original position. The atoms represented by the red circles have moved to their new position in the lattice and have reestablished metallic bonds. The atoms represented by the green circles are in the process of moving. It can be seen that only a portion of the bonds are broke at any given time. As was the case with the edge dislocation, movement in this manner requires a much smaller force than breaking all the bonds across the middle plane simultaneously. If the shear force is increased, the atoms will continue to slip to the right. A row of the green atoms will find there way back into a proper spot in the lattice (and become red) and a row of the blue atoms will slip out of position (and become green). In this way, the screw dislocation will move upward in the image, which is perpendicular to direction of the stress. Recall that the edge dislocation moves parallel to the direction of stress. As shown in the image below, the net plastic deformation of both edge and screw dislocations are the same, however. The dislocations move along the densest planes of atoms in a material, because the stress needed to move the dislocation increases with the spacing between the planes. FCC and BCC metals have many dense planes, so dislocations move relatively easy and these materials have high ductility. Metals are strengthened by making it more difficult for dislocations to move. This may involve the introduction of obstacles, such as interstitial atoms or grain boundaries, to pin the dislocations. Also, as a material plastically deforms, more

Appendix D. Page 10 of 116.

Menu

CM MOK cmmok128@yahoo.com
dislocations are produced and they will get into each others way and impede movement. This is why strain or work hardening occurs. In ionically bonded materials, the ion must move past an area with a repulsive charge in order to get to the next location of the same charge. Therefore, slip is difficult and the materials are brittle. Likewise, the low density packing of covalent materials makes them generally more brittle than metals.

Planar Defects
Stacking Faults and Twin Boundaries A disruption of the long-range stacking sequence can produce two other common types of crystal defects: 1) a stacking fault and 2) a twin region. A change in the stacking sequence over a few atomic spacing produces a stacking fault whereas a change over many atomic spacing produces a twin region. A stacking fault is a one or two layer interruption in the stacking sequence of atom planes. Stacking faults occur in a number of crystal structures, but it is easiest to see how they occur in close packed structures. For example, it is know from a previous discussion that face centered cubic (fcc) structures differ from hexagonal close packed (hcp) structures only in their stacking order. For hcp and fcc structures, the first two layers arrange themselves identically, and are said to have an AB arrangement. If the third layer is placed so that its atoms are directly above those of the first (A) layer, the stacking will be ABA. This is the hcp structure, and it continues ABABABAB. However it is possible for the third layer atoms to arrange themselves so that they are in line with the first layer to produce an ABC arrangement which is that of the fcc structure. So, if the hcp structure is going along as ABABAB and suddenly switches to ABABABCABAB, there is a stacking fault present.
Appendix D. Page 11 of 116.

Menu

CM MOK cmmok128@yahoo.com
Alternately, in the fcc arrangement the pattern is ABCABCABC. A stacking fault in an fcc structure would appear as one of the C planes missing. In other words the pattern would become ABCABCAB_ABCABC. If a stacking fault does not corrects itself immediately but continues over some number of atomic spacing, it will produce a second stacking fault that is the twin of the first one. For example if the stacking pattern is ABABABAB but switches to ABCABCABC for a period of time before switching back to ABABABAB, a pair of twin stacking faults is produced. The red region in the stacking sequence that goes ABCABCACBACBABCABC is the twin plane and the twin boundaries are the A planes on each end of the highlighted region. Grain Boundaries in Polycrystals Another type of planer defect is the grain boundary. Up to this point, the discussion has focused on defects of single crystals. However, solids generally consist of a number of crystallites or grains. Grains can range in size from nanometers to millimeters across and their orientations are usually rotated with respect to neighboring grains. Where one grain stops and another begins is know as a grain boundary. Grain boundaries limit the lengths and motions of dislocations. Therefore, having smaller grains (more grain boundary surface area) strengthens a material. The size of the grains can be controlled by the cooling rate when the material cast or heat treated. Generally, rapid cooling produces smaller grains whereas slow cooling result in larger grains. For more information, refer to the discussion on solidification.

Bulk Defects
Bulk defects occur on a much bigger scale than the rest of the crystal defects discussed in this section. However, for the sake of completeness and since they do affect the movement of dislocations, a few of the more common bulk defects will be mentioned. Voids are regions where there are a large number of atoms missing from the lattice. The image to the right is a void in a piece of metal The image was acquired using a Scanning Electron Microscope (SEM). Voids can occur for a number of reasons. When voids occur due to air bubbles becoming trapped when a material solidifies, it is commonly called porosity. When a void occurs due to the shrinkage of a material as it solidifies, it is called cavitation. Another type of bulk defect occurs when impurity atoms cluster together to form small regions of a different phase. The term phase refers to that region of space occupied by a physically homogeneous material. These regions are often called precipitates. Phases and precipitates will be discussed in more detail latter.

Appendix D. Page 12 of 116.

Menu

CM MOK cmmok128@yahoo.com

Pinning points - overview.


In a crystalline material, a dislocation is capable of traveling throughout the lattice when relatively small stresses are applied. This movement of dislocations results in the material plastically deforming. Pinning points in the material act to halt a dislocation's movement, requiring a greater amount of force to be applied to overcome the barrier. This results in an overall strengthening of materials.

Motion of Dislocations at Elevated Temperatures Abstract The movies, which are in MPG format, show the movement of dislocations in ferritic steels (body-centred cubic crystal structure), at elevated temperatures. The images are taken in a hot-stage in a transmission electron microscope. These observations have been used to conclude the molybdenum and tungsten have the same solid-solution strengthening effect on ferrite

Types of pinning points


Point defects

Point defects (as well as stationary dislocations, jogs, and kinks) present in a material create stress fields within a material that disallow traveling dislocations to come into direct contact. Much like two particles of the same electric charge feel a repulsion to one another when brought together, the dislocation is pushed away from the already present stress field.

Alloying

elements

The introduction of atom1 into a crystal of atom2 creates a pinning point for multiple reasons. An alloying atom is by nature a point defect, thus it must create a stress field when placed into a foreign crystallographic position, which could block the passage of a dislocation. However, it is possible that the allowing material is approximately the same size as the atom that is replaced, and thus its presence would not stress the lattice (as occurs in cobalt alloyed nickel). The different atom would, though, have a different elastic modulus, which would create a different terrain for the moving dislocation. A higher modulus would look like an energy barrier, and a lower like an energy trough both of which would stop its movement.

Appendix D. Page 13 of 116.

Menu

CM MOK cmmok128@yahoo.com

Second phase precipitates

The precipitation of a second phase within the lattice of a material creates physical blockades through which a dislocation cannot pass. The result is that the dislocation must bend (which requires greater energy, or a greater stress to be applied) around the precipitates, which inevitably leaves residual dislocation loops encircling the second phase material and shortens the original dislocation. This is a schematic shows how a dislocation interacts with solid phase precipitates. The dislocation moves from left to right in each frame.

Dislocation slip mechanism Grain boundaries

Dislocations require proper lattice ordering to move through a material. At grain boundaries, there is a lattice mismatch, and every atom that lies on the boundary is uncoordinated. This stops dislocations that encounter the boundary from moving.

Appendix D. Page 14 of 116.

Menu

CM MOK cmmok128@yahoo.com

Methods of strengthening

Following are brief description s on the methods of strengthening.

Work or Strain hardening The reason for strain hardening is that the dislocation density increases with plastic deformation (cold work) due to multiplication. The average distance between dislocations then decreases and dislocations start blocking the motion of dislocations. The primary species responsible for work hardening are dislocations. Dislocations interact with each other by generating stress fields in the material. The interaction between the stress fields of dislocations can impede dislocation motion by repulsive or attractive interactions. Additionally, if two dislocations cross, dislocation line entanglement occurs, causing the formation of a jog which opposes dislocation motion. These entanglements and jogs act as pinning points, which oppose dislocation motion. As both of these processes are more likely to occur when more dislocations are present, there is a correlation between dislocation density and yield strength,

Where G is the shear modulus, b is the Burgers vector, and

is the dislocation density.

Increasing the dislocation density increases the yield strength which results in a higher shear stress required to move the dislocations. This process is easily observed while working a material. Theoretically, the strength of a material with no dislocations will be extremely high (=G/2) because plastic deformation would require the breaking of many bonds simultaneously. However, at moderate dislocation density values of around 107-109 dislocations/m2, the material will exhibit a significantly lower mechanical strength. Analogously, it is easier to move a rubber rug across a surface by propagating a small ripple through it than by dragging the whole rug. At dislocation densities of 1014 dislocations/m2 or higher, the strength of the material becomes high once again. It should be noted that the dislocation density can't be infinitely high because then the material would lose its crystalline structure. Work hardening is an important strengthening process in steel, particularly in obtaining high strength levels in rod and wire, both in plain carbon and alloy steels. For example, the tensile strength of a 0.05% C steel subjected to 95% reduction in area by wire drawing, is raised by no less than 550 MPa while higher carbon steels are strengthened by up to twice this amount. Indeed, without the addition of special alloying elements, plain carbon steels can be raised to strength levels above 1500 MPa simply by the phenomenon of work hardening. Basic work on the deformation of iron has largely concentrated on the other end of the strength spectrum, namely pure single crystals and polycrystals subjected to small controlled deformations. The diversity of slip

Appendix D. Page 15 of 116.

Menu

CM MOK cmmok128@yahoo.com
planes leads to rather irregular wavy slip bands in deformed crystals, as the dislocations can readily move from one type of plane to another by cross slip, provided they share a common slip direction. The yield stress of iron single crystals are very sensitive to both temperature and strain rate and a similar dependence has been found for less pure polycrystalline iron. Therefore, the temperature sensitivity cannot be attributed to interstitial impurities. It is explained by the effect of temperature on the stress needed to move free dislocations in the crystal, the Peierls-Nabarro stress.

Case Study 1: The effect of cyclic torsion on the dislocation structure of drawn mild steel 1. Introduction Cold forming of metals usually causes their work hardening. The magnitude of this hardening depends on the area reduction, on the temperature and strain rate associated with the processing, and on the way the strain is imposed on the metal. Keeping all other variables constant, the work hardening of a metal submitted to a sequential straining under varying directions or of different natures is different from that resulting from monotonic straining. Changes in the way the material is deformed can alter the hardening rates and even cause strain softening of the metal1-16. Recent research results17-19 show that cyclic straining influences in various ways the mechanical behavior of annealed and drawn metal bars. Annealed Aluminum submitted to cyclic torsion displays higher flow stresses than the annealed material. On the other hand, cyclic torsion softens previously drawn Aluminum. Cyclic torsion also softens steel bars previously drawn in one or two passes and hardens the initially annealed material. Experimental results indicate that the stress-strain curve and the work hardening coefficient (n) of steel drawn in two passes and submitted to cyclic torsion are similar to those for the material submitted to only one drawing pass. This is similar to the case of the Aluminum alloy 6063, where the cyclic torsion after two drawing passes eliminates the hardening associated with the second drawing pass. It is also observed for both materials that their Ultimate Tensile Strength (UTS) tends to remain unaltered by cyclic torsion, in the case of initially annealed material, whereas their Yield Strength (YS) is considerably increased by cyclic torsion. The YS and UTS of both previously drawn materials are decreased by cyclic torsion, with the exception of the YS of Aluminum drawn in a single pass. The decrease in these properties is more pronounced after two drawing passes than after a single drawing pass. Finally, cyclic torsion increases the Tensile Elongation to Fracture of drawn material and decreases this property for initially annealed material. The present research analyzes the relationship between the mechanical effects described above and the dislocation structures in Low Carbon steel. 2. Materials and Experimental Methods The material was an AISI 1010 steel with the following chemical composition: 0.12%C, 0.47%Mn, 0.07%Si, 0.003%Sn, 0.01%Mo, 0.016%P e 0.013%S, received as cylindrical bars 6.4 mm in diameter. The bars were initially annealed and some of them were drawn in one or two passes. A fraction of these bars were then submitted to cyclic torsion. The effects of the strain path were analyzed by Transmission Electron Microscopy (TEM). Annealing was performed under vacuum, at 850 C for 2400 seconds, leading to an average hardness of 122.6 HV. Drawing was performed in a hydraulic draw bench, using Tungsten Carbide dies with semi-angle of 8?and abundant lubrication with a Molybdenum Disulfide paste. Different dies were employed, guaranteeing a

Appendix D. Page 16 of 116.

Menu

CM MOK cmmok128@yahoo.com
fixed reduction of area of 20% in each pass. Cyclic torsion (11.2% plastic strain per cycle, total of 10 cycles) was performed in an especially adapted lathe, where the chuck was manually actuated. All experiments were performed at room temperature, at a strain rate of 0.002 / s5. All deformed samples were stored at temperatures below 0 C in order to avoid static strain aging effects. TEM was performed in a JEOL-JEM microscope, operated at 200 kV. Analyses were performed in 3 mm samples taken from the cross-section of the bars. Sample preparation involved initial mechanical polishing, followed by electrolytic thinning with a perchloric acid and ethanol solution at room temperature. 3. Results and Discussion The dislocation structures of the annealed and of the drawn material (in one or two passes) are shown in Figures 1 and The effect of cyclic torsion on the dislocation structure of drawn mild steel2 respectively. The dislocation density is quite low for the annealed material, whereas the deformed material displays a much higher dislocation density. Drawn material shows an aligned cell structure, with irregular cell sizes and cell wall thickness. Dense dislocation networks can be observed inside the cells. Higher drawing strains lead to a smaller cell size, as expected.

Figure 1.

Figure 2a.

Figure 2b. TEM of annealed and drawn steel (8% and 20% per pass) a: 1 pass b: 2 passes

The dislocation structure of the material submitted only to cyclic torsion is broadly similar to that resulting from drawing (see Figure 3), but the cell size is higher and the tendency to cell alignment is less pronounced than in

Appendix D. Page 17 of 116.

Menu

CM MOK cmmok128@yahoo.com
drawing. It is important to realize that the total strain caused by cyclic torsion is much higher than in drawing, but leads to essentially similar dislocation structures. The analysis of Figure 4 indicates that the dislocation structure of the material after one drawing pass is altered by subsequent cyclic torsion. There is an increase in the cell size, a decrease in their alignment and in the dislocation density within the cells, and an overall evolution of alignment in only one direction to a "checkerboard" appearance, which is typical of the development of two sets of aligned cells, corresponding to the two directions of twisting. This is also the situation for the material initially annealed, drawn in 2 passes and cyclically twisted (Figure 5). Under these circumstances, the dislocation cells tend to be larger and the dislocation density inside the cells lower than for one drawing pass followed by cyclic torsion. Cyclic torsion promotes dynamic recovery of the material, involving the annihilation of cell walls and the decrease of dislocation density inside the cells. This is similar to results from the analysis of the Bauschinger effect20. where such dislocation annihilation stems from dislocation movements in opposing directions.

Appendix D. Page 18 of 116.

Menu

CM MOK cmmok128@yahoo.com

Considering the widely established relationship between the material flow stress G, and the corresponding dislocation density r20: one should expect higher flow stresses in the drawn or cyclic twisted material than in the annealed material. On the other hand, the recovery promoted by cyclic torsion of the previously drawn material should lead to their softening and consequent lower flow stresses. 4. Conclusions Drawing of low carbon steel leads to the formation of an aligned dislocation structure, displaying irregular cell sizes and cell wall thickness as well as dense networks of dislocations inside the cells. Cyclic torsion promotes the softening of material previously strained by drawing. This softening is associated with a restructuring of the previous dislocation arrangement, involving an increase in cell size, a decrease in the dislocation density inside the cells and a "checkerboard" dislocation wall structure.

Appendix D. Page 19 of 116.

Menu

CM MOK cmmok128@yahoo.com
Grain Boundary Strengthening In grain boundary strengthening the grain boundaries act as pinning points impeding further dislocation propagation. Since the lattice structure of adjacent grains differs in orientation, it requires more energy for a dislocation to change directions and move into the adjacent grain. The grain boundary is also much more disordered than inside the grain, which also prevents the dislocations from moving in a continuous slip plane. Impeding this dislocation movement will hinder the onset of plasticity and hence increase the yield strength of the material. Grain boundaries act as an impediment to dislocation motion for the following two reasons: Dislocation must change its direction of motion due to the differing orientation of grains. Discontinuity of slip planes from grain 1 to grain

Under an applied stress, existing dislocations and dislocations generated by Frank-Read Sources will move through a crystalline lattice until encountering a grain boundary, where the large atomic mismatch between different grains creates a repulsive stress field to oppose continued dislocation motion. As more dislocations propagate to this boundary, dislocation 'pile up' occurs as a cluster of dislocations are unable to move past the boundary. As dislocations generate repulsive stress fields, each successive dislocation will apply a repulsive force to the dislocation incident with the grain boundary. These repulsive forces act as a driving force to reduce the energetic barrier for diffusion across the boundary, such that additional pile up causes dislocation diffusion across the grain boundary, allowing further deformation in the material. Decreasing grain boundary size decreases the amount of possible pile up at the boundary, increasing the amount of applied stress necessary to move a dislocation across a grain boundary. The higher the applied stress to move the dislocation, the higher the yield strength. Thus, there is then an inverse relationship between grain boundary size and yield strength, as demonstrated by the Hall-Petch equation. A lower number of dislocations per grain results in a lower dislocation 'pressure' building up at grain boundaries. This makes it more difficult for dislocations to move into adjacent grains. This relationship can be mathematically described as follows:

, Where k is a constant, d is the average grain diameter and y,0 is the original yield stress. However, when there is a large direction change in the orientation of the two adjacent grains, the dislocation may not necessarily move from one grain to the other but instead create a new source of dislocation in the adjacent grain. The theory remains the same that more grain boundaries create more opposition to dislocation movement and in turn strengthens the material.

Appendix D. Page 20 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 1: Hall-Petch Strengthening is limited by the size of dislocations. Once the grain size reaches about 10 nm, grain boundaries start to slide.
Obviously, there is a limit to this mode of strengthening, as infinitely strong materials do not exist. Grain boundary sizes can range from about 100 m (large grains) to 1 m (small grains). Lower than this, the size of dislocations begins to approach the size of the grains. At a grain size of about 10 nm, only one or two dislocations can fit inside of a grain (see Figure 1 above). This scheme prohibits dislocation pile-up and never results in grain boundary diffusion. The lattice resolves the applied stress by grain boundary sliding, resulting in a decrease in the material's yield strength; A phenomenon known as grain-boundary sliding. To understand the mechanism of grain boundary strengthening one must understand the nature of dislocation-dislocation interactions. Dislocations create a stress field around them given by:

Appendix D. Page 21 of 116.

Menu

CM MOK cmmok128@yahoo.com

, Where G is the material's shear modulus, and b is the Burgers vector. If the dislocations are in the right alignment with respect to each other, the local stress fields they create will repel each other. This helps dislocation movement along grains and across grain boundaries. Hence, the more dislocations are present in a grain, the greater the stress field felt by a dislocation near a grain boundary:

This is a schematic roughly illustrating the concept of dislocation pile up and how it effects the strength of the material. A material with larger grain size is able to have more dislocation to pile up leading to a bigger driving force for dislocations to move from one grain to another. Thus you will have to apply less force to move a dislocation from a larger than from a smaller grain, leading materials with smaller grains to exhibit higher yield stress.
In a polycrystalline metal, grain size has a tremendous influence on the mechanical properties. Because grains usually have varying crystallographic orientations, grain boundaries arise. While an undergoing deformation, slip motion will take place.

The refinement of the grain size of ferrite provides one of the most important strengthening routes in the heal treatment of steels. The grain size effect on the yield stress can therefore be explained by assuming that a dislocation source operates within a crystal causing dislocations to move and eventually to pile up at the grain boundary. The pile-up causes a stress to be generated in the adjacent grain, which, when it reaches a critical value, operates a new source in that grain. In this way, the yielding process is propagated from grain to grain. The grain size determines the distance dislocations have to move to form grain boundary pile-ups, and thus the number of dislocations involved. With large grain sizes, the pile-ups will contain larger numbers of dislocations, which will in turn cause higher stress concentrations in neighboring grains. In practical terms, the finer the grain size, the higher the resulting yield stress and, as a result, in modern steel working much attention is paid to the final ferrite grain size. While a coarse grain size of d-1/2 = 2, i.e. d = 0.25

Appendix D. Page 22 of 116.

Menu

CM MOK cmmok128@yahoo.com
mm, gives a yield stress in mild steels of around 100 MPa, grain refinement to d-1/2 = 20. i.e. d = 0.0025 mm, raises the yield stress to over 500 MPa, so that achieving grain sizes in the range 2-10 m is extremely worthwhile.

Appendix D. Page 23 of 116.

Menu

CM MOK cmmok128@yahoo.com
Dispersion Strengthening (Mechanical Alloying) Dispersion strengthening is about the interaction between dislocations and finely dispersed particles in the metal matrix. Traditionally, dispersion particles are obtained as precipitates from phase transformations during traditional metallurgical processes (melting, solidification, and heat treating). This is the case of many high-strength commercial alloys. However, maximum service temperatures are limited by the limited thermal stability of particles or precipitates obtained from thermal processes.

A: Lattice distortion due to the presence of coherent precipitate: B: Noncohereny precipitates produce no lattice distortion

Particles, which may not be metallurgical compatible with a given metal, can be introduced in a metal matrix by violently deforming mixtures of different powders. Such technique is called mechanical alloying . An example is high-energy ball milling. Using this technique, oxides and other highly stable chemical species can be introduced uniformly into the metal microstructure. A further refinement of this technique allows oxidation and other chemical reactions to take place during the mixing process, a technique that allegedly promotes the formation of ultrafine dispersoids as well as allowing control of particle composition and distribution. After mixing, the powder mixture can then be consolidated and compacted using a number of powder metallurgy techniques to produce a solid with a very fine grain structure. Mechanical alloying methods permit the manufacturing of metallic alloys with a number of interesting properties. Some examples of materials include nanocrystalline and amorphous materials, metastable phases, and alloys with extended solubility limits. Mechanical alloying was originally developed as a means of raising the maximum service temperatures of nickel-based superalloys for aircraft gas-turbine applications. As turbine operating temperatures rose, so too did the demand for materials with increased high-temperature strength and oxidation resistance. Mechanical alloying avoids many of the problems associated with conventional melting and solidification processes, and is now used to make a variety of oxide-dispersion-strengthened (ODS) Fe-Cr, Ni-Cr, and Ni-Cr-gamma superalloys for turbine-engine industrial applications and aluminum alloys for aircraft structural components. Mechanical alloying may play a key role in the development of future aerospace systems by enabling the production of even higher performance materials that are difficult or impossible to make by other methods. The effectiveness of dispersion strengthening largely depends on (a) mechanical and geometrical characteristics of the dispersoids, such as hardness, continuity, size, and shape, and (b) their density and distribution (dispersion factor) in the metal matrix. By controlling process parameters, such as characteristics

Appendix D. Page 24 of 116.

Menu

CM MOK cmmok128@yahoo.com
of the raw powder materials, pressing pressures, atmospheres, cycle time, and temperatures, the microstructure (e.g., grain size, dispersion size, volume fraction, and distribution) can be tailored for specific performance requirements. New research initiatives in this field include the synthesis of refractory carbide nanoparticles and assessment of their use as dispersion strengthening agents, as well as alternative methods to inoculate the metal matrix.

In all steels there is normally more than one phase present, and indeed it is often the case that several phases can be recognized in the microstructure. The matrix, which is usually ferrite (bcc structure) or austenite (fcc structure) strengthened by grain size refinement and by solid solution additions, is further strengthened, often to a considerable degree, by controlling the dispersions of the other phases in the microstructure. The commonest other phases are carbides formed as a result of the low solubility of carbon in -iron. In plain carbon steels this carbide is normally Fe3C (cementite), which can occur, in a wide range of structures from coarse lamellar form (pearlite), to fine rod or spheroidal precipitates (tempered steels). In alloy steels, the same range of structures is encountered, except that in many cases iron carbide is replaced by other carbides, which are thermodynamically more stable. Other dispersed phases which are encountered include nitrides, intermetallic compounds, and, in cast irons, graphite. Most dispersions lead to strengthening, but often they can have adverse effects on ductility and toughness. In fine dispersions (where ideally small spheres are randomly dispersed in a matrix) are well-defined relationships between the yield stress, or initial flow stress, and the parameters of the dispersion. These relationships can be applied to simple dispersions sometimes found in steels, particularly after tempering, when, in plain carbon steels, the structure consists of spheroidal cementite particles in a ferritic matrix. However, they can provide approximations in less ideal cases, which are the rule in steels, where the dispersions vary over the range from fine rods and plates to irregular polyhedral. Perhaps the most familiar structure in steels is that of the eutectoid pearlite, usually a lamellar mixture of ferrite and cementite. This can be considered as an extreme form of dispersion of one phase in another, and undoubtedly provides a useful contribution to strengthening.

Appendix D. Page 25 of 116.

Menu

CM MOK cmmok128@yahoo.com

General behaviour of the powder particles during mechanical alloying processing

Case Study: Effect of mechanical alloying and Ti addition on solution and ageing treatment of an AA7050 aluminium alloy http://www.scielo.br/scielo.php?pid=S1516-14392007000200017&script=sci_arttext

Mechanical Alloys & Milling http://www.scribd.com/doc/3629131/Mechanical-alloying-and-milling-Suryanarayana

Appendix D. Page 26 of 116.

Menu

CM MOK cmmok128@yahoo.com
Solid solution strengthening

This is a schematic illustrating how the lattice is strained by the addition of substitutional and interstitial solute. Notice the strain in the lattice that the solute atoms cause. The interstitial solute could be carbon in iron for example. The carbon atoms in the interstitial sites of the lattice create a stress field that impedes dislocation movement. http://en.wikipedia.org/wiki/Strengthening_mechanisms_of_materials

A grain boundary in a 2D lattice is the interface between two regions of crystalline order. Each region or 'grain' has a different orientation with respect to some arbitrary axis perpendicular to the plane of the lattice.

Grain boundaries A vacancy is a point defect that arises when an atom is 'missing' from the ideal crystal structure.

A vacancy A solute atom in a crystal structure is an atomic species that is different from the majority of atoms that form the

Appendix D. Page 27 of 116.

Menu

CM MOK cmmok128@yahoo.com
structure. Solute atoms of similar size to those in the host lattice may substitute for host atoms - these are known as substitutional solutes. Solute atoms that are much smaller than the host atoms may exist within normally empty regions (interstices) in the host lattice, where they are called interstitial solutes.

Substitutional and interstitial solutes. Note that some distortion of the host lattice occurs around the solutes. A dislocation in a 2D close-packed plane can be described as an extra 'half-row' of atoms in the structure. Dislocations can be characterised by the Burgers vector which gives information about the orientation and magnitude of the dislocation.

Dislocation

http://www.msm.cam.ac.uk/doitpoms/tlplib/dislocations/dislocations_in_2D.php

Appendix D. Page 28 of 116.

Menu

CM MOK cmmok128@yahoo.com

Solid solution strengthening by interstitials The formation of interstitial atmospheres at dislocations requires diffusion of the solute. As both carbon and nitrogen diffuse much more rapidly in iron than substitutional solutes, it is not surprising that strain ageing can take place readily in the range from 20C to 150C. Consequently the atmosphere condenses to form rows of interstitial atoms along the cores of the dislocations. These arise because the temperature is high enough to allow interstitial atoms to diffuse during deformation, and to form atmospheres around dislocations generated throughout the stress-strain curve. Steels tested under these conditions also show low ductility, due partly to the high dislocation density and partly to the nucleation of carbide particles on the dislocations where the carbon concentration is high. The phenomenon is often referred to as blue brittleness, blue being the interference color of the steel surface when oxidized in this temperature range. The break away of dislocations from their carbon atmospheres as a cause of the sharp yield point became a controversial aspect of the theory because it was found that the provision of free dislocations, for example, by scratching the surface of a specimen, did not eliminate the sharp yield point. An alternative theory was developed which assumed that, once condensed carbon atmospheres are formed in iron, the dislocations remain locked, and the yield phenomena arise from the generation and movement of newly formed dislocations. To summarize, the occurrence of a sharp yield point depends on the occurrence of a sudden increase in the number of mobile dislocations. However, the precise mechanism by which this takes place will depend on the effectiveness of the locking of the pre-existing dislocations. If the pinning is weak, then the yield point can arise as a result of unpinning. However, if the dislocations are strongly locked, either by interstitial atmospheres or precipitates, the yield point will result from the rapid generation of new dislocations. Under conditions of dynamic strain ageing, where atmospheres of carbon atoms form continuously on newly-generated dislocations, it would be expected that a higher density of dislocations would be needed to complete the deformation, if it is assumed that most dislocations which attract carbon atmospheres are permanently locked in position.

Strengthening at high interstitial concentrations Austenite can take into solid solution up to 10% carbon, which can be retained in solid solution by rapid quenching. However, in these circumstances the phase transformation takes place, not to ferrite but to a tetragonal structure referred to as martensite. This phase forms as a result of diffusion less shear transformation leading to characteristic laths or plates. If the quench is sufficiently rapid, the martensite is essentially a supersaturated solid solution of carbon in a tetragonal iron matrix, and as the carbon concentration can be greatly in excess of the equilibrium concentration in ferrite, the strength is raised very substantially. High carbon martensites are normally very hard but brittle, the yield strength reaching as much as 1500 MPa; much of this increase can be directly attributed to increased interstitial solid solution hardening, but there is also a contribution from the high dislocation density, which is characteristic of martensitic transformations in iron-carbon alloys. Substitutional solid solution strengthening of iron Many metallic elements form solid solutions in - and -iron. These are invariably substitutional solid solutions, but for a constant atomic concentration of alloying elements there are large variations in strength. Using single crystal data for several metals, Fig. 1 shows that an element such as vanadium has a weak strengthening effect on -iron at low concentrations (< 2%), while silicon and molybdenum are much more effective

Appendix D. Page 29 of 116.

Menu

CM MOK cmmok128@yahoo.com
strengthened. Other data indicates that phosphorus; manganese, nickel and copper are also effective strengtheners. However, it should be noted that the relative strengthening might alter with the temperature of testing, and with the concentrations of interstitial solutes present in the steels.

Figure 1. Solid solution strengthening of iron crystals by substitutional solutes. Ratio of the critical resolved shear stress 0 to shear modulus as a function of atomic concentration.

The strengthening achieved by substitutional solute atoms is, in general, greater the larger the difference in atomic size of the solute from that of iron, applying the Hume-Rothery size effect. However, from the work of Fleischer and Takeuchi it is apparent that differences in the elastic behavior of solute and solvent atoms are also important in determining the overall strengthening achieved. In practical terms, the contribution to strength from solid solution effects is superimposed on hardening from other sources, e.g. grain size and dispersions. Also it is a strengthening increment, like that due to grain size, which need not adversely affect ductility. In industrial steels, solid solution strengthening is a far from negligible factor in the overall strength, where it is achieved by a number of familiar alloying elements, e.g. manganese, silicon, nickel, molybdenum, several of which are frequently present in a particular steel and are additive in their effect. These alloying elements arc usually added for other reasons, e.g. Si to achieve deoxidation, Mn to combine with sulphur or Mo to promote hardenability. Therefore, the solid solution hardening contribution can be viewed as a useful bonus.

Appendix D. Page 30 of 116.

Menu

CM MOK cmmok128@yahoo.com

Precipitation Hardening In most binary systems, alloying above a concentration given by the phase diagram will cause the formation of a second phase. A second phase can also be created by mechanical or thermal treatments. The particles that compose the second phase precipitates act as pinning points in a similar manner to solutes, though the particles are not necessarily single atoms. The dislocations in a material can interact with the precipitate atoms in one of two ways (see Figure 2). If the precipitate atoms are small, the dislocations would cut through them. As a result, new surfaces (b in Figure 2) of the particle would get exposed to the matrix and the particle/matrix interfacial energy would increase. For larger precipitate particles, looping or bowing of the dislocations would occur which results in dislocations getting longer. Hence, at a critical radius of about 5nm, dislocations will preferably cut across the obstacle while for a radius of 30nm, the dislocations will readily bow or loop to overcome the obstacle.

Appendix D. Page 31 of 116.

Menu

CM MOK cmmok128@yahoo.com

Chapter Two:

The Effects of Alloying Elements on Iron-Carbon Alloys

Effects of Alloying Additions to Steel

Element

Influence

Uses

Carbon

Hardness - Strength - Wear Most important alloying element. Is essential to the formation of cementite and other carbides, bainite and Added to construction steels to increase iron-carbon martensite. Within limits increasing the strength, hardness and hardenability. carbon content increases the strength and hardness of a steel while reducing its toughness and ductility. Deoxidation - Ease of Nitriding Hardenability Corrosion Resistance - Strength Machinability Strength Deoxidation - Hardenability Machinability Machinabilty Toughness - Strength - Hardenability Stabilises gamma phase by raising A4 and lowering A3. Refines grains in steels and some non-ferrous alloys. Strengthens ferrite by solid solution. Unfortunatly is a powerful graphitiser. Can take into solid solution larger proportions of important elements such as chromium, molybdenum and tungsten than can iron. Used up to help refine grain size. Used in large amounts in stainless and heat-resisting steels. Nickel based alloys can offer corrosion resistance in more aggressive environments and nickel is used as the basis of complex superalloys for high temperature service.

Aluminum Boron Copper Lead Phosphorus Silicon Sulfur Tellurium

Nickel

Manganese

Strength - Hardenability - More Response To Heat Treatment High manganese (Hadfield) steel contains Deoxidises the melt. Greatly increases the hadenability 12.5% Mn and is austenitic but hardens on of steels. Stabilises gamma phase. Forms stable abrasion. carbides. De-oxidises melt. Helps casting fluidity. Improves Up to 0.3% in steels for sandcasting, up to oxidation resistance at higher temperatures. 1% in heat resisting steels.

Silicon

Appendix D. Page 32 of 116.

Menu

CM MOK cmmok128@yahoo.com Corrosion Resistance - Strength Stabilises alpha phase by raising A3 and depressing A4. Forms hard stable carbides. Strengthens ferrite by solid solution. In amounts above 13% it imparts stainless properties. Unfortunately increases grain growth.

Chromium

Small amounts in constructional and tool steels. About 1.5% in ball and roller bearings. Larger amounts in Stainless and heat-resisting steels.

High Temperature Strength - Hardenability Strong carbide-stabilising influence. Raises high temperature creep strength of some alloys. Slows tempering response. When added to stainless steels it greatly improves the Molybdenum pitting and crevice corrosion resistance. There are limits to the proportion that can be taken into an iron based matrix. However up to almost 30% can be incorporated into nickel based alloys which provides excellent corrosion resistance in many aqueous environments.

Reduces 'temper brittleness' in nickel-chromium steels. Increases red-hardness of tool steels. Now used to replace some tungsten in high-speed steels.

Vanadium

Fine Grain - Toughness Strong carbide forming tendency. Stabilises martensite Used to retain high temperature hardness, and increases hardenability. Restrains grain growth. e.g. in dies for hot-forging and die casting Improves resistance to softening at elevated dies. Increasingly used in high speed steels. temperatures after hardening. Used in high-speed steels and other tool and die steels, particularly those for use at high temperatures. Stabilises alpha phase and forms stable, very hard Used in a few stainless steels, in carbides, which improves creep resistance and renders combination with molybdenum. to improve transformations very sluggish, hence hardened steels pitting and crevice corrosion resistance. It resist tempering influences. is also used in some high temperature nickel based alloys and in some high temperature austenitic stainless steels. Hardness - Wear Has similar corrosion resistance to that of Nickel, but higher cost means that it is not normally used for such applications. Used in super high speed steels and Provides matrix - strengthening characteristics to maraging steels, permanent magnet steels stainless and nickel based alloys designed for high and alloys. temperature applications. Slows the transformation of martensite, hence increases 'red hardness' which is useful in tool steels. Elimination Of Carbide Precipitation Used to stabilize stainless steels. In low alloy steels it acts as a carbide former and

Tungsten

Cobalt

Niobium

Appendix D. Page 33 of 116.

Menu

CM MOK cmmok128@yahoo.com improves creep resistance. In stainless steels it combines with carbon, stabilising the steel and reducing the susceptibility to intergranular corrosion Elimination Of Carbide Precipitation Used in stabilized stainless steels. In stainless steels combines with excess carbon In nickel based alloys it is used with reducing the risk of intergranular corrosion. aluminium to promote age hardening.

Titanium

Appendix D. Page 34 of 116.

Menu

CM MOK cmmok128@yahoo.com

Introduction
Typical 'mild' steels have a small carbon content, usually under 0.2%. Increasing the carbon content hardens the steel and its ability to take and hold an edge but at the expense of toughness. Adding other alloying elements can alter these properties, though usually at the expense of increased cost and manufacturability. The main alloying elements, and their effects are describe in next section. Carbon steels usually contain less than 1 to 2% carbon and small quantities of manganese, copper, silicon, sulfur, and phosphorus. Alloy steels are carbon steel with other metals added specifically to improve the properties of the steel significantly. Stainless steel is considered a separate group. Plain carbon steel is produced with a wide range of mechanical properties with comparatively low cost. To extend the range of properties of steel, alloys have been developed. The benefits resulting include The maximum UTS is increased. Thick sections steels are available with high hardness throughout the section. More controllable quenching with minimum risk of shape distortion or cracking. Improved impact resistance at high temperature range. Improved corrosion resistance. Improved high temperature performance.

The principle elements that are used in producing alloy steel include nickel, chromium, molydenenum, manganese, silicon and vanadium. Cobalt, copper and lead are also used as alloying elements. Effect of alloying elements Elements may encourage formation of graphite from the carbide. Only a small proportion of these elements can be added to the steel before graphite forms destroying the properties of the steel, unless elements are added to counteract the effect. Elements which encourage the formation of graphite include silicon, cobalt, aluminium and nickel Alloying elements may go into solid solution in the iron, enhancing the strength. Elements which go into solid solution include silicon, molybdenum, chromium, nickel and magnesium. Hard carbides (cementite) associated with iron and carbon may be formed with alloying elements. Elements which tend to form carbides include chromium, tungsten, titanium, columbium, vanadium, molybdenum and manganese. Elements which stabilize austenite include manganese, nickel, cobalt and copper. These increase the range over which austenite is stable e.g. by lowering the eutectoid temperature, and this retards the separation or carbides. If these alloys are present is certain high levels the austenite phase is dramatically reduced and the ferrite ( ) phase exists down to ambient temperatures e.g.18% chromium .

Appendix D. Page 35 of 116.

Menu

CM MOK cmmok128@yahoo.com

Elements which tend to stabilize ferrite include chromium, tungsten, molybdenum, vanadium and silicon. They reduce the amount of carbon soluble in the austenite and thus increase the volume of free carbide in the steel at a given carbon content. The effectively reduce the austenite ( ) phase by raising the eutectoid temperature and lowering the peritectic temperature Intermediate compounds with iron may be formed e.g. FeCr Alloying elements may adjust the characteristics such as eutectoid content, quenching rate which produces bainite or martensite.

Relative effect alloying elements The combined effect of alloying elements results from many complex interactions resulting from the processing history, the number and quantities of constituents, the heat treament, the section shape etc. Some basic rules can be identified. Nickel has reduced carbide forming tendency than iron and dissolves in ferrite. Silicon combines with oxygen to form nonmetallic inclusions or dissolves in the ferrite. Most of the manganese in alloy steels dissolves in the ferrite . carbides result in (Fe,Mn)3C. Chromium spreads between the ferrite and carbide phases the spread depending on the amount of carbon and other carbide generating elements present. Tungsten and molybdenum form carbides if sufficient carbon is present which has not already formed carbides with other stronger carbide forming elements. Vanadium , titanium, and Colombian are strong carbide forming elements and are present in steel as carbides. Aluminium combines with oxygen and nitrogen to form Al2O an AlN Any manganese that form

Appendix D. Page 36 of 116.

Menu

CM MOK cmmok128@yahoo.com

Characteristics of alloying
Effects of Elements on Steel Steels are among the most commonly used alloys. The complexity of steel alloys is fairly significant. Not all effects of the varying elements are included. The following text gives an overview of some of the effects of various alloying elements. Additional research should be performed prior to making any design or engineering conclusions. The Periodic Table:

The Atomic Radii:

Appendix D. Page 37 of 116.

Menu

CM MOK cmmok128@yahoo.com
In specifying values for the radius of an atom, one must keep in mind the fact that atoms are not hard spheres, and the electron distribution in the outer part of the atom does not have a sharp cutoff radius. You could characterize the radius of the atom as a limiting radius where a certain percentage of the electron charge will be found. The illustration above is a plot of "covalent radii" (from Ebbing) which are determined by measuring the bond lengths in the molecules of chemical compounds. Another way to determine characteristic radii is to measure ionic radii in crystals using x-ray diffraction. If the crystalline composition is such that the ions can be considered to be in contact with each other, and you can determine the lattice spacing from x-ray diffraction, then you can imply the ionic radius. As might be expected, the ionic radius of negative ions is slightly larger than the covalent radius since they have extra electronic charge, and that of positive ions is slightly smaller.

Atoms and Nuclei http://www.practicalphysics.org/go/Topic_40.html?topic_id=40

Appendix D. Page 38 of 116.

Menu

CM MOK cmmok128@yahoo.com

Carbon

Carbon has a major effect on steel properties. Carbon is the primary hardening element in steel. Hardness and tensile strength increases as carbon content increases up to about 0.85% C as shown in the figure above. Ductility and weldability decrease with increasing carbon. Carbon is essential in steels which have to be hardened by quenching and for example, in austenitic manganese steel which is required to have high resistance to wear. The maximum hardness obtainable in any carbon steel is a function of the carbon content which may vary up to about 2% according to the purpose for which the steel is to be used. It occurs in varying forms according to the percentage present, and the heat treatment to which the steel has been submitted. (See allotropy and transformation range). Cast irons usually contain from about 1.8% to 4.5% carbon, present either as free carbon (graphite) and/or combined carbon (cementite), the varying distribution of the carbon between these two forms considerably influencing the strength and hardness. Carbon is a strong austenite former and strongly promotes an austenitic structure. It also substantially increases the mechanical strength. Carbon reduces the resistance to intergranular corrosion. In ferritic stainless steels carbon will strongly reduce both toughness and corrosion resistance. In the martensitic and martensitic-austenitic steels carbon increases hardness and strength. In the martensitic steels an increase in hardness and strength is generally accompanied by a decrease in toughness and in this way carbon reduces the toughness of these steels. Carbon: Present in all steels, it is the most important hardening element. Also increases the strength of the steel. We usually want knife-grade steel to have >.5% carbon, which makes it "high-carbon" steel. it has by far the greatest influence of any of the elements. Steel could not exist without carbon. Martensite, along with banite gives steel a microstructure of hard, tough carbide. None of the other elements so dramatically alter the strength and hardness as do small changes in carbon content. Carbon iron crystalline structures have the widest number and variety known to exist in metallurgy. They also combine with other elements to furnish steel with an assortment of iron alloy carbide systems.

Appendix D. Page 39 of 116.

Menu

CM MOK cmmok128@yahoo.com

Manganese
Manganese fulfils a variety of functions in steel. It is used as a deoxidizing agent in nearly all steels. It forms manganese sulphide inclusions which in the ingot are spherical. In the absence of manganese sulphur forms interdendritic films of iron sulphide causing brittleness at forging temperature (hot shortness). It effectively increases harden ability and up to 1.5% is added for this purpose. (d) In larger amounts it is used to stabilize austenite, as in 14% manganese steel. MANGANESE (Mn): Is normally present in all steel and functions as a deoxidizer. It also imparts strength and responsiveness to heat treatment. It is usually present in quantities of 0.5 to 2.0 percent. Range 0.3% to 1.5% always present in steels to reduce the negative effects of impurities carried out forward from the production process e.g. sulphur embrittlement. It promotes the formation of stable carbides in quenched-hardened steels. Alloys containing manganese are pearlitic. Up to 1% acts as hardening agent and from 1% to 2% improves strength and toughness. Alloys containing more than 5% are non-magnetic. Alloys containing large proportions of up to 12.5% manganese have the property that they spontaneously form hard skins when subject to abrasion. (Self-hardening) All commercial steels contain 0.3-0.8% manganese, to reduce oxides and to counteract the harmful influence of iron sulfide. Any manganese in excess of these requirements partially dissolves in the iron and partly forms Mn3C which occurs with the Fe3C. There is a tendency nowadays to increase the manganese content and reduce the carbon content in order to get steel with an equal tensile strength but improved ductility If the manganese is increased above 1,8% the steel tends to become air hardened, with resultant impairing of the ductility. Up to this quantity, manganese has a beneficial effect on the mechanical properties of oil hardened and tempered 0.4% carbon steel. The manganese content is also increased in certain alloy steels, with a reduction or elimination of expensive nickel, in order to reduce costs. Steels with 0.3-0.4% carbon, 1,3-1,6% manganese and 0.3% molybdenum have replaced 3% nickel steel for some purposes. Non-shrinking tool steel contains up to 2% manganese, with 0.8-0.9% carbon. Steels with 5 to 12% manganese are martensitic after slow cooling and have little commercial importance. Hadfield`s manganese steel a specially steel which is austenitic and usually contains approximately 12% Manganese. It is used in mining, earth- moving equipment and in railroad track work. . Hadfield`s manganese steel contains 12 to 14% of manganese and 1,0% of carbon. It is characterized by a great resistance to wear and is therefore used for railway points, rock drills and stone crushers. Austenite is completely retained by quenching the steel from 1000C, in which soft condition it is used, but abrasion raises the hardness of the surface layer from 200 to 600 VPN (with no magnetic change), while the underlying material remains rough. Annealing embrittles the steel by the formation of carbides at the grain boundaries. Nickel is added to electrodes for welding manganese steel and 2% Mo sometimes added, with a prior carbide dispersion treatment at 600C, to minimize initial distortion and spreading.

Appendix D. Page 40 of 116.

Menu

CM MOK cmmok128@yahoo.com

http://www.arema.org/eseries/scriptcontent/custom/e_arema/library/2 005_Conference_Proceedings/00040.pdf

Manganese is generally beneficial to surface quality especially in resulfurized steels. Manganese contributes to strength and hardness, but less than carbon. The increase in strength is dependent upon the carbon content. Increasing the manganese content decreases ductility and weldability, but less than carbon. Manganese has a significant effect on the hardenability of steel. Manganese aids the grain structure, and contributes to hardenability, strength & wear resistance. Improves the steel (e.g. deoxidizes) during the steel's manufacturing (hot working and rolling). Present in most cutlery steel except for A-2, L-6 and CPM 420V. Manganese slightly increases the strength of ferrite, and also increases the hardness penetration of steel in the quench by decreasing the critical quenching speed. This also makes the steel more stable in the quench. Steels with manganese can be quenched in oil rather than water, and therefore are less susceptible to cracking because of a reduction in the shock of quenching. Manganese is present in most commercially made steels. Manganese is generally used in stainless steels in order to improve hot ductility. Its effect on the ferrite/austenite balance varies with temperature: at low temperature manganese is a austenite stabiliser but at high temperatures it will stabilize ferrite. Manganese increases the solubility of nitrogen and is used to obtain high nitrogen contents in austenitic steels.

Figure 1: Non-metallic inclusion in steel: oxides-dark gray and sulfides-light gray

Manganese is generally used in stainless steels in order to improve hot ductility. Its effect on the ferrite/austenite balance varies with temperature: at low temperature manganese is a austenite stabiliser but at high temperatures it will stabilise ferrite. Manganese increases the solubility of nitrogen and is used to obtain high nitrogen contents in austenitic steels.

Appendix D. Page 41 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 2: Typical duplex oxidesulfide inclusion (particle A, B and C) and plate-like MnS (particle D) in conventional continuous casting silicon steel.

SEM of an inclusion.

Appendix D. Page 42 of 116.

Menu

CM MOK cmmok128@yahoo.com

Nickel

Increase Strength. Improve Toughness. Unable to increase Hardness Ferrite Former. Increases strength and toughness but is ineffective in increasing hardness. It is generally

NICKEL (Ni):

added in amounts ranging from 1 percent to 4 percent. In some stainless steels it is sometimes as high as 20 percent. It is used for strength, corrosion resistance, and toughness, nickel increases the strength of ferrite, therefore increasing the strength of the steel. It is used in low alloy steels to increase toughness and hardenability. Nickel also tends to help reduce distortion and cracking during the quenching phase of heat treatment. Nickel is a ferrite strengthener. Nickel does not form carbides in steel. It remains in solution in ferrite, strengthening and toughening the ferrite phase. Nickel increases the hardenability and impact strength of steels. Range 0.2% to 5% Improves strength, toughness, and hardenability without seriously affecting the ductility. It encourages grain refinement. Nickel and chromium together have opposing properties and are used together to advantage in nickel-chrome steels. The resulting steels have their advantages combined and their undesirable features cancel each other At 5% nickel provides high fatigue resistance. When alloyed at higher proportions significant corrosion resistance results and at 27% a non magnetic stainless steel results. The addition of nickel, in amounts up to 8% or 10 %, to low carbon steel, increases the tensile strength and considerably raises the impact resistance. 9% nickel steels are useful at very low temperatures. In engineering steels it is widely used, often with chromium and molybdenum. High nickel increases resistance to corrosion, and in combination with chromium, is used in the austenitic corrosion-resisting steels. Certain iron-nickel alloys have unique properties. 25% nickel steel is practically non-magnetic. Alloys with about 36% nickel have very low coefficients of expansion, whilst with 50% to 78.5% nickel; alloys are obtained having very high magnetic permeability in low fields. An alloy containing 29% nickel, 17% cobalt is used for sealing with certain borosilicate glasses. For stainless steel, the main reason for the nickel addition is to promote an austenitic structure. Nickel generally increases ductility and toughness. It also reduces the corrosion rate and is thus advantageous in acid environments. In precipitation hardening steels nickel is also used to form the intermetallic compounds that are used to increase the strength. Nickel and manganese are very similar in behavior and both lower the eutectoid temperature. This change point on heating is lowered progressively with increase of nickel (approximately 10C for 1% of nickel), but the lowering of the change on cooling is greater and irregular. The temperature of this change (Ar1) is plotted for different nickel contents for 0.2% carbon steels in Fig. 1, It will be seen that the curve takes a sudden plunge round about 8% nickel. A steel with 12% nickel begins to transform below 300C on cooling, but on reheating the reverse change does not occur until about 650C. Such steels are said to exhibit pronounced lag or hysteresis and are called irreversible steels. This characteristic is made use of in maraging steels and 9% Ni

Appendix D. Page 43 of 116.

Menu

CM MOK cmmok128@yahoo.com
cryogenic steel. It Increases strength and toughness but is ineffective in increasing hardness. It is generally added in amounts ranging from 1 percent to 4 percent. In some stainless steels it is sometimes as high as 20 percent.

Maraging steels are a class of high-strength steel with low carbon content and the use of
substitutional (as opposed to interstitial) elements to produce hardening from formation of nickel martensites. The name maraging has resulted from the combination of

Martensite + Age hardening


Maraging steels contain 18% nickel, along with a amounts of molybdenum, cobalt, and titanium and aluminium, and almost no carbon. These alloys can be strengthened significantly by a Weldability is excellent. Fracture precipitation reaction at a relatively low temperature. They can be formed and machined in the solution-annealed condition but not without difficulty. toughness of the maraging steels is considerably higher than that of the conventional high-strength steels. Maraging steels are hardened by a metallurgical reaction that does not involve carbon. Maraging steels are strengthened by intermetallic compounds such as Ni 3Ti and Ni 3Mo which precipitate at about 500C. The carbon content provides no real benefit and is kept low as possible in order to minimize the formation of titanium carbide which can adversely affect mechanical properties. Toughness is superior to all low alloy carbon steels of similar strength, particularly the low temperature toughness. These steels are easy to machine and heat treat, so some cost savings result in component production to compensate for the high cost of the steel. A high strength maraging steel (extrusion section MIL-S-46850 grade 300) can have a 0.2% proof stress of 1930MPa and Ultimate Tensile strength of 2068 MPa with an elongation of 4%

The addition of nickel acts similarly to increasing the rate of cooling of a carbon steel. Thus with a constant rate of cooling the 5 to 8% nickel steels become troostitic; at 8 to 10% nickel, where the sudden drop appears, the structure is martensitic, while above 24% nickel the critical point is depressed below room temperature and austenite remains. The lines of demarcation are not so sharp as indicated by Fig. 1, but a gradual transition occurs from one constituent to another.

Appendix D. Page 44 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure1. Effect of nickel on change points and mechanical properties of 0.2% carbon steels cooled at a constant rate.

The mechanical properties change accordingly as shown in the lower part of Fig. 1. Steels with 0.5% nickel are similar to carbon steel, but are stronger, on account of the finer pearlite formed and the presence of nickel in solution in the ferrite. When 10% nickel is exceeded the steels have a high tensile strength, great hardness, but are brittle, as shown by the Izod and elongation curves. When the nickel is sufficient to produce austenite the steels become non-magnetic, ductile, tough and workable, with a drop in strength and elastic limit. Carbon intensives the action of nickel and the change points shown in Fig. 1 will vary according to the carbon content. The influences of carbon and nickel on the structure are shown in the small inset (Guillet) diagram in Fig. 1, for one rate of cooling. Steels containing 2 to 5% nickel and about 0.1% carbon are used for case hardening; those containing 0.25 to 0.40% carbons are used for crankshafts, axles and connecting rods. The superior properties of low nickel steels are best brought out by quenching and tempering (550-650C). Since the Ac3 point is lowered, a lower hardening temperature than for carbon steels is permissible and also a wider range of hardening temperatures above Ac3 without excessive grain growth, which is hindered by the slow rate of diffusion of the nickel. Martensitic nickel steels are not utilized and the austenitic alloys cannot compete with similar manganese steels owing to the higher cost. Maraging steels have fulfilled a high tensile requirement in aero and space fields. High nickel alloys are used for special purposes, owing to the marked

Appendix D. Page 45 of 116.

Menu

CM MOK cmmok128@yahoo.com
influence of nickel on the coefficient of expansion of the metal. With 36% nickel, 0.2% carbon, 0.5% manganese, the coefficient is practically zero between 0C and 100C. This alloy ages with time, but this can be minimized by heating at 100C for several days. The alloy is called Inver and it is used extensively in clocks, tapes and wire measures, differential expansion regulators, and in aluminum pistons with a split skirt in order to give an expansion approximating to that of cast iron. A carbon-free alloy containing 78.5% nickel and 21.5% iron has a high permeability in small magnetic fields.

Figure 6: Vertical section of Fe-Cr-C diagram for 0.1C wt%.

Figure 7: Schaeffler diagram for weld metals.

Appendix D. Page 46 of 116.

Menu

CM MOK cmmok128@yahoo.com

Chromium

Chromium is commonly added to steel to increase corrosion resistance and oxidation resistance, to increase hardenability, or to improve high-temperature strength. As a hardening element, Chromium is frequently used with a toughening element such as nickel to produce superior mechanical properties. At higher temperatures, chromium contributes increased strength. Chromium is strong carbide former. Complex chromium-iron carbides go into solution in austenite slowly; therefore, sufficient heating time must be allowed for prior to quenching. It is added for wear resistance, hardenability, and (most importantly) for corrosion resistance. As with manganese, chromium has a tendency to increase hardness penetration. When 5 percent chromium or more is used in conjunction with manganese, the critical quenching speed is reduced to the point that the steel becomes air hardening. Chromium can also increase the toughness of steel, as well as the wear resistance. As an alloying element in steel, chromium increases the hardenability and in association with high carbon gives resistance to abrasion and wear. 4%is present in high speed steel and up to 5% is present in hot die steels. In Structural steels it may be present in amounts up to about 3 %. Simple chromium-carbon steels are used for ball bearings having high elastic limit and high uniform hardness due to the uniform distribution of the hard carbide particles, but for most structural purposes chromium is used in conjunction with up to 4 % nickel and small amounts of molybdenum or vanadium. In heat-resisting steels, chromium is present in amounts up to 30%, and it is an important element in many of the highly alloyed heat-resisting materials, whose iron contents are so low that they cannot be regarded as steel. Chromium is also used as an alloying addition to high duty cast irons. This is the most important alloying element in stainless steels. It is this element that gives the stainless steels their basic corrosion resistance. The corrosion resistance increases with increasing chromium content. It also increases the resistance to oxidation at high temperatures. Chromium promotes a ferritic structure. Chromium is unique in its effect on resistance to corrosion and scaling and is an essential constituent in all stainless steels, e.g., stainless cutlery steels contain 12% to 14% chromium, whilst in steels of the austenitic corrosion-resisting type, 18% chromium is associated with 8% nickel, and small amounts of other elements. Steel with at least 13% chromium is deemed "stainless" steel. Despite the name, all steel can rust if not maintained properly. Chromium can dissolve in either alpha- or gama-iron, but, in the presence of carbon, the carbides formed are cementite (FeCr)3C in which chromium may rise to more than 15%; chromium carbides (CrFe)3C2 (CrFe)7C3 (CrFe)4C, in which chromium may be replaced by a few per cent, by a maximum of 55% and by 25% respectively. Stainless steels contain Cr4C. The pearlitic chromium steels with, say, 2% chromium are extremely sensitive to rate of cooling and temperature of heating before quenching; It increases the depth penetration of hardening and also the responsiveness to heat treatment. It is usually added with nickel (Ni) for use in stainless steels. Most of the chromium (Cr) bearing alloys contain 0.50 to 1.50 percent chromium; some stainless steels contain as much as 20 percent or more. It can affect forging, causing a tendency in the steel to crack.

Appendix D. Page 47 of 116.

Menu

CM MOK cmmok128@yahoo.com
For example: Critical Hardening Rate (Min. to cool from 836C to 546C) 3.5 S 6.5 S 13 S

Temperature of Initial Heating, C 836 1010 1200

The reason is that the chromium carbides are not readily dissolved in the austenite, but the amount increases with increase of temperature. The effect of the dissolved chromium is to raise the critical points on heating (Ac) and also on cooling (Ar) when the rate is slow. Faster rates of cooling quickly depress the Ar points with consequent hardening of the steel. Chromium imparts a characteristic form of the upper portion of the isothermal transformation curve. The percentage of carbon in the pearlite is lowered. Hence the proportion of free cementite (hardest constituent) is increased in high carbon steel and, when the steel is properly heat-treated, it occurs in the spheroidised form which is more suitable when the steel is used for ball bearings. The pearlite is rendered fine. When the chromium exceeds 1.1% in low-carbon steels an inert passive film is formed on the surface which resists attack by oxidizing reagents. Still higher chromium contents are found in heat-resisting steel. Chromium steels are easier to machine than nickel steels of similar tensile strength. The steels of higher chromium contents are susceptible to temper brittleness if slowly cooled from the tempering temperature through the range 550/450C. These steels are also liable to form surface markings, generally referred to as "chrome lines". The chrome steels are used wherever extreme hardness is required, such as in dies, ball bearings, plates for safes, rolls, files and tools. High chromium content is also found in certain permanent magnets.

Figure 2. Effect of alloying with chromium on the critical temperature of steel and austenite (g -iron) phase transformation zone on the iron-iron carbide diagram.

Appendix D. Page 48 of 116.

Menu

CM MOK cmmok128@yahoo.com

Nickel and chromium


Nickel steels are noted for their strength, ductility and toughness, while chromium steels are characterized by their hardness and resistance to wear. The combination of nickel and chromium produces steels having all these properties, some intensified, without the disadvantages associated with the simple alloys. The depth of hardening is increased, and with 4,5% nickel, 1,25% chromium and 0.35% carbon the steel can be hardened simply by cooling in air. Low nickel-chromium steels with small carbon content are used for casehardening, while for most constructional purposes the carbon content is 0.25-0.35%, and the steels are heat-treated to give the desired properties. Considerable amounts of nickel and chromium are used in steel for resisting corrosion and oxidation at elevated temperatures. Embattlement. The effects of tempering a nickel-chromium steel are shown in Fig. 2, from which it will be noticed that the Izod impact curve No. 1 reaches a dangerous minimum in the range 250-450C in common with many other steels. This is known as 350C embattlement. Phosphorus and nitrogen have a significant effect while other impurities (As, Sb, Sn) and manganese in larger quantity may also contribute to the embattlement.

Figure 2. Effect of tempering on the mechanical properties of nickel-chromium steel, C 0.26, Ni 3, Cr 1,2, 29 mm diam, bars hardened in oil from 830C. Izod (2) for steel with 0.25% molybdenum added Temper brittleness is usually used to describe the notch impact intergranular brittleness (Grain boundaries are revealed in temper brittle samples by etching in 1 gm cetyl trimethyl ammonium bromide; 20 gm picric acid; 100 cc distilled water, 100 cc ether. Shake mixture, allow to stand for 24 hrs; use portion of top layer and return to tube afterwards) induced in some steels by slow cooling after tempering above about 600C and also from prolonged soaking of tough material between about 400?and 550C. Temper brittleness seems to be due to grain boundary enrichment with alloying elements-Mn, Cr, Mo-during

Appendix D. Page 49 of 116.

Menu

CM MOK cmmok128@yahoo.com
austenitising which leads to enhanced segregation of embattling elements P, Sn, Sb, As-by chemical interaction on slow cooling from 600C. The return to the tough condition, obtained by rehearing embattled steel to temperatures above 600C and rapidly cooling, is due to the redistribution and retention in solution of the embattling segregation. Antimony (0-001 %), phosphorus (0-008 %), arsenic, tin, manganese increase, while molybdenum decreases the susceptibility of a steel to embattlement. 0-25 % molybdenum reduces the brittleness as shown by Izod curve No. 2. Table 1 illustrates the effect rate of cooling after tempering and the influence of an addition of 0-45 % molybdenum: Table 1. Steel 0.3% C, 3,5 % Ni, 0.7%, Cr, tempered at 630C Steel Ni-Cr Ni-Cr Ni-Cr-Mo Cooling Rate Oil Furnance Furnance TS MPa 896 880 896 Elongation 18 18 18 RA 60 60 61 Izod ft lbf 64 19 59 Izod J 87 25 80

Appendix D. Page 50 of 116.

Menu

CM MOK cmmok128@yahoo.com

Molybdenum
Molybdenum: A carbide former, prevents brittleness & maintains the steel's strength at high temperatures. Present in many steels, and air-hardening steels (e.g. A-2, ATS-34) always have 1% or more molybdenum -molybdenum is what gives those steels the ability to harden in air. It is Adds greatly to the penetration of hardness and increases toughness of an alloy. It causes steel to resist softening at high temperatures, which defeats the purpose of forging. If the alloy has below 0.020 percent molybdenum (Mo), you should be able to forge this alloy with little difficulty. Molybdenum is used very widely because of its powerful effect in increasing hardenability and also because in low alloy steels it reduces susceptibility to temper brittleness. It forms stable carbides, raises the temperature at which softening takes place on tempering and increases resistance to creep. In high speed steel it can be used to replace approximately twice its weight of tungsten. The corrosion resistance of stainless steel is improved by molybdenum additions. Molybdenum increases the hardenability of steel. Molybdenum may produce secondary hardening during the tempering of quenched steels. It enhances the creep strength of low-alloy steels at elevated temperatures. Molybdenum substantially increases the resistance to both general and localized corrosion. It increases the mechanical strength somewhat and strongly promotes a ferritic structure. Molybdenum also promotes the formation secondary phases in ferritic, ferritic-austenitic and austenitic steels. In martensitic steels it will increase the hardness at higher tempering temperatures due to its effect on the carbide precipitation. Molybdenum dissolves in both alpha- and gama-iron and in the presence of carbon forms complex carbides (FeMo)6C, Fe21Mo2C6, Mo2C. Molybdenum is similar to chromium in its effect on the shape of the TTT-curve but up to 0.5% appears to be more effective in retarding pearlite and increasing bainite formation. Additions of 0.5% molybdenum have been made to plain carbon steels to give increased strength at boiler temperatures of 400C, but the element is mainly used in combination with other alloying elements. Ni-Cr-Mo steels are widely used for ordnance, turbine rotors and other large articles, since molybdenum tends to minimize temper brittleness and reduces mass effect. Molybdenum is also a constituent in some high-speed steels, magnet alloys, heat-resisting and corrosion-resisting steels.

Appendix D. Page 51 of 116.

Menu

CM MOK cmmok128@yahoo.com

Vanadium
Ferrite Promoter. Carbide and Nitride Former.

Vanadium acts as a scavenger for oxides, forms vanadium carbide VC, and has a beneficial effect on the mechanical properties of heat-treated steels, especially in the presence of other elements. It slows up tempering in the range of 500-600C and can induce secondary hardening. Chromium-vanadium (0.15%) steels are used for locomotive forging, automobile axles, coil springs, torsion bars and creep resistance. Vanadium increases the yield strength and the tensile strength of carbon steel. The addition of small amounts of Vanadium can significantly increase the strength of steels. Vanadium is one of the primary contributors to precipitation strengthening in microalloyed steels. When thermomechanical processing TMCP is properly controlled the ferrite grain size is refined and there is a corresponding increase in toughness. The impact transition temperature also increases when vanadium is added. All microalloy steels contain small concentrations of one or more strong carbide and nitride forming elements. Vanadium, niobium, and titanium combine preferentially with carbon and/or nitrogen to form a fine dispersion of precipitated particles in the steel matrix. The presence of vanadium in steel raises the temperature at which grain coarsening sets in and under certain conditions increases the hardenability. It also lessens softening on tempering and confers secondary hardness on high speed and other steels. Vanadium carbide is intensely hard and as much as 5% vanadium may be added to high speed and high chromium tool steel where it improves abrasion resistance. Vanadium is an important constituent in many types of steel, for widely varying applications, e.g., nitriding, heat-resistance, tools, wearing plates and other fully hardened parts. In conjunction with molybdenum, vanadium has a marked effect in enhancing creep resistance. Vanadium increases the hardness of martensitic steels due to its effect on the type of carbide present. It also increases tempering resistance. Vanadium stabilises ferrite and will, at high contents, promote ferrite in the structure. It is only used in hardenable stainless steels. It retards grain growth within steel even after long exposures at high temperatures, and helps to control grain structures while heat treating. It is usually present in small quantities of 0.15 to 0.20 percent. Most tool steels which contain this element seem to absorb shock better that those that do not contain vanadium (V). Vanadium contributes to wear resistance and hardenability. A carbide former that helps produce fine-grained steel. A number of steels have vanadium, but M-2, Vascowear, and CPM T440V and 420V (in order of increasing amounts) have high amounts

Appendix D. Page 52 of 116.

Menu

CM MOK cmmok128@yahoo.com

Tungsten
TUNGSTEN (W): Also known as wolfram, is used as an alloying element in tool steels, as it tends to impart a tight, small, and dense grain pattern and keen cutting edges when used in relatively small amounts. It will also cause steel to retain its hardness at higher temperatures and hence will have a detrimental effect upon the steel's forgeability (otherwise known as "red hard") Tungsten dissolves in gama-iron and in alpha-iron. With carbon it forms WC and W2C, but in the presence of iron it forms Fe3W3C or Fe4W2C. A compound with iron, Fe3W2 provides an age-hardening system. Tungsten raises the critical points in steel and the carbides dissolve slowly over a range of temperature. When completely dissolved, the tungsten renders transformation sluggish, especially to tempering, and use is made of this in most hot-working tool ("high speed") and die steels. It Increases wear resistance. When combined properly with chromium or molybdenum, tungsten will make the steel to be a high-speed steel. The high-speed steel M-2 has a high amount of tungsten. Tungsten refines the grain size and produces less tendency to decarburisation during working. Tungsten is also used in magnet, corrosion- and heat-resisting steels.

The effect of the addition of this metal to steel is to increase the strength at normal and elevated temperatures. Owing to the hardness of tungsten carbide and its influence on secondary hardening, tungsten is used as the main alloy addition in high speed tool steels, molybdenum being its only substitute. In addition, tungsten finds considerable application in general tool steels, die and precipitation hardening steels. It has found a useful application in valves and other steels required for use at high temperatures. Tungsten is an essential constituent in the sintered hard metals.

Appendix D. Page 53 of 116.

Menu

CM MOK cmmok128@yahoo.com

Silicon

Ferrite Former. Encourage Brittleness.

Silicon Si, is one of the principal deoxidizers used in steelmaking. Silicon is less effective than manganese in increasing as-rolled strength and hardness. In low-carbon steels, silicon is generally detrimental to surface quality. Silicon increases the resistance to oxidation, both at high temperatures and in strongly oxidising solutions at lower temperatures. It promotes a ferritic structure.

It has a beneficial effect upon tensile strength and improves hardenability of an alloy. It has a toughening effect when used in combination with certain other elements. Silicon (Si) is usually added to improve electrical conductivity of an alloy. Its average concentration is between 1.5 and 2.5 percent.
Silicon is used as a deoxidizer in the manufacture of steel. It slightly increases the strength of ferrite, and when used in conjunction with other alloys can help increase the toughness and hardness penetration of steel. Silicon is a powerful deoxidizer, and as such is used in steel making processes in amounts up to about 08%. When used as an alloying element, silicon in small percentages will increase the tensile strength and yield point of structural steels. It is used in amounts of 15% to 2% in silicon-manganese spring steels and ultra-high tensile steels due to its effect in raising the limit of proportionality and resistance to tempering. Up to 4% in heat resisting steels improves scale resistance owing to the formation of a protective layer (see also Ihrigizing). The higher the silicon, the higher the temperature at which protection against further atmospheric oxidation is given. Water vapour and carbon dioxide, however, attack the layer. Alloys of iron and silicon, containing 15% of the element, are used as acid-resisting materials, but have the properties of cast irons rather than of steels. Carbon-free alloys with up to 4% silicon have a high electrical resistance and low hysteresis loss, and are used as transformer steels. In cast iron, silicon not only serves as a deoxidizer but also has a marked graphitizing effect, thus improving machinability.

Silicon dissolves in the ferrite, of which it is a fairly effective hardener, and raises the Ac change points and the Ar points when slowly cooled and also reduces the gama-alpha volume change. Only three types of silicon steel are in common use-one in conjunction with manganese for springs; the second for electrical purposes, used in sheet form for the construction of transformer cores, and poles of dynamos and motors, that demand high magnetic permeability and electrical resistance; and the third is used for automobile valves.
C 1. Silico-manganese 2. Silicon steel 3. Silichrome 0.5 0.07 0.4 Si 1,5 4,3 3,5 Mn 0.8 0.09 8

It contributes oxidation resistance in heat-resisting steels and is a general purpose deoxidizes.


Appendix D. Page 54 of 116.

Menu

CM MOK cmmok128@yahoo.com

Copper

Austenite Former. Impair Forging. Use as precipitation hardening alloy e.g17-4PH

Copper dissolves in the ferrite to a limited extent; not more than 3.5% is soluble in steels at normalizing temperatures, while at room temperature the ferrite is saturated at 0.35%. It lowers the critical points, but insufficiently to produce martensite by air cooling. The resistance to atmospheric corrosion is improved and copper steels can be temper hardened. Copper Cu, in significant amounts is detrimental to hot-working steels. Copper negatively affects forge welding, but does not seriously affect arc or oxyacetylene welding. Copper can be detrimental to surface quality. Copper is beneficial to atmospheric corrosion resistance when present in amounts exceeding 0.20%. Weathering steels are sold having greater than 0.20% Copper. The addition of about 0.20% copper to low carbon steel may increase its resistance to atmospheric corrosion by as much as 20% to 30%. In amounts of about 050% copper appreciably increases the tensile and yield strengths. The addition of increasing amounts of copper leads to defects in rolling. High yield point structural steels containing copper, in association with chromium and appreciable percentages of silicon and phosphorus have been developed. Copper is also added to some stainless steels to improve corrosion resistance. Copper enhances the corrosion resistance in certain acids and promotes an austenitic structure. In precipitation hardening steels copper is used to form the intermetallic compounds that are used to increase the strength.

Phosphorous

Embrittlement Effects Increase Machinability. Undesirable Element.

Phosphorus P, increases strength and hardness and decreases ductility and notch impact toughness of steel. The adverse effects on ductility and toughness are greater in quenched and tempered higher-carbon steels. Phosphorous levels are normally controlled to low levels. Higher phosphorus is specified in low-carbon free-machining steels to improve machinability. Although it has been used to increase the tensile strength of steel and to improve resistance to atmospheric corrosion, phosphorus is usually regarded as an undesirable impurity because of its embrittling effect. In most British specifications the maximum permitted is 0.05 %, but in steel for

Appendix D. Page 55 of 116.

Menu

CM MOK cmmok128@yahoo.com
nitriding it may be restricted to a maximum of 002 % since during the nitriding treatment phosphorus has a temper embrittling effect.

Sulphur
Increase Machinability. Undesirable element. Embrittlement Effects. Impair Ductility. Impair Weldability.

SULFUR S, Is usually regarded as an impurity in most alloys and its addition to steel is held to a minimum as it is damaging to the hot forming characteristics of steel. It is, however added to increase machinability. A word of caution, some alloys are offered in different forms, an example is E52100. This particular steel can be had in either a "Bearing Quality" or "Machining Quality" the latter having sulfur added to increase machinability. Sulfur decreases ductility and notch impact toughness especially in the transverse direction. Weldability decreases with increasing sulfur content. Sulfur is found primarily in the form of sulfide inclusions. Sulfur levels are normally controlled to low levels. The only exception is free-machining steels, where sulfur is added to improve machinability. A non-metal, which combines with iron to form iron sulphides, in which form its effect is to make the steel red short but combined with manganese its influence is less injurious. In steel the sulphur content is usually specified as less than 0.05 % but it may be added deliberately to improve machinability. Sulphur is added to certain stainless steels, the free-machining grades, in order to increase the machinability. At the levels present in these grades sulphur will substantially reduce corrosion resistance, ductility and fabrication properties, such as weldability and formability.

Appendix D. Page 56 of 116.

Menu

CM MOK cmmok128@yahoo.com

Lead

Improve Machinability. Undesirable Element. Impair Ductility. Impair Toughness. Impair Creep Strength.

Lead Pb, increase the machinability of steel and has no effect upon the other properties of the metal. Lead is virtually insoluble in liquid or solid steel. However, lead is sometimes added to carbon and alloy steels by means of mechanical dispersion during pouring to improve the machinability. The addition of about 025% lead improves machinability. It also causes a reduction in fatigue strength, ductility and toughness but this only becomes serious in the transverse direction and at high tensile levels. In creep resisting alloys very small amounts of lead may be harmful.

Hydrogen

Hydrogen H, in steel is an undesirable impurity which is introduced from moisture in the atmosphere or the charge during melting. If a large amount of hydrogen is present in the liquid steel, some may be liberated on freezing giving an unsound ingot, evolution of hydrogen subsequently when the solid steel cools may cause hair line cracks. Hydrogen can be reduced to safe proportions by casting in vacuum or by prolonged annealing. It may also be introduced into steel by electrolytic action or by pickling and may then cause brittleness.

Selenium

Selenium Se, A metalloid closely resembling sulphur in its properties. It is sometimes added to steels to the extent of 02 % to 03 % to improve machinability.

Appendix D. Page 57 of 116.

Menu

CM MOK cmmok128@yahoo.com

Tantalum

Tantalum Ta, This metal is associated with niobium and is very similar to it chemically. As an alloying addition to steel, niobium is preferred. Tellurium is added to steel either alone or together with selenium to promote machinability. It is a powerful carbide stabilizer and has been also added to cast iron where it is said to increase the depth of chill and to prevent shrinkage. It may be added in small amounts to the molten iron or by the use of cores dipped or painted with washes containing tellurium in suspension.

Tin

Tin Sn, Owing to its good resistance to corrosion in many conditions, the major use of tin is in the form of coatings for steel and copper alloys. It is an undesirable impurity in steel giving rise to temper brittleness, but is less harmful than phosphorus.

Calcium

Calcium Ca, This metal in the form of calcium silicide is sometimes added to steel as a deoxidizer and degasefier.

Cerium

Cerium Ce, a metal of the rare earth class which in many respects resemble the alkali metals. The hot working properties of high alloy corrosion- and heat-resistant steels maybe improved by the addition of cerium, whilst in cast iron, cerium acts as a deoxidizer and desulphurizer but when the sulphur content has been reduced to a value of about 0015%, the cerium enters into solution in the cast iron and functions as a powerful carbide stabilizer. In amounts above 002%, cerium is the operative factor in the production of nodular graphite structures in cast iron. Cerium is one of the rare earth metals (REM) and is added in small amounts to certain heat resistant temperature steels and alloys in order to increase the resistance to oxidation and high temperature corrosion.

Appendix D. Page 58 of 116.

Menu

CM MOK cmmok128@yahoo.com

Nitrogen

Nitrogen N, Nitrogen can combine with many metals to form nitrides and is thus applied to the case hardening of steel, the usual source for this purpose being ammonia. The incorporation of nitrogen in austenitic chromium-nickel steels stabilizes the austenite and increases the strength. In carbon steels it has an influence on creep. (See Abnormal Steels) Nitrogen is a very strong austenite former and strongly promotes an austenitic structure. It also substantially increases the mechanical strength. Nitrogen increases the resistance to localised corrosion, especially in combination with molybdenum. In ferritic stainless steels nitrogen will strongly reduce toughness and corrosion resistance. In the martensitic and martensitic-austenitic steels nitrogen increases both hardness and strength but reduces the toughness.

Boron

Boron B, is added to fully killed steel to improve hardenability. Boron-treated steels are produced to a range of 0.0005 to 0.003%. Whenever boron is substituted in part for other alloys, it should be done only with hardenability in mind because the lowered alloy content may be harmful for some applications. The addition of about 0.003 % of boron confers increased harden ability to steels in the quenched and tempered condition. Further, it has been found that the addition of 0003% boron to low carbon, 050% molybdenum steel in the normalized condition doubles the yield strength and gives a 30% increase in tensile strength, but the advantage due to boron is very slight when molybdenum is less than 0.35% causes difficulty in forging. As much as 2% may be added to steels used in nuclear engineering. Boron is a potent alloying element in steel. A very small amount of boron (about 0.001%) has a strong effect on hardenability. Boron steels are generally produced within a range of 0.0005 to 0.003%. effective in lower carbon steels. Boron. In recent years, especially in USA, 0.003-0.005% boron has been added to previously fully killed, fine-grain steel to increase the hardenability of the steel. The yield ratio and impact are definitely improved, provided advantage is taken of the increased hardenability obtained and the steel is fully hardened before tempering. In conjunction with molybdenum boron forms a useful group of high tensile bainitic steels. Boron is used in some hard facing alloys and for nuclear control rods. Boron is most

Appendix D. Page 59 of 116.

Menu

CM MOK cmmok128@yahoo.com

Aluminium

Aluminum Al, is widely used as a deoxidizer. Aluminum can control austenite grain growth in reheated steels and is therefore added to control grain size. Aluminum is the most effective alloy in controlling grain growth prior to quenching. Titanium, zirconium, and vanadium are also valuable grain growth inhibitors, but there carbides are difficult to dissolve into solution in austenite. As a deoxidizer, up to 0.05% aluminum may be added to steel. For increasing fine grain characteristics or sub-zero impact properties, up to 010% may be added. Nitriding steels contain about 1% aluminum for promoting a high surface hardness when heated in ammonia. Still larger additions made to heat resisting steels promote resistance to scaling. Approximately 5% added to chromium steel increases electrical resistivity. Aluminium improves oxidation resistance, if added in substantial amounts. It is used in certain heat resistant alloys for this purpose. In precipitation hardening steels aluminium is used to form the intermetallic compounds that increase the strength in the aged condition.

Zirconium

Zirconium Zr, can be added to killed high-strength low-alloy steels to achieve improvements in inclusion characteristics. Zirconium causes sulfide inclusions to be globular rather than elongated thus improving toughness and ductility in transverse bending. Zirconium acts as a deoxidizing element in steel and combines with the sulphur.

Niobium

Niobium Nb (Columbium) increases the yield strength and, to a lesser degree, the tensile strength of carbon steel. The addition of small amounts of Niobium can significantly increase the yield strength of steels. Niobium can also have a moderate precipitation strengthening effect. Its main contributions are to form precipitates above the transformation temperature, and to retard the recrystallization of austenite, thus promoting a fine-grain microstructure having improved strength and toughness. The metal is also known as columbium. It occurs in association with tantalum, to which it is closely related. Niobium is a strong carbide-forming element and as such is added to certain austenitic corrosion-resistant steels of the 18/8 chromium-nickel type for the prevention of intercrystalline corrosion. Where niobium is used as the stabilizer, it is usually specified that it should be present in an amount at least 8 times that of the carbon content. Further, niobium is often used as a constituent of the electrodes used in the welding of such steels. Niobium is added to heat-resisting steels and

Appendix D. Page 60 of 116.

Menu

CM MOK cmmok128@yahoo.com
enhances creep strength. In small amounts, of the order of 005%, it increases the yield strength of mild steel. Niobium is both a strong ferrite and carbide former. As titanium it promotes a ferritic structure. In austenitic steels it is added to improve the resistance to intergranular corrosion but it also enhances mechanical properties at high temperatures. In martensitic steels niobium lowers the hardness and increases the tempering resistance. In U.S. it is also referred to as Columbium (Cb).

Titanium

Titanium Ti, is used to retard grain growth and thus improve toughness. Titanium is also used to achieve improvements in inclusion characteristics. Titanium causes sulfide inclusions to be globular rather than elongated thus improving toughness and ductility in transverse bending. Titanium is a strong ferrite former and a strong carbide former, thus lowering the effective carbon content and promoting a ferritic structure in two ways. In austenitic steels it is added to increase the resistance to intergranular corrosion but it also increases the mechanical properties at high temperatures. In ferritic stainless steels titanium is added to improve toughness and corrosion resistance by lowering the amount of interstitials in solid solution. In martensitic steels titanium lowers the martensite hardness and increases the tempering resistance. In precipitation hardening steels titanium is used to form the intermetallic compounds that are used to increase the strength. The principal use of titanium is to stabilize carbon by forming titanium carbide. In austenitic stainless steels it is used in this way to prevent inter crystalline corrosion, the titanium addition being at least four times the carbon content. It is also added to low carbon steels to prevent blistering during vitreous enameling. Titanium carbide is used with tungsten carbide in the manufacture of hard metal tools.

Cobalt

Cobalt Co, Increases strength and hardness, permits quenching at higher temperatures. In some steels used for nuclear engineering cobalt is an undesirable impurity, even in amounts as small as 002%. Unlike most other alloying elements cobalt reduces hardenability. It raises the red hardness of steel and this is the reason for adding 5% to 10% cobalt to certain types of high speed steels, developed for the specific purpose of cutting exceptionally hard materials. Heat resisting alloys with high cobalt contents have been developed for use in gas turbines. Cobalt is added to the extent of up to 40 % to magnet steels requiring high coercive force and it is used in electrical- resistance alloys. In the sintered hard metals Cobalt acts as the binding metal. Cobalt only used as an alloying element in martensitic steels where it increases the hardness and tempering resistance, especially at higher temperatures.

Appendix D. Page 61 of 116.

Menu

CM MOK cmmok128@yahoo.com
Cobalt has a high solubility in alpha- and gama-iron but a weak carbide-forming tendency. It decreases hardenability but sustains hardness during tempering. It is used in "Stellite" type alloys, gas turbine steel, magnets and as a bond in hard metal.

Appendix D. Page 62 of 116.

Menu

CM MOK cmmok128@yahoo.com

Carbon content, steel classifications, and alloy steels


Generally, carbon is the most important commercial steel alloy. Increasing carbon content increases hardness and strength and improves hardenability. But carbon also increases brittleness and reduces weldability because of its tendency to form martensite. This means carbon content can be both a blessing and a curse when it comes to commercial steel. And while there are steels that have up to 2 percent carbon content, they are the exception. Most steel contains less than 0.35 percent carbon. To put this in perspective, keep in mind thats 35/100 of 1 percent. Now, any steel in the 0.35 to 1.86 percent carbon content range can be hardened using a heat-quench-temper cycle. Most commercial steels are classified into one of three groups: 1. Plain carbon steels 2. Low-alloy steels 3. High-alloy steels Plain Carbon Steels These steels usually are iron with less than 1 percent carbon, plus small amounts of manganese, phosphorus, sulfur, and silicon. The weldability and other characteristics of these steels are primarily a product of carbon content, although the alloying and residual elements do have a minor influence. Plain carbon steels are further subdivided into four groups: 1. Low 2. Medium 3. High 4. Very high Low. Often called mild steels, low-carbon steels have less than 0.30 percent carbon and are the most commonly used grades. They machine and weld nicely and are more ductile than higher-carbon steels. Medium. Medium-carbon steels have from 0.30 to 0.45 percent carbon. Increased carbon means increased hardness and tensile strength, decreased ductility, and more difficult machining. High. With 0.45 to 0.75 percent carbon, these steels can be challenging to weld. Preheating, postheating (to control cooling rate), and sometimes even heating during welding become necessary to produce acceptable welds and to control the mechanical properties of the steel after welding. Very High. With up to 1.50 percent carbon content, very high-carbon steels are used for hard steel products such as metal cutting tools and truck springs. Like high-carbon steels, they require heat treating before, during, and after welding to maintain their mechanical properties. Low-alloy Steels When these steels are designed for welded applications, their carbon content is usually below 0.25 percent and often below 0.15 percent. Typical alloys include nickel, chromium, molybdenum, manganese, and silicon, which add strength at room temperatures and increase low-temperature notch toughness.

Appendix D. Page 63 of 116.

Menu

CM MOK cmmok128@yahoo.com
These alloys can, in the right combination, improve corrosion resistance and influence the steels response to heat treatment. But the alloys added can also negatively influence crack susceptibility, so its a good idea to use low-hydrogen welding processes with them. Preheating might also prove necessary. This can be determined by using the carbon equivalent formula, which well cover in a later issue. High-alloy Steels For the most part, were talking about stainless steel here, the most important commercial high-alloy steel. Stainless steels are at least 12 percent chromium and many have high nickel contents. The three basic types of stainless are: 1. Austenitic 2. Ferritic 3. Martensitic Martensitic stainless steels make up the cutlery grades. They have the least amount of chromium, offer high hardenability, and require both pre- and postheating when welding to prevent cracking in the heat-affected zone (HAZ). Ferritic stainless steels have 12 to 27 percent chromium with small amounts of austenite-forming alloys. Austenitic stainless steels offer excellent weldability, but austenite isnt stable at room temperature. Consequently, specific alloys must be added to stabilize austenite. The most important austenite stabilizer is nickel, and others include carbon, manganese, and nitrogen. Special properties, including corrosion resistance, oxidation resistance, and strength at high temperatures, can be incorporated into austenitic stainless steels by adding certain alloys like chromium, nickel, molybdenum, nitrogen, titanium, and columbium. And while carbon can add strength at high temperatures, it can also reduce corrosion resistance by forming a compound with chromium. Its important to note that austenitic alloys cant be hardened by heat treatment. That means they dont harden in the welding HAZ.

Appendix D. Page 64 of 116.

Menu

CM MOK cmmok128@yahoo.com

* Stainless steels always have a high chromium content, often considerable amounts of nickel, and sometimes contain molybdenum and other elements. Stainless steels are identified by a three-digit number beginning with 2, 3, 4, or 5.

Figure 1 Be sure to check the appropriate AISI and SAE publications for the latest revisions.

Steel Classification Systems Before we look at a couple of common steel classification systems, lets consider one more high-carbon metal, cast iron. The carbon content of cast iron is 2.1 percent or more. There are four basic types of cast iron:

1. Gray cast iron, which is relatively soft. Its easily machined and welded, and youll find it used for engine cylinder blocks, pipe, and machine tool structures. 2. White cast iron, which is hard, brittle, and not weldable. It has a compressive strength of more than 200.000 pounds per square inch (PSI), and when its annealed, it becomes malleable cast iron.

Appendix D. Page 65 of 116.

Menu

CM MOK cmmok128@yahoo.com
3. Malleable cast iron, which is annealed white cast iron. It can be welded, machined, is ductile, and offers good strength and shock resistance. 4. Ductile cast iron, which is sometimes called nodular or spheroidal graphite cast iron. It gets this name because its carbon is in the shape of small spheres, not flakes. This makes it both ductile and malleable. Its also weldable. Now lets take a look at a typical steel classification system (see Figure 1). Both the Society of Automotive Engineers (SAE) and the American Iron and Steel Institute (AISI) use virtually identical systems. Both are based on a four-digit system with the first number usually indicating the basic type of steel and the first two numbers together indicating the series within the basic alloy group. Keep in mind there may be a number of series within a basic alloy group, depending on the amount of the principal alloying elements. The last two or three numbers refer to the approximate permissible range of carbon content in points (hundredths of a percent). These classification systems can become fairly complex, and Figure 1 is just a basic representation. Be sure to reference the most recent AISI and SAE publications for the latest revisions. Thats a look at some basics concerning the iron-carbon-steel relationship and its influences on welding and metal alloys. Next time well look at hardening and ways to make metals stronger. Well also consider the influences of some key alloying elements and the effects of welding on metallurgy.

Carbon Steel

Carbon steels and alloy steels are designated by a four digit number, where the first two digits indicate the alloying elements and the last two digits indicate the amount of carbon, in hundredths of a percent by weight. For example, a 1060 steel is a plain carbon steel containing 0.60 wt% C.
designation Carbon steels 10xx 11xx 12xx 15xx Manganese steels 13xx Nickel steels 23xx 25xx Nickel-chromium steels Ni 3.50% Ni 5.00% Mn 1.75% Plain carbon (Mn 1.00% max) Resulfurized Resulfurized and rephosphorized Plain carbon (Mn 1.00% to 1.65%) Type

Appendix D. Page 66 of 116.

Menu

CM MOK cmmok128@yahoo.com
31xx 32xx 33xx 34xx Molybdenum steels 40xx 44xx Mo 0.20% or 0.25% or 0.25% Mo & 0.042 S Mo 0.40% or 0.52% Ni 1.25%, Cr 0.65% or 0.80% Ni 1.25%, Cr 1.07% Ni 3.50%, Cr 1.50% or 1.57% Ni 3.00%, Cr 0.77%

Chromium-molybdenum (Chromoly) steels 41xx Cr 0.50% or 0.80% or 0.95%, Mo 0.12% or 0.20% or 0.25% or 0.30%

Nickel-chromium-molybdenum steels 43xx 43BVxx 47xx 81xx 81Bxx 86xx 87xx 88xx 93xx 94xx 97xx 98xx Ni 1.82%, Cr 0.50% to 0.80%, Mo 0.25% Ni 1.82%, Cr 0.50%, Mo 0.12% or 0.35%, V 0.03% min Ni 1.05%, Cr 0.45%, Mo 0.20% or 0.35% Ni 0.30%, Cr 0.40%, Mo 0.12% Ni 0.30%, Cr 0.45%, Mo 0.12% Ni 0.55%, Cr 0.50%, Mo 0.20% Ni 0.55%, Cr 0.50%, Mo 0.25% Ni 0.55%, Cr 0.50%, Mo 0.35% Ni 3.25%, Cr 1.20%, Mo 0.12% Ni 0.45%, Cr 0.40%, Mo 0.12% Ni 0.55%, Cr 0.20%, Mo 0.20% Ni 1.00%, Cr 0.80%, Mo 0.25%

Nickel-molybdenum steels 46xx 48xx Chromium steels 50xx 50xxx 50Bxx 51xx 51xxx 51Bxx Cr 0.27% or 0.40% or 0.50% or 0.65% Cr 0.50%, C 1.00% min Cr 0.28% or 0.50% Cr 0.80% or 0.87% or 0.92% or 1.00% or 1.05% Cr 1.02%, C 1.00% min Cr 0.80% Ni 0.85% or 1.82%, Mo 0.20% or 0.25% Ni 3.50%, Mo 0.25%

Appendix D. Page 67 of 116.

Menu

CM MOK cmmok128@yahoo.com
52xxx Cr 1.45%, C 1.00% min

Chromium-vanadium steels 61xx Cr 0.60% or 0.80% or 0.95%, V 0.10% or 0.15% min

Tungsten-chromium steels 72xx W 1.75%, Cr 0.75%

Silicon-manganese steels 92xx Si 1.40% or 2.00%, Mn 0.65% or 0.82% or 0.85%, Cr 0.00% or 0.65%

High-strength low-alloy steels 9xx xxBxx xxLxx Various SAE grades Boron steels Leaded steels

Stainless steel

200 Series: austenitic chromium-nickel-manganese alloys 300 Series: austenitic chromium-nickel alloys Type 301: highly ductile, for formed products. Also hardens rapidly during mechanical working. Type 303: free machining version of 304 via addition of sulfur Type 304: the most common; the classic 18/8 stainless steel. Type 316: the next most common; for food and surgical stainless steel uses; alloy addition of molybdenum prevents specific forms of corrosion. 316 steel is more resistant to corrosion than 18-8 stainless steels. 316 steel is used in the handling of certain food and pharmaceutical products where it is often required in order to minimize metallic contamination. 316 steel is also known as "marine grade" stainless steel due to its increased ability to resist saltwater corrosion compared to type 304. SS316 is often used for building nuclear reprocessing plants.

400 Series: ferritic and martensitic chromium alloys Type 408: heat-resistant; poor corrosion resistance; 11% chromium, 8% nickel. Type 409: cheapest type; used for automobile exhausts; ferritic (iron/chromium only). Type 410: martensitic (high-strength iron/chromium). Type 416: the most machinable stainless steel; achieved by the addition of extra sulfur which reduces corrosion resistance. Often used for "stainless" rifle barrels Type 420: "Cutlery grade" martensitic; similar to the Brearley's original "rustless steel". Also known as "surgical steel". Type 430: decorative, e.g., for automotive trim; ferritic. Type 440: a higher grade of cutlery steel, with more carbon in it, which allows for much better edge retention when the steel is heat treated properly.

500 Series: heat resisting chromium alloys 600 Series: martensitic precipitation hardening alloys

Appendix D. Page 68 of 116.

Menu

CM MOK cmmok128@yahoo.com
Type 630: most common PH stainless, better known as 17-4; 17% chromium, 4% nickel

Stainless steel designations SAE UNS % Cr % Ni %C % Mn % Si %P %S %N Other

designation designation Austenitic 201 202 205 301 302 302B 303 303Se 304 304L 304Cu 304N 305 308 309 309S 310 310S 314 316 316L 316F 316N 317 317L S20100 S20200 S20500 S30100 S30200 S30215 S30300 S30323 S30400 S30403 S30430 S30451 S30500 S30800 S30900 S30908 S31000 S31008 S31400 S31600 S31603 S31620 S31651 S31700 S31703

1618 1719 16.518 1618 1719 1719 1719 1719 1820 1820 1719 1820 1719 1921 2224 2224 2426 2426 2326 1618 1618 1618 1618 1820 1820

3.55.5 46 11.75 68 810 810 810 810 810.50 812 810 810.50

0.15 0.15

5.57.5

0.75

0.06 0.06 0.06

0.03 0.03 0.03

0.25 0.25

7.510.0 0.75

0.120.25 1415.5 0.75 0.15 0.15 0.15 0.15 0.15 0.08 0.03 0.08 0.08 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 0.75 0.75 2.03.0 1 1 0.75 0.75 0.75 0.75 0.75 1 1 1 1.5 1.5 1.53.0 0.75 0.75 1 0.75 0.75 0.75

0.320.40 0.1 0.1 0.1 Mo (optional) 0.15 Se min 34 Cu 0.60

0.045 0.03 0.045 0.03 0.045 0.03 0.2 0.2 0.15 min 0.06

0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.045 0.03 0.2 0.10 min

0.100.16 0.10 0.10 2.03.0 Mo 2.03.0 Mo 1.752.50 Mo

10.5013 0.12 1012 1215 1215 1922 1922 1922 1014 1014 1014 1014 1115 1115 0.08 0.2 0.08 0.25 0.08 0.25 0.08 0.03 0.08 0.08 0.08 0.03

0.045 0.03 0.045 0.03 0.045 0.03

0.100.16 2.03.0 Mo 0.10 max 3.04.0 Mo 0.10 max 3.04.0 Mo

Appendix D. Page 69 of 116.

Menu

CM MOK cmmok128@yahoo.com
Ti 5(C+N)

321 329 330 347

S32100 S32900 N08330 S34700

1719 2328 1720 1719

912 2.55 3437 913

0.08 0.08 0.08 0.08

2 2 2 2

0.75 0.75

0.045 0.03 0.04 0.03 0.03

0.10 max -

min, 0.70 max 12 Mo Nb + Ta, 10 x C min, 1 max Nb + Ta, 10 x

0.751.50 0.04 0.75

0.045 0.030 -

348

S34800

1719

913

0.08

0.75

0.045 0.030 -

C min, 1 max, but 0.10 Ta max; 0.20 Ca

384 Ferritic 405

S38400

1517

1719

0.08

0.045 0.03

S40500

11.514.5

0.08

0.04

0.03

0.10.3 0.60 max

Al,

409 429 430 430F 430FSe 434

S40900 S42900 S43000 S43020 S43023 S43400

10.511.75 0.05 1416 1618 1618 1618 1618 0.75 0.75 -

0.08 0.12 0.12 0.12 0.12 0.12

1 1 1 1.25 1.25 1

1 1 1 1 1 1

0.045 0.03 0.04 0.04 0.06 0.06 0.04 0.03 0.03 0.15 min 0.06 0.03

Ti 6 x C, but 0.75 max 0.60 (optional) 0.15 Se min 0.751.25 Mo 0.751.25 Mo

436

S43600

1618

0.12

0.04

0.03

Mo; Nb+Ta 5 x C min, 0.70 max

442 446 Martensitic 403 410 414 416 416Se

S44200 S44600

1823 2327

0.25

0.2 0.2

1 1.5

1 1

0.04 0.04

0.03 0.03

S40300 S41000 S41400 S41600 S41623

11.513.0 11.513.5 11.513.5 1214 1214

0.60 0.75

0.15 0.15

1 1 1 1.25 1.25

0.5 1 1 1 1

0.04 0.04 0.04 0.06 0.06

0.03 0.03 0.03 0.15 min 0.06

0.060 (optional) 0.15 Se min Mo

1.252.50 0.15 0.15 0.15

Appendix D. Page 70 of 116.

Menu

CM MOK cmmok128@yahoo.com
420 420F S42000 S42020 1214 1214 0.15 min 0.15 min 1 1.25 1 1 0.04 0.06 0.03 0.15 min 0.60 Mo max (optional) 0.901.25 422 S42200 11.012.5 0.501.0 0.200.25 0.51.0 0.5 0.025 0.025 Mo; 0.200.30 V; 0.901.25 W 431 440A 440B 440C S41623 S44002 S44003 S44004 1517 1618 1618 1618 1.252.50 0.2 1 1 1 1 1 0.04 0.04 0.04 0.04 0.03 0.03 0.03 0.03 0.75 Mo 0.75 Mo 0.75 Mo

0.600.75 1 0.750.95 1 0.951.20 1

Heat resisting 501 502 S50100 S50200 46 46 0.10 min 0.1 1 1 1 1 0.04 0.04 0.03 0.03 0.400.65 Mo 0.400.65 Mo

Unified numbering system

Introduction to the Unified Numbering System of Ferrous Metals and Alloys http://www.key-to-steel.com/Articles/Art111.htm http://en.wikipedia.org/wiki/Unified_numbering_system

Appendix D. Page 71 of 116.

Menu

CM MOK cmmok128@yahoo.com

Strength in Steel

Strength in steels arises from several phenomena, which usually contribute collectively to the observed mechanical properties. The heat treatment of steels is aimed at adjusting these contributions so that the required balance of mechanical properties is achieved. Fortunately the / phase change allows great variations in microstructure to be produced, so that a wide range of mechanical properties can be obtained even in plain carbon steels. The additional use of metallic alloying elements, primarily as a result of their influence on the transformation, provides an even greater control over microstructure, with consequent benefits in the mechanical properties

The simplest version of analyzes the effects of alloying elements on iron-carbon alloys would require analysis of a large number of ternary alloy diagrams over a wide temperature range. However, Wever pointed out that iron binary equilibrium systems fall into four main categories (Fig. 1): open and closed -field systems, and expanded and contracted -field systems. This approach indicates that alloying elements can influence the equilibrium diagram in two ways: by expanding the -field, and encouraging the formation of austenite over wider compositional limits. These elements are called -stabilizers. by contracting the -field, and encouraging the formation of ferrite over wider compositional limits. These elements are called -stabilizers. The form of the diagram depends to some degree on the electronic structure of the alloying elements which is reflected in their relative positions in the periodic classification. Class 1: open -field. To this group belong the important steel alloying elements nickel and manganese, as well as cobalt and the inert metals ruthenium, rhodium, palladium, osmium, iridium and platinum. Both nickel and manganese, if added in sufficiently high concentration, completely eliminate the bcc -iron phase and replace it, down to room temperature, with the -phase. So nickel and manganese depress the phase transformation from to to lower temperatures (Fig. 1a), i.e. both Ac1 and Ac3 are lowered. It is also easier to obtain metastable austenite by quenching from the -region to room temperature, consequently nickel and manganese are useful elements in the formulation of austenitic steels. Class 2: expanded -field. Carbon and nitrogen are the most important elements in this group. The -phase field is expanded, but its range of existence is cut short by compound formation (Fig.1b). Copper, zinc and gold have a similar influence. The expansion of the -field by carbon, and nitrogen, underlies the whole of the heat treatment of steels, by allowing formation of a homogeneous solid solution (austenite) containing up to 2.0 wt % of carbon or 2.8 wt % of nitrogen. Class 3: closed -field. Many elements restrict the formation of -iron, causing the -area of the diagram to contract to a small area referred to as the gamma loop (Fig. 1c). This means that the relevant elements are encouraging the formation of bcc iron (ferrite), and one result is that the - and -phase fields become continuous. Alloys in which this has taken place are, therefore, not amenable to the normal heat treatments involving cooling through the /-phase transformation. Silicon, aluminium, beryllium and phosphorus fall into this category, together with the strong carbide forming elements, titanium, vanadium, molybdenum and

Appendix D. Page 72 of 116.

Menu

CM MOK cmmok128@yahoo.com
chromium.

Figure 1. Classification of iron alloy phase diagrams: a. open -field; b. expanded -field; c. closed -field Class 4: contracted y-field. Boron is the most significant element of this group, together with the carbide forming elements tantalum, niobium and zirconium. The -loop is strongly contracted, but is accompanied by compound

Appendix D. Page 73 of 116.

Menu

CM MOK cmmok128@yahoo.com
formation (Fig. 1d). The distribution of alloying elements in steels. Although only binary systems have been considered so far, when carbon is included to make ternary systems the same general principles usually apply. For a fixed carbon content, as the alloying clement is added the y-field is either expanded or contracted depending on the particular solute. With an element such as silicon the -field is restricted and there is a corresponding enlargement of the -field. If vanadium is added, the -field is contracted and there will be vanadium carbide in equilibrium with ferrite over much of the ferrite field. Nickel does not form a carbide and expands the -field. Normally elements with opposing tendencies will cancel each other out at the appropriate combinations, but in some cases anomalies occur. For example, chromium added to nickel in a steel in concentrations around 18% helps to stabilize the -phase, as shown by 18Cr8Ni austenitic steels. One convenient way of illustrating quantitatively the effect of an alloying element on the -phase field of the Fe-C system is to project on to the Fe-C plane of the ternary system the -phase field boundaries for increasing concentration of a particular alloying element. For more precise and extensive information, it is necessary to consider series of isothermal sections in true ternary systems Fe-C-X, but even in some of the more familiar systems the full information is not available, partly because the acquisition of accurate data can be a difficult and very time-consuming process. Recently the introduction of computer-based methods has permitted the synthesis of extensive thermochemical and phase equilibria data, and its presentation in the form, for example, of isothermal sections over a wide range of temperatures. If only steels in which the austenite transforms to ferrite and carbide on slow cooling are considered, the alloying elements can be divided into three categories: elements which enter only the ferrite phase elements which form stable carbides and also enter the ferrite phase elements which enter only the carbide phase. In the first category there are elements such as nickel, copper, phosphorus and silicon which, in transformable steels, are normally found in solid solution in the ferrite phase, their solubility in cementite or in alloy carbides being quite low. The majority of alloying elements used in steels fall into the second category, in so far as they are carbide formers and as such, at low concentrations, go into solid solution in cementite, but will also form solid solutions in ferrite. At higher concentrations most will form alloy carbides, which are thermodynamically more stable than cementite. Typical examples are manganese, chromium, molybdenum, vanadium, titanium, tungsten and niobium. Manganese carbide is not found in steels, but instead manganese enters readily into solid solution in Fe3C. The carbide-forming elements are usually present greatly in excess of the amounts needed in the carbide phase, which are determined primarily by the carbon content of the steel. The remainder enters into solid solution in the ferrite with the non-carbide forming elements nickel and silicon. Some of these elements, notably titanium, tungsten, and molybdenum, produce substantial solid solution hardening of ferrite. In the third category there are a few elements which enter predominantly the carbide phase. Nitrogen is the most important element and it forms carbo-nitrides with iron and many alloying elements. However, in the presence of certain very strong nitride forming elements, e.g. titanium and aluminum, separate alloy nitride phases can occur. While ternary phase diagrams, Fe-C-X, can be particularly helpful in understanding the phases which can exist in simple steels, isothermal sections for a number of temperatures are needed before an adequate picture of the equilibrium phases can be built up. For more complex steels the task is formidable and equilibrium diagrams can

Appendix D. Page 74 of 116.

Menu

CM MOK cmmok128@yahoo.com
only give a rough guide to the structures likely to be encountered. It is, however, possible to construct pseudobinary diagrams for groups of steels, which give an overall view of the equilibrium phases likely to be encountered at a particular temperature. Structural changes resulting from alloying additions. The addition to iron-carbon alloys of elements such as nickel, silicon, manganese, which do not form carbides in competition with cementite, does not basically alter the microstructures formed after transformation. However, in the case of strong carbide-forming elements such as molybdenum, chromium and tungsten, cementite will be replaced by the appropriate alloy carbides, often at relatively low alloying element concentrations. Still stronger carbide forming elements such as niobium, titanium and vanadium are capable of forming alloy carbides, preferentially at alloying concentrations less than 0.1 wt%. It would, therefore, be expected that the microstructures of steels containing these elements would be radically altered. It has been shown how the difference in solubility of carbon in austenite and ferrite leads to the familiar ferrite/cementite aggregates in plain carbon steels. This means that, because the solubility of cementite in austenite is much greater than in ferrite, it is possible to redistribute the cementite by holding the steel in the austenite region to take it into solution, and then allowing transformation to take place to ferrite and cementite. Examining the possible alloy carbides, and nitrides, in the same way, shows that all the familiar ones are much less soluble in austenite than is cementite. Chromium and molybdenum carbides are not included, but they are substantially more soluble in austenite than the other carbides. Detailed consideration of such data, together with practical knowledge of alloy steel behavior, indicates that, for niobium and titanium, concentrations of greater than about 0.25 wt % will form excess alloy carbides which cannot be dissolved in austenite at the highest solution temperatures. With vanadium the limit is higher at 1-2%, and with molybdenum up to about 5%. Chromium has a much higher limit before complete solution of chromium carbide in austenite becomes difficult. This argument assumes that sufficient carbon is present in the steel to combine with the alloying element. If not, the excess metallic element will go into solid solution both in the austenite and the ferrite. In general, the fibrous morphology represents a closer approach to an equilibrium structure so it is more predominant in steels which have transformed slowly. In contrast, the interphase precipitation and dislocation nucleated structures occur more readily in rapidly transforming steels, where there is a high driving force, for example, in microalloyed steels. The clearest analogy with pearlite is found when the alloy carbide in lath morphology forms nodules in association with ferrite. These pearlitic nodules are often encountered at temperatures just below Ac1, in steels which transform relatively slowly. For example, these structures are obtained in chromium steels with between 4% and 12% chromium and the crystallography is analogous to that of cementitic pearlite. It is, however, different in detail because of the different crystal structures of the possible carbides. The structures observed are relatively coarse, but finer than pearlite formed under equivalent conditions, because of the need for the partition of the alloying element, e.g. chromium between the carbide and the ferrite. To achieve this, the interlamellar spacing must be substantially finer than in the equivalent iron-carbon case. Interphase precipitation. Interphase precipitation has been shown to nucleate periodically at the / interface during the transformation. The precipitate particles form in bands which are closely parallel to the interface, and which follow the general direction of the interface even when it changes direction sharply. A further characteristic is the frequent development of only one of the possible Widmansttten variants, for example VC plates in a particular region are all only of one variant of the habit, i.e. that in which the plates are most nearly parallel to the interface.

Appendix D. Page 75 of 116.

Menu

CM MOK cmmok128@yahoo.com
The extremely fine scale of this phenomenon in vanadium steels, which also occurs in Ti and Nb steels, is due to the rapid rate at which the / transformation takes place. At the higher transformation temperatures, the slower rate of reaction leads to coarser structures. Similarly, if the reaction is slowed down by addition of further alloying elements, e.g. Ni and Mn, the precipitate dispersion coarsens. The scale of the dispersion also varies from steel to steel, being coarsest in chromium, tungsten and molybdenum steels where the reaction is relatively slow, and much finer in steels in which vanadium, niobium and titanium are the dominant alloying elements and the transformation is rapid. Transformation diagrams for alloy steels. The transformation of austenite below the eutectoid temperature can best be presented in an isothermal transformation diagram, in which the beginning and end of transformation is plotted as a function of temperature and time. Such curves are known as time-temperature-transformation, or TTT curves, and form one of the important sources of quantitative information for the heat treatment of steels. In the simple case of a eutectoid plain carbon steel, the curve is roughly C-shaped with the pearlite reaction occurring down to the nose of the curve and a little beyond. At lower temperatures bainite and martensite are formed. The diagrams become more complex for hypo- and hyper-eutectoid alloys as the ferrite or cementite reactions have also to be represented by additional lines

Appendix D. Page 76 of 116.

Menu

CM MOK cmmok128@yahoo.com

Alloying and Its Effects on the Critical Temperature, Hardness and Tensile Strength

Alloying elements have significant effect on the iron-iron carbide equilibrium diagram. The addition of some of these alloying elements will widen the temperature range through which austenite (gamma -iron) is stable while other elements will constrict the temperature range. What this means is that some elements will raise and some elements will lower the critical temperature of steel. Manganese, cobalt, and nickel increase the temperature range through which austenite is stable. This also means that the lower critical temperature of steel will be lowered by these alloying elements. Other alloying elements that lower the critical temperature of steel are carbon, copper and zinc. The alloying elements that are used to reduce the critical temperature are highly soluble in the gamma iron (austenite). Figure 1 shows the effect of manganese on the critical temperature of steel.

Figure 1. The effect of alloying with manganese on the critical temperature of steel and austenite (-iron) phase transformation zone on the iron-iron carbide diagram.. Alloys such as aluminum, chromium, molybdenum, phosphorus, silicon, tungsten tend to form solid solutions with alpha iron (ferrite). This constricts the temperature region through which gamma iron (austenite) is stable. As shown in Figure 2, chromium at different percentages constricts the critical temperature range which results in a marked reduction of the region where austenite is stable.

Appendix D. Page 77 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 2. Effect of alloying with chromium on the critical temperature of steel and austenite (g -iron) phase transformation zone on the iron-iron carbide diagram. The elements shown in Figure 3 have the greatest solubility in ferrite and also influence the hardenability of iron when in the presence of carbon. With a slight increase in the carbon content, they respond markedly to heat treating, because carbon acts as a ferrite strengthener. As indicated in Figure 3, Phosphorus will improve the hardness of the ferrite significantly by adding only a very small percentage of Phosphorus, while Chromium will not strengthen the ferrite that well even at very high percentage of Chromium addition to the steel

Appendix D. Page 78 of 116.

Menu

CM MOK cmmok128@yahoo.com
Figure 3. The effect of various alloying elements on the hardness of steel. Figure 4 shows the effect of furnace cooling vs. air cooling on the tensile strength of steel for three different percentages of carbon in the presence of chromium. As this figure indicates, furnace cooling has very little effect on the tensile strength of the material. The addition of chromium does not change the tensile strength properties when the steel is cooled in the furnace. If the same steels are air cooled at the same rate, the slope of the curves increases significantly which means that a slight increase in the chromium content increases the strength drastically when air cooling is applied.

Figure 4. Effect of different percentages of carbon on the tensile strength of steel in the presence of chromium.

Appendix D. Page 79 of 116.

Menu

CM MOK cmmok128@yahoo.com

Control of High Strength Low Alloy (HSLA) Steel Properties

Most HSLA steels are furnished in the as-hot-rolled condition with ferritic-pearlitic microstructure. The exceptions are the controlled-rolled steels with an acicular ferrite microstructure and the dual-phase steels with martensite dispersed in a matrix of polygonal ferrite. These two types of HSLA steels use the formation of eutectoid structures for strengthening, while the ferritic-pearlitic HSLA steels generally require strengthening of the ferrite. Pearlite is generally an undesirable strengthening agent in structural steels because it reduces impact toughness and requires higher carbon contents. Moreover, yield strength is largely unaffected by a higher pearlite content. Strengthening Mechanisms in Ferrite The ferrite in HSLA steels is typically strengthened by grain refinement, precipitation hardening, and, to a lesser extent, solid-solution strengthening. Grain refinement is the most desirable strengthening mechanism because it improves not only strength but also toughness. Grain refinement is influenced by the complex effects of alloy design and processing methods. For example, the various methods of grain refinement used in the three different stages of hot rolling (that is, reheating, hot rolling, and cooling) include: The addition of titanium or aluminum to retard austenite grain growth when the steel is reheated for hot deformation or subsequent heat treatment The controlled rolling of microalloyed steels to condition the austenite so that it transforms into fine-grain ferrite The use of alloy additions and/or faster cooling rates to lower the austenite-to-ferrite transformation temperature. The use of higher cooling rates for grain refinement may require consideration of its effect on precipitation strengthening and the possibility of undesirable transformation products. Precipitation strengthening occurs from the formation of finely dispersed carbonitrides developed during heating and cooling. Because precipitation strengthening is generally associated with a reduction in toughness, grain refinement is often used in conjunction with precipitation strengthening to improve toughness. Precipitation strengthening is influenced by the type of carbonitride, its grain size, and, of course, the number of carbonitrides precipitated. The formation of MC is the most effective metal carbide in the precipitation strengthening of microalloyed niobium, vanadium, and/or titanium steels. The number of fine MC particles formed during heating and cooling depends on the solubility of the carbides in austenite and on cooling rates. Steelmaking Precise steelmaking operations are also essential in controlling the properties and chemistry of HSLA steels. Optimum property levels depend on such factors as the control of significant alloying elements and the reduction of impurities and nonmetallic inclusions. Developments in secondary steelmaking such as desulphurization, vacuum degassing, and argon shrouding have enabled better control of steel chemistry and the effective use of microalloyed elements. Compositional limits for HSLA steel grades described in ASTM specifications the use of vacuum degassing equipment allows the production of interstitial-free (IF) steels. The IF steels exhibit excellent formability, high elongation, and good deep draw/ability. Compositions and Alloying Elements Chemical compositions for the HSLA steels are specified by ASTM standards. The principal function of alloying elements in these ferrite-pearlite HSLA steels, other than corrosion resistance, is strengthening of the ferrite by

Appendix D. Page 80 of 116.

Menu

CM MOK cmmok128@yahoo.com
grain refinement, precipitation strengthening, and solid-solution strengthening. Solid-solution strengthening is closely related to alloy contents, while grain refinement and precipitation strengthening depend on the complex effects of alloy design and thermo-mechanical treatment. Alloying elements are also selected to influence transformation temperatures so that the transformation of austenite to ferrite and pearlite occurs at a lower temperature during air cooling. This lowering of the transformation temperature produces a finer-grain transformation product, which is a major source of strengthening. At the low carbon levels typical of HSLA steels, elements such as silicon, copper, nickel, and phosphorus are particularly effective for producing fine pearlite. Element such as, manganese and chromium, which are present in both the cementite and ferrite, also strengthen the ferrite by solid-solution strengthening in proportion to the amount, dissolved in the ferrite. In the presence of alloying elements, the practical maximum carbon content at which HSLA steels can be used in the as-cooled condition is approximately 0.20%. Higher levels of carbon tend to form martensite or bainite in the microstructure of as-rolled steels, although some of the higher-strength low-alloy steels have carbon contents that approach 0.30%. The required strength is developed by the combined effect of: Fine grain size developed during controlled hot roiling and enhanced by microalloyed elements (especially niobium) Precipitation strengthening caused by the presence of vanadium, niobium, and titanium in the composition. Nitrogen additions to high-strength steels containing vanadium are limited to 0.005% and have become commercially important because such additions enhance precipitation hardening. The precipitation of vanadium nitride in vanadium-nitrogen steels also improves grain refinement because it has a lower solubility in austenite than vanadium carbide. Manganese is the principal strengthening element in plain carbon high-strength structural steels. It functions mainly as a mild solid-solution strengthener in ferrite, but it also provides a marked decrease in the austenite-to-ferrite transformation temperature. In addition, manganese can enhance the precipitation strengthening of vanadium steels and. to a lesser extent, niobium steels. One of the most important applications of silicon is its use as a deoxidizer in molten steel. Silicon has a strengthening effect in low-alloy structural steels. In larger amounts, it increases resistance to scaling at elevated temperatures. Silicon has a significant effect on yield strength enhancement by solid-solution strengthening and is widely used in HSLA steels for riveted or bolted structures. Copper in levels in excess of 0.50% also increases the strength of both low- and medium-carbon steels by virtue of ferrite strengthening, which is accompanied by only slight decreases in ductility. Copper can be retained in solid solution even at the slow rate of cooling obtained when large sections are normalized, but it is precipitated out when the steel is reheated to about 510 to 605C (950 to 1125F). At about 1% copper, the yield strength is increased by about 70 to 140 MPa regardless of the effects of other alloying elements. Copper in amounts up to 0.75% is considered to have only minor adverse effects on notch toughness or weldability. Copper precipitation hardening gives the steel the ability to be formed extensively and then precipitation hardened as a complex shape or welded assembly. The atmospheric-corrosion resistance of steel is increased appreciably by the addition of phosphorus, and when small amounts of copper are present in the steel, the effect of the phosphorus is greatly enhanced. When both phosphorus and copper are present, there is a greater beneficial effect on corrosion resistance than the sum of the effects of the individual elements. Chromium is often, added with copper to obtain improved atmospheric-corrosion resistance.

Appendix D. Page 81 of 116.

Menu

CM MOK cmmok128@yahoo.com
Nickel is often added to copper-bearing steels to minimize hot shortness. Molybdenum in hot-rolled HSLA steels is used primarily to improve hardenability when transformation products other than ferrite-pearlite are desired. Molybdenum (0.15 to 0.30%) in microalloyed steels also increases the solubility of niobium in austenite, thereby enhancing the precipitation of NbC(N) in the ferrite. This increases the precipitation-strengthening effect of NbC(N). Aluminum is widely used as a deoxidizer and was the first element used to control austenite grain growth during reheating. During controlled rolling, niobium and titanium are more effective grain refiners than aluminum. Vanadium strengthens HSLA steels by both precipitation hardening the ferrite and refining the ferrite grain size. The precipitation of vanadium carbonitride in ferrite can develop a significant increase in strength that depends not only on the rolling process used, but also on the base composition. Carbon contents above 0.13 to 0.15% and manganese content of 1% or more enhances the precipitation hardening, particularly when the nitrogen content is at least 0.01%. Titanium is unique among common alloying elements in that it provides both precipitation strengthening and sulfide shape control. Small amounts of titanium (<0.025%) are also useful in limiting austenite grain growth. However, it is useful only in fully killed (aluminum deoxidized) steels because of its strong deoxidizing effects, the versatility of titanium is limited because variations in oxygen, nitrogen, and sulfur affect the contribution of titanium as carbide strengthened. Zirconium can also be added to killed high-strength low-alloy steels to improve inclusion characteristics, particularly in the case of sulfide inclusions, for which changes in inclusion shape improve ductility in transverse bending. Boron has no effect on the strength of normal hot-rolled steel but can considerably improve hardenability when transformation products such as acicular ferrite are desired in low-carbon hot-rolled plate. Treatment with calcium is preferred for sulfide inclusion shape control. Controlled Rolling The hot-rolling process has gradually become a much more closely controlled operation, and controlled rolling is now being increasingly applied to microalloyed steels with compositions carefully chosen to provide optimum mechanical properties at room temperature. Controlled rolling is a procedure whereby the various stages of rolling are temperature controlled, with the amount of reduction in each pass predetermined and the finishing temperature precisely defined. This processing is widely used to obtain reliable mechanical properties in steels for pipelines, bridges, offshore platforms, and many other engineering applications. The use of controlled rolling has resulted in improved combinations of strength and toughness and further reductions in the carbon content of microalloyed HSLA steels

Appendix D. Page 82 of 116.

Menu

CM MOK cmmok128@yahoo.com

Influence of Alloying Elements on Steel Microstructure


It is a long-standing tradition to discuss the various alloying elements in terms of the properties they confer on steel. For example, the rule was that Chromium (Cr) makes steel hard whereas Nickel (Ni) and Manganese (Mn) make it tough. In saying this, one had certain types of steel in mind and transferred the properties of particular steel to the alloying element that was thought to have the greatest influence on the steel under consideration. This method of reasoning can give false impressions and the following examples will illustrate this point. When we say that Cr makes steel hard and wear-resisting we probably associate this with the 2% C, 12% Cr tool steel grade, which on hardening does in fact become very hard and hard-wearing. But if, on the other hand, we choose a steel containing 0.10% C and 12% Cr, the hardness obtained on hardening is very modest. It is quite true that Mn increases steel toughness if we have in mind the 13% manganese steel, so-called Hadfield steel. In concentrations between l% and 5%, however, Mn can produce a variable effect on the properties of the steel it is alloyed with. The toughness may either increase or decrease. A property of great importance is the ability of alloying elements to promote the formation of a certain phase or to stabilize it. These elements are grouped as austenite-forming, ferrite-forming, carbide-forming and nitride-forming elements. Austenite-forming elements The elements C, Ni and Mn are the most important ones in this group. Sufficiently large amounts of Ni or Mn render a steel austenitic even at room temperature. An example of this is the so-called Hadfield steel which contains 13% Mn, 1,2% Cr and l% C. In this steel both the Mn and C take part in stabilizing the austenite. Another example is austenitic stainless steel containing 18% Cr and 8% Ni. The equilibrium diagram for iron-nickel, Figure 1, shows how the range of stability of austenite increases with increasing Ni-content.

Appendix D. Page 83 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 1. Fe-Ni equilibrium diagram More diagrams: http://www.calphad.com/phase_diagrams.html

An alloy containing 10% Ni becomes wholly austenitic if heated to 700C. On cooling, transformation from g to a takes place in the temperature range 700-300C. Ferrite-forming elements The most important elements in this group are Cr, Si, Mo, W and Al. The range of stability of ferrite in iron-chromium alloys is shown in Figure 2. Fe-Cr alloys in the solid state containing more than 13% Cr are ferritic at all temperatures up to incipient melting. Another instance of ferritic steel is one that is used as transformer sheet material. This is a low-carbon steel containing about 3% Si.

Appendix D. Page 84 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 2. Cr-Fe equilibrium diagram

Multi-alloyed steels The great majority of steels contain at least three components. The constitution of such steels can be deduced from ternary phase diagrams (3 components). The interpretation of these diagrams is relatively difficult and they are of limited value to people dealing with practical heat treatment since they represent equilibrium conditions only. Furthermore, since most alloys contain more than three components it is necessary to look for other ways of assessing the effect produced by the alloying elements on the structural transformations occurring during heat treatment. One approach that is quite good is the use of Schaeffler diagrams (see Figure 3). Here the austenite formers are set out along the ordinate and the ferrite formers along the abscissa. The original diagram contained only Ni and Cr but the modified diagram includes other elements and gives them coefficients that reduce them to the equivalents of Ni or Cr respectively. The diagram holds good for the rates of cooling which result from welding.

Appendix D. Page 85 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 3. Modified Schaeffler diagram

Appendix D. Page 86 of 116.

Menu

CM MOK cmmok128@yahoo.com
A 12% Cr steel containing 0.3% C is martensitic, the 0.3% C gives the steel a nickel equivalent of 9. An 18/8 steel (18% Cr, 8% Ni) is austenitic if it contains 0-0.5% C and 2% Mn. The Ni content of such steels is usually kept between 9% and 10%. Hadfield steel with 13% Mn (mentioned above) is austenitic due to its high carbon content. Should this be reduced to about 0.20% the steel becomes martensitic. Carbide-forming elements Several ferrite formers also function as carbide formers. The majority of carbide formers are also ferrite formers with respect to Fe. The affinity of the elements in the line below for carbon increases from left to right. Cr, W, Mo, V, Ti, Nb, Ta, Zr. Some carbides may be referred to as special carbides, i.e. non-iron-containing carbides, such as Cr7C3 W2C, VC, Mo2C. Double or complex carbides contain both Fe and a carbide-forming element, for example Fe4W2C. High-speed and hot-work tool steels normally contain three types of carbides, which are usually designated M6C, M23C6 and MC. The letter M represents collectively all the metal atoms. Thus M6C represents Fe4W2C or Fe4Mo2C; M23C6 represents Cr23C6 and MC represents VC or V4C3. Carbide stabilizers The stability of the carbides is dependent on the presence of other elements in the steel. How stable the carbides are depends on how the element is partitioned between the cementite and the matrix. The ratio of the percentage, by weight, of the element contained in each of the two phases is called the partition coefficient K. The following values are given for K: Al 0 Cu 0 P 0 Si 0 Co 0.2 Ni 0.3 W 2 Mo 8 Mn 11,4 Cr 28 Ti Nb Ta

Increasing

Note that Mn, which by itself is a very weak carbide former, is a relatively potent carbide stabilizer. In practice, Cr is the alloying element most commonly used as a carbide stabilizer. Malleable cast iron (i.e. white cast iron that is rendered soft by a graphitizing heat treatment called malleablizing) must not contain any Cr. Steel containing only Si or Ni is susceptible to graphitization, but this is most simply prevented by alloying with Cr. Nitride-forming elements All carbide formers are also nitride formers. Nitrogen may be introduced into the surface of the steel by nitriding. By measuring the hardness of various nitrided alloy steels it is possible to investigate the tendency of the different alloying elements to form hard nitrides or to increase the hardness of the steel by a mechanism known as precipitation hardening. The results obtained by such investigations are shown in Figure 4, from which it can be seen that very high hardnesses result from alloying a steel with Al or Ti in amounts of about 1,5%.

Appendix D. Page 87 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 4. Effect of alloying element additions on hardness after nitriding Base composition: 0.25% C, 0.30% Si, 0.70% Mn On nitriding the base material in Figure 4, hardness of about 400 HV is obtained and according to the diagram the hardness is unchanged if the steel is alloyed with Ni since this element is not a nitride former and hence does not contribute to any hardness increase

Appendix D. Page 88 of 116.

Menu

CM MOK cmmok128@yahoo.com

Martensite in Austenitic Stainless Steel Welds

Background
Martensite is a crystal structure that forms in steels during rapid cooling. Cooling rates are dependent on the particular chemistry of the steel. Certain conditions can be met that will cause martensite formation in austenitic stainless steels. Long, needle-like clusters of crystals in the metal characterize martensite. The martensite crystals have a highly stressed body centered tetragonal structure. In austenitic stainless steels and the austenite phase of a magnetic (austenitic/ferritic) stainless, when martensite is present in a weld, don't expect it to pass the bend test. The weld will not have the toughness and ductility that we normally expect from a stainless steel, and in some circumstances there may be unanticipated corrosion.

Fillet Composition
With only one exception, adding alloying elements to steel allows martensite to form at a slower cooling rate. That exception is cobalt, which works the other way. With cobalt, martensite will form at a faster cooling rate. We recognize that the actual composition of the weld fillet will vary along a line running across or bisecting the weld fillet. On the centerline, we expect the fillet composition to be closer to the wire composition. Approaching each base metal, the composition shifts from the filler metal composition towards each base metal. In welding metallurgy, we mark on the phase diagram the chemistry of one base metal and the chemistry of the filler wire. We find that the actual fillet compositions lie on the line drawn on the phase diagram. When the base metals are different, it takes different lines on the phase diagram to represent the compositions approaching each base metal.

Predicting Martensite
The results and microstructural consequences of this sort of exercise in physical metallurgy appeared in 1949 as the "Schaeffler Diagram". Our technology did not stand still and the Diagram iterated through several updates. The 1994 Winter Addendum to the ASME Code brought us the Welding Research Council's "WRC-1992 Diagram" which continues to be extensively used. Still, there has been a problem associated with manganese, which brought yet another modification into use. The analysis behind the modification appears, for example, in D.J. Kotecki's, "A martensite boundary on the WRC-1992 diagram" (Welding Journal , Vol. 78, No. 5, pp 180-192). In lots of cases the 1% manganese line satisfactorily predicts martensite or not martensite. However, we sometimes encounter steels having higher manganese and quite often encounter much lower. High side

Appendix D. Page 89 of 116.

Menu

CM MOK cmmok128@yahoo.com
examples include 1.0-1.5% manganese in a 309L filler wire, 4% in a 307, and 6% in the European 18 8 Mn filler wire. On the low side, if you are joining stainless to a modern carbon steel, the manganese can be quite low, perhaps even 0.3%.

(The 1949 Schaeffler diagram - click image to enlarge) To enter the modified Diagram, we need to calculate two numbers roughly based on chromium and nickel content. We also have to apply a lot of welding "know-how" as to the mixing of the metals. Calculate the

Nickel Ni Chrome Cr

Equivalent = %Ni + 35%C + 20%Mn + 0.25%Cu Equivalent = %Cr + Mo + 1.5%Si + 0.7%Nb

The Amount of Nitrogen


One of the problems we have to also confront is the amount of nitrogen. The best shielding practice with a wire electrode might not introduce nitrogen. Flux cored electrodes tend to add nitrogen. For example, the metal in a flux cored electrode might analyze to 0.05% nitrogen while it deposits as 0.075% nitrogen. Most of us deal with the nitrogen by first plotting the points with only the known nitrogen and then we plot the nearby point based on

Appendix D. Page 90 of 116.

Menu

CM MOK cmmok128@yahoo.com
our estimate of the actual nitrogen. The compositions at risk are those on the Diagram and below the indicated manganese brands. Within the bands the diagram is known to be imprecise. In practice, you have to cope with the range of compositions between one base metal and the filler metal, and between the other base metal and the filler metal. You could put all of your predicted nickel and chromium equivalent compositions on the graph. However, examining the diagram we see that the compositions more at risk for forming martensite are those with low nickel and chromium equivalent numbers, coupled with low manganese. After a little practice with the diagram we know pretty well which compositions are likely to be at risk and we tend to plot only those numbers. A warning: Notice that all of the alloying elements in the calculation contribute to hardenability. Should either base metal or the filler metal contain any unlisted elements which are known to contribute to hardenability then the Diagram doesn't apply. After decades of absence, tungsten is showing up mostly as a substitute for molybdenum and mostly in Russian and Chinese metals. In doing the calculation the practice is to lump the tungsten with the moly on a 1:1 basis. When forced to do it, welding metallurgists may tweak the calculations a little to reflect other unlisted elements, however, there isn't much of a research foundation on which to base such adjustments.

Ferrite Number
The 300 series of stainless steels is austenitic (non-magnetic) while the duplex stainless steels are mixed austenite and ferrite. In the field, the amount of ferrite is measured through its magnetic response. The portable meter is calibrated for % ferrite and it is called the "ferrite number". Martensite gives a magnetic response, but not as strongly as ferrite, so when it is present it contributes to the "ferrite number". Recognize also that perfectly good austenitic stainless steel that has been heavily cold worked can become slightly magnetic and give a ferrite number even though ferrite nor martensite are present. In the upper right side of the diagram there is a cluster of lines of constant ferrite number. The upper left line is 1% ferrite number which is nominally 99% austenite and 1% ferrite. The last line along the lower right side of the cluster is nominally 98% ferrite and 2% austenite.

Learning by Doing (and "cya")


Good record keeping means photocopying a bunch of diagrams and every time you confront the martensite issue you make a record of how you calculated the nickel and chromium equivalents. You plot these on the graph, draw the lines between and estimate an adjustment for nitrogen. You note the ferrite number and whether the bend test is pass or fail. You don't just do the exercise for samples that fail the bend test. Otherwise you won't learn the limits for welds that pass every time! If testing the weld for its ferrite number is worthwhile then making the record is also worthwhile. The day will come when you note the manganese is low and you make yourself into a hero when you predict the need for a little preheat.

Appendix D. Page 91 of 116.

Menu

CM MOK cmmok128@yahoo.com

(The WRC-1992 Diagram modified to reflect experience with manganese. Martensite is predicted to form with composition equivalents below and to the left of the manganese bands. Reprinted from Advanced Materials&Process, June 1000, p. 75 - click image to enlarge) There will be discrepancies. When it happens it most likely means that you misjudged the effect of cooling rate, or you didn't get the chemistry right, or the basic diagram is imperfect!

Appendix D. Page 92 of 116.

Menu

CM MOK cmmok128@yahoo.com

Effects on the martensite, pearlite and bainite formation

Effect on the temperature of martensite formation

All alloying elements with the possible exception of Co, lower Ms the temperature of the start of the martensite formation, as well as Mf, the finish of the martensite formation, i.e. at 100% martensite. For the majority of steels containing more than 0.50% C, Mf lies below room temperature. This implies that after hardening these steels practically always contain some residual austenite. Ms may be calculated from the equation given below, by inserting the percentage concentration of each alloying element in the appropriate term. The equation is valid only if all the alloying elements are completely dissolved in the austenite. Ms = 561 - 474C - 33Mn - 17Ni - 17Cr - 21Mo For high-alloy and medium-alloy steels Stuhlmann has suggested the following equation: Ms(C) = 550 - 350C - 40Mn - 20Cr - 10Mo - 17Ni - 8W - 35V - 10Cu + 15Co + 30Al It can be noted that carbon has the strongest influence on the Ms temperature. Figures 1 and 2 show diagrams with an example of experimental results of the effect of Mn and Ni on the Ms temperature of various types of steel.

Figure 1. Effect of Mn on the Ms - temperature (after Russel and McGuire, Payson and Savage, Zyuzin, Grange and Stewart)

Appendix D. Page 93 of 116.

Menu

CM MOK cmmok128@yahoo.com

Figure 2. Effect of Ni on the Ms - temperature (after Russel and McGuire, Payson and Savage, Zyuzin, Grange and Stewart)

Effect on the formation of pearlite and bainite during the isothermal transformation All alloying elements except Co delay the formation of ferrite and cementite. It is very difficult to formulate any general rules regarding the influence exerted by the various alloying elements. However, it has definitely been found that some elements affect the bainite transformation more than the pearlite transformation, while other elements act in the opposite manner. Certain elements will, paradoxically, accelerate the transformations if their concentration increases beyond a certain limiting value, this limit been affected by other alloying elements present. For case-hardening and tool steels the time taken to initiate the pearlite-bainite transformation is reduced as the carbon content exceeds about 1%. For tool steels and constructional steels Si-concentrations of 1,5% and above have been found to promote pearlite formation. As a general principle it may be stated that by increasing the concentration of one alloying element by some few percent and the basic carbon content being kept about 0.50%, only a relatively small retardation of the transformation rates is noticed. For plain carbon steels a successive increase in C from 0.30% to 1% produces but a negligible effect. It is only in conjunction with several alloying elements that a more noticeable effect is produced. The diagram in Figure 3, applicable to steel W 1 (l% C) will serve as a basis for this discussion. The shortest transformation time for this steel is less than 1/8th second. Note that the time scale is logarithmic; hence there is no zero time. As has been mentioned previously, both pearlite and bainite form simultaneously in this steel at about 550C. Since the curves overlap it is customary to draw only one curve. With increasing contents of certain alloying elements, however, the noses of the pearlite and bainite curves will separate. The structures shown in Figure 3 are obtained by austenitizing samples of steel W 1 at 780C for 10 min and quenching in a salt bath at various temperatures. After holding them for predetermined times at various temperatures they are finally quenched in water. Before the salt-bath quenching the steel contains undissolved

Appendix D. Page 94 of 116.

Menu

CM MOK cmmok128@yahoo.com
carbides but in view of the composition of the austenite the steel may be regarded as an eutectoid one. The diagram should be studied with the aid of the explanatory text below.

Figure 3. TTT diagram for isothermal transformation of steel W 1 (1% C) A = austenite, B = bainite, Ms = start of martensite transformation, M50 = 50% M, P = pearlite

1. Quenching in a liquid bath at 700C; holding time 4 min. During this interval the C has separated out, partly as pearlite lamellae and partly as spheroidized cementite. Hardness 225 HV. 2. Quenching to 575C, holding time 4 s. A very fine, closely spaced pearlite as well as some bainite has formed. Note that the amount of spheroidized cementite is much less than in the preceding case. Hardness 380 HV. 3. Quenching to 450C, holding time 60 s. The structure consists mainly of bainite. Hardness 410 HV.

4. Quenching to 20C (room temperature). The matrix consists of, roughly, 93% martensite and 7% retained austenite. There is some 5% cementite as well which has not been included in the matrix figure. Hardness 850 HV.

Appendix D. Page 95 of 116.

Menu

CM MOK cmmok128@yahoo.com

Steel Alloys

Below is a list of some SAE-AISI designations for Steel (the xx in the last two digits indicate the carbon content in hundredths of a percent)
Carbon Steels 10xx 11xx 12xx Manganese steels 13xx Nickel steels 23xx 25xx Nickel Chromium Steels 31xx 32xx 33xx 34xx Chromium Molybdenum steels 41xx Nickel 43xx 47xx 86xx Nickel steels 46xx 48xx Chromium steels 50xx 51xx Cr 0.27- 0.65 Cr 0.80 1.05 Illustration of effect of Carbon content on Steel Hardness Ni 0.85-1.82 Mo 0.20 Ni 3.50 Mo 0.25 Molybdenum Chromium Ni 1.82 Cr 0.50-0.80 Mo 0.25 Ni 1.05 Cr 0.45 Mo 0.20 0.35 Ni 0.55 Cr 0.50 Mo 0.20 Molybdenum steels Cr 0.50-0.95 Mo 0.12-0.30 Ni 1.25 Cr 0.65-0.80 Ni 1.75 Cr 1.07 Ni 3.50 Cr 1.50-1.57 Ni 3.00 Cr 0.77 Ni 3.5 Ni 5.0 Mn 1.75 Plain Carbon Resulfurized Resulfurized and rephosphorized

Appendix D. Page 96 of 116.

Menu

CM MOK cmmok128@yahoo.com

Carbon Steel to Austenitic Steel


When a weld is made using a filler wire or consumable, there is a mixture in the weld consisting of approximately 20% parent metal and 80% filler metal alloy ( percentage depends on welding process, type of joint and welding parameters). Any reduction in alloy content of 304 / 316 type austenitic is likely to cause the formation of martensite on cooling. This could lead to cracking problems and poor ductility. To avoid this problem an over alloyed filler metal is used, such as a 309, which should still form austenite on cooling providing dilution is not excessive. The Shaeffler diagram can be used to determine the type of microstructure that can be expected when a filler metal and parent metal of differing compositions are mixed together in a weld. The Shaeffler Diagram

The Nickel and other elements that form Austenite, are plotted against Chrome and other elements that form ferrite, using the following formula:Nickel Equivalent = %Ni + 30%C + 0.5%Mn Chrome Equivalent = %Cr + Mo + 1.5%Si + 0.5%Nb Example, a typical 304L = 18.2%Cr, 10.1%Ni, 1.2%Mn, 0.4%Si, 0.02%C Ni Equiv = 10.1 + 30 x 0.02 + 0.5 x 1.2 = 11.3 Cr Equiv = 18.2 + 0 + 1.5 x 0.4 + 0 = 18.8

Appendix D. Page 97 of 116.

Menu

CM MOK cmmok128@yahoo.com
A typical 309L welding consumable Ni Equiv = 14.35, Cr Equiv = 24.9 The main disadvantage with this diagram is that it does not represent Nitrogen, which is a very strong Austenite former. Ferrite Number The ferrite number uses magnetic attraction as a means of measuring the proportion of delta ferrite present. The ferrite number is plotted on a modified Shaeffler diagram, the Delong Diagram. The Chrome and Nickel equivalent is the same as that used for the Shaeffler diagram, except that the Nickel equivalent includes the addition of 30 times the Nitrogen content.

Examples

Appendix D. Page 98 of 116.

Menu

CM MOK cmmok128@yahoo.com
The Shaeffler diagram above illustrates a carbon steel C.S , welded with 304L filler. Point A represents the anticipated composition of the weld metal, if it consists of a mixture of filler metal and 25% parent metal. This diluted weld, according to the diagram, will contain martensite. This problem can be overcome if a higher alloyed filler is used, such as a 309L, which has a higher nickel and chrome equivalent that will tend to pull point A into the austenite region. If the welds molten pool spans two different metals the process becomes more complicated. First plot both parent metals on the shaeffler diagram and connect them with a line. If both parent metals are diluted by the same amount, plot a false point B on the diagram midway between them. (Point B represents the microstructure of the weld if no filler metal was applied.)

Next, plot the consumable on the diagram, which for this example is a 309L. Draw a line from this point to false point B and mark a point A along its length equivalent to the total weld dilution. This point will give the approximate microstructure of the weld metal. The diagram below illustrates 25% total weld dilution at point A, which predicts a good microstructure of Austenite with a little ferrite.

Appendix D. Page 99 of 116.

Menu

CM MOK cmmok128@yahoo.com

The presence of martensite can be detected by subjecting a macro section to a hardness survey, high hardness levels indicate martensite. Alternatively the weld can be subjected to a bend test ( a side bend is required by the ASME code for corrosion resistant overlays), any martensite present will tend to cause the test piece to break rather than bend. However the presence of martensite is unlikely to cause hydrogen cracking, as any hydrogen evolved during the welding process will be absorbed by the austenitic filler metal. More reading: http://nhml.com/search.cfm?zoom_query=stainless+steel&zoom_per_page=10&zoom_and=0&zoom_sort=0

Appendix D. Page 100 of 116.

Menu

CM MOK cmmok128@yahoo.com

Appendix D. Page 101 of 116.

Menu

CM MOK cmmok128@yahoo.com

Selection of Age-Hardenable Superalloys

By Richard B. Frank High-Temperature R&D Carpenter Technology Corp., Reading, PA, USA Superalloys are high-performance materials designed to provide high mechanical strength and resistance to surface degradation at high temperatures of 1200F (650C) or above. They combine high tensile, creep-rupture, and fatigue strength; good ductility and toughness, with excellent resistance to oxidation and hot corrosion. Furthermore, superalloys are designed to retain these properties during long-term exposures at the elevated temperatures. This article focuses on the wrought age-hardenable alloys, which are the most commonly used superalloys. Wrought materials can be formed using hot and cold working operations. Not discussed here are the cast, powder (P/M), and oxide dispersion strengthened (ODS) superalloys that can also offer enhanced properties. The first age-hardenable, high-temperature alloy dates back to about 1929 when various developers added titanium and aluminum to the standard 80% nickel/20% chromium resistance wire alloy. This was a precursor to the 80A nickel-base superalloy, developed in 1940-1944, but still in use today. Little was done to advance the original age-hardenable alloys until the time period of 1935-1944 when World War II spurred demand for improved alloys that could be used in the early aircraft gas turbine engines. Alloy development activity exploded in the 1950s and 1960s to keep pace with the demands of the gas turbine engine industry. Progress in superalloy development not only made the jet engine possible, but allowed for constantly increasing thrust-to-weight ratios over the last 60 years. Applications The primary application for superalloys is still in hot sections of aircraft gas turbine engines, accounting for over 50% of the weight of advanced engines. However, the excellent performance of these materials at elevated temperatures has expanded their application far beyond one industry. In addition to the aerospace industry, these alloys are used in turbine engines for marine, industrial, land-based power generation, and vehicular applications. Specific engine parts using superalloys include turbine discs, blades, compressor wheels, shafts, combustor cans, afterburner parts and engine bolts. Beyond the gas turbine engine industries, superalloys are commonly used for applications in rocket engines, space, petrochemical/energy production, internal combustion engines, metal forming (hot-working tools and dies), heat-treating equipment, nuclear power reactors, and coal conversion.

Appendix D. Page 102 of 116.

Menu

CM MOK cmmok128@yahoo.com
While these alloys are primarily used for service at elevated temperatures above 1000F (540C), the characteristics of high strength and excellent environmental resistance have made some superalloys an excellent choice for lower-temperature applications. Examples are prosthetic devices in the medical industry and components for deep sour gas wells in the oil/gas exploration industry. Chemical Composition Table 1 contains the nominal compositions of the most common wrought age-hardenable superalloys. These alloys contain various combinations of nickel, iron, cobalt, and chromium with lesser amounts of other elements including molybdenum, niobium, titanium, and aluminum. With minor additions of beneficial elements such as boron and zirconium, these alloys may contain up to 12 intentional additions. All these additions help to impart and maintain the desired properties at elevated temperatures.

Many other elements such as silicon, phosphorus, sulfur, oxygen, nitrogen and a larger number of tramp elements (like lead, bismuth, selenium) must be tightly controlled in superalloys to avoid detrimental effects on high-temperature properties. These minor and tramp elements are controlled during raw material selection prior to melting, as well as during the melting/remelting processes.

Appendix D. Page 103 of 116.

Menu

CM MOK cmmok128@yahoo.com
Superalloys can be classified into nickel-base, iron-base, and cobalt-based groups. Nickel-based superalloys (>50% Ni) are the most common group. About half of the alloys in Table 1 are considered nickel-base alloys and the others contain large additions of nickel. The nickel base has a high tolerance for alloy additions that might otherwise cause phase instability leading to loss of strength, ductility, and/or environmental resistance. Iron-based superalloys are less costly, but are less tolerant of alloying additions and typically have lower mechanical properties and maximum temperature limitations. Examples are Pyromet Alloy A-286 and NCF 3015 (Ni-30) alloy. These alloys contain an austenitic stainless steel base with additions of nickel, titanium, and aluminum to promote age hardening. Pyromet Alloy 706 and Pyromet Alloy 901 have similar amounts of nickel and iron and can be considered nickel-iron-base superalloys. The higher nickel levels of 901 and 706 alloys allow for larger additions of strengthening elements without undesirable effects. Although there are some cobalt-base superalloys, they are significantly higher in cost and typically cannot be age hardened to high strength levels. However, cobalt is an important alloying addition to nickel-based alloys because it extends the maximum temperature for usage by reducing the solubility of the age-hardening phase. Waspaloy and Pyromet Alloy 41 and Pyromet Alloy 720 are nickel-base alloys with 10-15% cobalt additions. These alloys have the highest temperature capability of the common wrought age-hardenable superalloys. Chromium, usually in the range of 14 to 23 weight percent, is a critical alloying addition to nearly all superalloys. As in stainless steels, chromium forms a tightly-adherent, protective oxide film (Cr2O3) on the alloy surface to resist oxidation and corrosion at high temperatures as well as corrosion at lower temperatures. This surface layer protects the alloy from the harmful effects of the elements oxygen, nitrogen, and sulfur. Although most superalloys contain at least 14% chromium, in some applications, it is critical to minimize thermal expansion. Pyromet CTX-909 and Thermo-Span alloys are considered low-expansion superalloys that have low chromium contents to minimize expansion of the nickel-cobalt-iron base. Resistance to oxidation and hot corrosion are reduced so high-temperature coatings are often applied prior to service. Of the two alloys, 909 alloy provides the lowest expansion coefficient while Thermo-Span alloy (5.5% chromium) provides improved environmental resistance. Refractory elements like molybdenum, tungsten, and niobium, with their large atomic diameters, increase high temperature strength and stiffness by straining the nickel/iron base matrix. Alloys 901 and 41 contain larger additions of molybdenum to increase this solid solution strengthening effect. Other alloying additions such as chromium and aluminum also contribute to solid solution strengthening but to a lesser extent. The elements titanium, aluminum, and niobium are added to the nickel or nickel-iron matrix to form an intermetallic Ni3 (Al, Ti, Nb) phase during age-hardening heat treatments. The resultant gamma prime or gamma double prime phases are the primary strengthening agents in superalloys. This will be discussed in more detail in the next section on age-hardening. Although elements such as boron, zirconium, and magnesium may be added at levels less than 0.1 weight percent, the beneficial effects can be very potent. These elements segregate to and stabilize grain boundaries, which significantly improves hot workability, high temperature strength and ductility. Small additions of carbon

Appendix D. Page 104 of 116.

Menu

CM MOK cmmok128@yahoo.com
also may be added to form carbides that restrict grain growth and grain boundary sliding during high temperature exposure. Age-Hardening The major strengthening method in superalloys is age-hardening. Yield strength of nickel alloys is typically increased by a factor of two or three by precipitation of the gamma prime and/or gamma double prime, Ni3 (Al, Ti, Nb) hardening phase. Although the phase is based on the nickel aluminide (Ni3Al) intermetallic, up to 60% of the aluminum can be replaced by titanium or niobium, which actually increases strength of the alloy. The gamma prime phase is rather unique in that its strength actually increases with temperature up to 1200F (650C) and it is relatively ductile and resistant to oxidation. Gamma prime precipitates as very fine spheroidal or cuboidal particles in the nickel-iron matrix during aging. While most of the superalloys employ the titanium-rich gamma prime phase for age hardening, a niobium-rich variant called gamma double prime is the primary strengthening phase in some superalloys such as Pyromet Alloy 706 and Pyromet Alloy 718. The niobium-rich phase provides higher strength up to 1200F (650C) but is unstable above 1200F. Thus, 706 and 718 alloys have a lower temperature limit than the alloys strengthened with the titanium-rich gamma prime phase. Since the gamma double prime reaction is more sluggish, these alloys also tend to have better hot workability and weldability. Heat Treatment Proper heat treatment is critical to achieving the desired level of properties in age-hardenable superalloys. Typical heat treatments for these alloys are listed in the mechanical property Tables 2 and 3. The initial solution heat treatment typically dissolves all precipitated phases except for some primary carbide and nitride phases. The typical range for the wrought age-hardenable superalloys is 1650-2100F (900-1150C) for 1 to 4 hours followed by a rapid air cool or a quench in water, polymer or oil.

Appendix D. Page 105 of 116.

Menu

CM MOK cmmok128@yahoo.com

Appendix D. Page 106 of 116.

Menu

CM MOK cmmok128@yahoo.com

The selection of solution treatment time and temperature varies with the alloy and its phase solvus temperatures, and also depends on the specific properties that are most important for the intended application. Alloys with higher hardener contents (Ti, Al, Nb) require higher temperatures to solution any hardener phase that may have precipitated during hot working or cooling. Best tensile and fatigue properties are typically obtained with lower solution temperatures that result in a finer grain size. In contrast, better long-term stress-rupture and creep properties are generally obtained with higher-temperature solution treatments that result in coarser grain size and lower tensile yield strength. For these reasons, it is common to specify two or more preferred heat treatments for superalloys. In some cases, another objective of the solution treatment is to form a desirable distribution of a second phase such as carbide in Pyromet 41 alloy and delta phase (Ni3Nb) in Pyromet 718 alloy. After solution treatment, one or more aging treatments are applied to precipitate the hardening phase and possibly other phases in the

Appendix D. Page 107 of 116.

Menu

CM MOK cmmok128@yahoo.com
desired amount and distribution. As with solution treatment, the selection of aging temperatures is dependent on the alloy and the combination of properties desired. The aging range for age-hardenable superalloys is 1150-1600F (620-870C). Aging times range from 4 hours to 24 hours. Double-aging treatments are quite common to maximize strength and to develop the best combination of short-term tensile and long-term creep-rupture properties. The primary aging treatment precipitates a coarser distribution of the hardener phase and may also improve the type and distribution of carbides on grain boundaries. The secondary age is typically about 200F below the primary aging temperature, precipitating a finer dispersion of the gamma prime phase. For some higher-strength applications, the alloy is direct aged after hot, warm, or cold working without an intermediate solution treatment. The strain from working is used to further enhance tensile and fatigue properties with some sacrifice in creep-rupture properties. Mechanical Properties For the design engineer or materials specifier, a review of terms defining applicable mechanical properties may be helpful: Tensile Properties The design of load-bearing structures is often based on yield strength or, in some cases, the ultimate tensile strength of the material. Yield strength is a measure of the maximum stress a material can withstand before it permanently deforms. Tensile strength is a measure of the maximum stress a material can withstand before it fractures. Elevated temperature tensile properties are most applicable to short-time exposures at higher temperatures. Creep and stress-rupture properties are more applicable for longer exposures. Creep and Rupture Properties Creep and rupture strengths become important when the material must withstand the combined effects of high temperature and stress for long periods of time. At elevated temperatures, metals will stretch or "creep" at stresses well below the yield strength. Superalloys are more resistant to creep than low-alloy or stainless steels, but creep will still occur above about 1000F (540C). Creep properties are a measure of the alloys resistance to stretching under a constant load. Stress-rupture or creep-rupture properties are a measure of resistance to fracture under a constant load (creep test taken to fracture). Both properties are expressed as stress or strength values that will cause a given amount of creep (0.1%-1%) or rupture in a given amount of time (100 to 100,000 hours). Tables 2 and 3 list typical tensile (yield) and stress-rupture strength properties of the age-hardenable superalloys at temperatures of 1200-1600F (650-870C). Yield strengths at room temperature are also listed in Table 2. These properties are shown graphically in Figures 1 and 2. It should be noted that the data represents approximate nominal strength values for specific heat treatments. Actual values can vary by up to 35% due to differences in composition, hot/cold working practices, and heat treatment.

Appendix D. Page 108 of 116.

Menu

CM MOK cmmok128@yahoo.com

Appendix D. Page 109 of 116.

Menu

CM MOK cmmok128@yahoo.com

For example, superalloys like Pyromet 718 and Waspaloy may contain several different aim compositions within the broader industry ranges to optimize properties for specific applications. Higher levels of the age-hardening elements titanium, aluminum, and niobium result in higher strength. Hot or cold working an alloy to obtain a finer grain size typically increases tensile yield strength but decreases stress-rupture strength. As discussed previously, properties of all age-hardenable superalloys are dependent on heat treatment. Alloys like Pyromet Alloy X-750 and Waspaloy have two or more preferred heat treatments (see Table 2) depending on whether the application requires better short-time tensile and fatigue properties or long-time creep and stress-rupture properties. Examples of alternative heat treatments have been shown for Waspaloy and X-750 alloys but the reader should refer to manufacturers datasheets for a more complete listing of alternative heat treatments for the other superalloys. Other Properties While tensile and creep-rupture are the most basic mechanical properties considered for high-temperature applications, design criteria may also consider resistance to fatigue (low- and high-cycle), crack growth, and wear/erosion. Hardness and hot hardness tests are sometimes used as a rough measure of yield strength and wear/erosion. Alloy Selection

Appendix D. Page 110 of 116.

Menu

CM MOK cmmok128@yahoo.com
A simplified method known as the Carpenter Selectaloy system can help designers and engineers select the most suitable superalloy based on strength and maximum temperature requirements. Figures 3 and 4 contain Selectaloy diagrams for the 15 superalloys discussed in this article. Yield strength (Figure 3) or stress-rupture strength (Figure 4) increases vertically on the Selectaloy diagram, and temperature increases from left to right. The alloys are shown multiple times on the diagrams since the alloys are useful over a range of temperatures. The diagram can be used to estimate how the strength of an alloy decreases with temperature, but also how the strength of different alloys compare at different temperatures.

Appendix D. Page 111 of 116.

Menu

CM MOK cmmok128@yahoo.com

It should be noted that the alloys were positioned on the Selectaloy diagrams based on average strength values representative of compositions and heat treatments commonly used for each alloy. An alloys relative position could move up or down, left or right, with relatively minor modifications of composition, processing and heat treatment. Temperature limits should be considered approximate. Therefore, while the Selectaloy diagrams are useful tools to screen candidate alloys, they are not a substitute for a more detailed evaluation of the critical properties required for an intended application.

Pyromet A-286 alloy is the most basic age-hardenable superalloy in terms of properties and cost. A-286 provides the lowest strength levels, but still higher by a factor of two than other non-age-hardenable stainless alloys. When increased strength or temperature resistance is required, higher nickel alloys are typically preferred. Alloys with the highest levels of strength and temperature resistance typically contain the highest alloy contents and significant levels of cobalt. Relative cost of these alloys will be discussed in the next section. The Selectaloy diagrams presented in this article provide a method to compare basic strength properties and temperature limitations of common wrought age-hardenable superalloys. However, alloy selection will undoubtedly depend on many other considerations, including other physical and mechanical properties as well as environmental resistance and cost. For example, Thermo-Span and Pyromet CTX-909 alloys provide a benefit of much lower expansion during heating but at the expense of oxidation and corrosion resistance in the

Appendix D. Page 112 of 116.

Menu

CM MOK cmmok128@yahoo.com
uncoated condition. Pyromet 31V and Pyromet 751 alloys provide similar strength and temperature resistance, but the higher chromium content of 31V alloy results in much improved resistance to sulfidation and other forms of hot corrosion. Alloy Cost From the users standpoint, alloy selection must be based on expected cost effectiveness. In todays competitive global environment, overdesign is less common. The trend is to select the lowest-cost material to meet design requirements for the application. However, a higher-cost alloy may be justified to minimize overall life cycle cost or for longer service of certain components in a system that is critical or too expensive to be shut down for maintenance. Surely, knowledge of alloy capabilities is critical in making the best decision. As temperature and strength requirements increase, so does the necessary alloy content. Figure 5 compares the relative alloying costs of the 14 alloys using Pyromet A-286 alloy as a base (cost factor of 1.0). The cost factors are based on 10-year averages of the intrinsic alloying element costs at market prices. Higher temperature strength and resistance typically require higher nickel and cobalt contents. Nickel and cobalt prices have historically been volatile, with high and low prices varying by a factor of 4 to 5. More recently, the price of molybdenum, a potent solid solution strengthener, has increased in price by a factor of nearly ten over the last two years.

Appendix D. Page 113 of 116.

Menu

CM MOK cmmok128@yahoo.com

As discussed above, the cost factors in Figure 5 are based only on raw material elemental costs (10-year averages) that fluctuate significantly with time. Differences in melting, working, and other processing costs, which can be substantial, are not included in these factors. Processing yields and specific end user requirements (grain size, ultrasonic testing, etc.) significantly impact product cost. However, the cost comparisons are useful because alloying costs typically represent a large portion of superalloy product cost. Since superalloys are designed for high temperature strength and resistance to deformation, processing difficulty and cost also increase with hot strength and maximum temperature capability. Figure 6 shows the relationships of stress-rupture strength at 1200-1500F (650-815C) with raw material cost factor. It is apparent that the alloys that provide higher levels of strength, temperature resistance, and/or specialized properties also cost more, which reinforces the importance of the alloy selection process.

Appendix D. Page 114 of 116.

Menu

CM MOK cmmok128@yahoo.com

More Technical Article: http://crswnew.cartech.com/wnew/techarticles/TechLibrarySelector.html

Appendix D. Page 115 of 116.

Menu

Microplasticity: 1 dislocations
MSoM / MEM 2nd year. 6 lectures. Steve Roberts The basics revisited. Burgers vector and Burgers circuit. Line sense and FS/RH convention. Geometry of dislocations: edge, screw and mixed. Conservation of Burgers vector at nodes. Elastic properties of dislocations. Stress fields of screw and edge dislocations, strain energy of a dislocation. The force on a dislocation. Forces between dislocations. Image forces. Small-angle grain boundaries and epitaxial interfaces. Slip. Dislocation sources and multiplication. Climb and jogs. Strength of crystalline solids. The Peierls stress and lattice resistance. Kinks and thermally activated glide. Temperature and strain-rate dependence of flow stress. Plastic strain due to dislocation movement. Dislocations in c.c.p. metals. Perfect dislocations, Shockley partial dislocations and intrinsic stacking faults. The Thompson tetrahedron. Lomer-Cottrell locks. Cross-slip. Frank partial dislocations. Dislocations in h.c.p. metals. Basal and non-basal slip. Dislocations in b.c.c. metals. Absence of stable stacking faults. Non-planar screw dislocation cores. Dislocations in ordered intermetallics. Superdislocations, antiphase boundaries and the yield stress anomaly. Dislocations in non-metals. Ionic and covalent crystals. Followed by Microplasticity: 2 strength of materials
Steve Roberts Microplasticity S.G. Roberts 1: 1

Books and other Resources


Introduction to Dislocations, D. Hull and D.J. Bacon (3rd edn.). Theory of Dislocations, J.P. Hirth and H. Lothe. Dislocations, J. Freidel. Physical Metallurgy, R.W. Cahn and P. Haasen. Crystallography and Crystal Defects, A. Kelly and G.W. Groves. Worked Examples in Dislocations, M.J. Whelan. Physical Metallurgy Principles, R.E. Reed-Hill (3rd edn.). Microplasticity: Lecture notes by A.P. Sutton, Materials Dept. Library.

Steve Roberts Microplasticity

S.G. Roberts

1: 2

Menu

Recap Burgers vector


Perform an atom-to-atom circuit around the dislocation line, returning to the start point.

Convention: choice of line direction is arbitrary, but having chosen it, the circuit is done in the right-hand sense.

Repeat the atom-to-atom circuit in a perfect crystal. Circuit does not close. Closure vector is the Burgers vector b. Convention: b goes from Finish to Start of the circuit. Convention: FSRH b is defined w.r.t. lattice of perfect crystal

Steve Roberts Microplasticity

S.G. Roberts

1: 3

Edge dislocations where is the half plane ? Silicon


Structures shown so far have been grossly simplified, as simple cubic lattices. In materials with more realistic structures, geometry of even simple dislocations may be more subtle. Silicon has a face centred cubic lattice, with two atoms per lattice point. In the High Resolution Transmission Electron Micrograph shown, each bright spot represents one lattice point. There is an edge dislocation present. However, these lines are only one way of indicating lattice planes crossing the dislocation core. The extra half plane concept isnt quite as simple as in the simple cubic case. Nonetheless, the half plane is in both cases on the same side of the slip plane.
Steve Roberts Microplasticity S.G. Roberts 1: 4

Menu

Screw dislocations more detailed geometry


l
n+ n+

l b

n+ n+0

b b
The standard picture above makes it appear that the dislocation is confined to one plane (unless it reconstructs in some way). Real situation for a pure screw dislocation is rather more like that to the left. Displacement parallel to b accumulates uniformly as one progresses round the core of the dislocation.

b
n+ n+

n+0

n+

In real materials this helical structure gets further complicated

Steve Roberts Microplasticity

S.G. Roberts

1: 5

Mixed dislocations
The Burgers vector b is the principal defining characteristic of a dislocation. When a dislocation moves, every atom it crosses on the slip plane moves by b. This happens whatever the line vector. In many real materials, especially those with soft bonding, there is no preferred crystal direction for the dislocation line. Most dislocations will then not be pure screw or pure edge, but somewhere in between mixed dislocations. For stress analyses, we can treat mixed dislocations as equivalent to a vector sum of pure edge and pure screw components, which we can treat independently:
bs be

bs be

l bs = b cos be = b sin

b l b
be bs

M S E

The reason we can do this is because the stress fields of the edge and screw components are orthogonal they have no components in common.

Steve Roberts Microplasticity

S.G. Roberts

1: 6

Menu

Dislocation nodes
b1 b3 b1 1 b2 2 2 Since the only places the crystal structure is interrupted are at the dislocation cores, displacements in the Burgers circuits 2 and 3 must add to give the displacement associated with either of the two variants of Burgers circuit 1. b2 b1 b2 b3 3 1 b1 b3 3

b1 = b2 + b3

b1 = b2 + b3
(N.B. care is needed with the senses of the line vectors.)

Steve Roberts Microplasticity

S.G. Roberts

1: 7

Dislocation loops
One type of dislocation loop forms the boundary of a completely enclosed patch of slipped material on the slip plane.

View down direction 1

b l b b
1 b must be the same at all points on the loop.

b l b b
2 View down direction 2

Shear stresses acting on this loop will tend either to expand or to contract it.

Steve Roberts Microplasticity

S.G. Roberts

1: 8

Menu

Electron microscopy of dislocations


Electron beam wavelength Diffraction contrast in the T.E.M. is now the commonest way of seeing dislocations and how they move and interact. An electron beam is shone onto a crystal so that a Bragg condition is satisfied. = 2d sin A strong diffracted beam results, and illuminates the imaging screen. Near the core of the dislocation, the lattice planes are bent. The Bragg condition is no longer satisfied. The diffracted beam is very weak or absent. So in this bright field image the region near the dislocation core appears dark. Visiblity is due to the dislocations strain field.

Steve Roberts Microplasticity

S.G. Roberts

1: 9

Some TEM pictures of dislocations


Dislocation network in graphite
(Amelinckx 1960)

10 m 1 m Dislocation tangles in strained stainless steel.


(Whelan, 1958)

Moving dislocations in molybdenum


(Robertson 1999)

Steve Roberts Microplasticity

S.G. Roberts

1: 10

Menu

Strains and Stresses around Dislocations


Dislocations have a long-range stress field that can be analysed using linear elasticity diffuse strain energy stored in a large volume no variation with core position relative to atomic structure a core within which the strains are too great to be treated using linear elasticity intense strain energy stored in a small volume may be large energy fluctuations with core position The elastic field controls how dislocations react to distant microstructural features with their own elastic stress fields other dislocations mis-fitting precipitates mis-fitting solute atoms twins applied stresses The core structure controls how the dislocation interacts with the crystals lattice & atomic structure: dislocation dissociation core spreading mobility cross-slip defects on the core (kinks and jogs) details of interaction with point defects

Steve Roberts Microplasticity

S.G. Roberts

1: 11

Elasticity - recap
Displacement vector:

u = [ux, uy, uz]


Strain tensor:

Stress tensor:

xx = 2G xx + yy = 2G yy + zz = 2G zz + xy = 2G xy xz = 2G xz yz = 2G yz = 2G (1 2 )

xx = xy
Dilatation, :

u x x u y 1 u x =2 y + x

etc.

= xx+ yy+ zz
Elastic energy:

1 dE el = 2 dV

i= x,y,z j= x,y,z

ij ij

y x
Steve Roberts Microplasticity S.G. Roberts 1: 12

Menu

Strain field of straight screw dislocation


z

xx = yy = zz = xy = yx = 0 xz = b uz y tan 1 = 4 x x x b 1 y = 4 1 + y 2 x 2
1 2

( x)

b y b sin = 4 x 2 + y 2 4 r
u z b y = tan 1 y 4 y x

Recipe : - take a hollow cylinder, axis along z: - cut on a plane parallel to the z-axis; -displace the free surfaces by b in the z-direction. By inspection: u = u = 0 x y

yz = =

1 2

b 1 4 1 + y

( x)

1 x

uz =

b 2 b y = tan 1 2 x
S.G. Roberts

b x b cos = = 4 x 2 + y 2 4 r

Steve Roberts Microplasticity

1: 13

Stress field of straight screw dislocation


xx = yy = zz = xy = yx = 0 xz b y b sin = = 4 x 2 + y 2 4 r b x b cos = 4 x 2 + y 2 4 r
In Polar coordinates: (either by direct inspection, or by transforming the strains and stresses from Cartesian co-ordinates)

yz =

z = z = z = z

b 4r Gb = 2r

All other components of the stress tensor are zero.

xx = yy = zz = xy = yx = 0 =0 xz = 2G xz = yz = 2G yz = Gb y Gb sin = 2 x 2 + y 2 2 r

Gb x Gb cos = 2 x 2 + y 2 2 r

Note: Stress and strain fields are pure shear Fields have radial symmetry Stresses and strains are proportional to 1/r: extend to infinity tend to infinite values as r0 Infinite stresses cannot exist in real materials: the dislocation core radius r0 is that within which our assumption of linear elastic behaviour breaks down. Typically r0 1 nm.

Steve Roberts Microplasticity

S.G. Roberts

1: 14

Menu

Stress field of straight edge dislocation


z

xz = zx = yz = zy = 0 xx = D y yy = D y
y

(x
2

3x 2 + y 2
2

+ y2

with : D =

Gb 2(1 )

(x

x2 y2 + y2

x Recipe : - take a hollow cylinder, axis along z: - cut on a plane parallel to the z-axis; - displace the free surfaces LMNO by b in the x-direction. Situation is plane-strain. No displacements in z-direction. Derivation of stress tensor is complicated. (see Hirth and Lothe for the full works)
Steve Roberts Microplasticity

xy = yx = D x

(x

x2 y2
2

+ y2

zz = ( xx + yy ) Hydrostatic Stress, P :
1 P=3 ( xx + yy + zz ) =

2(1 + ) y D 2 3 x + y2

Note: Stress and strain fields are not pure shear Stresses and strains are proportional to 1/r: extend to infinity tend to infinite values as r0
S.G. Roberts 1: 15

Stress field around edge dislocation


y

+
400-500 300-400 200-300 100-200 0-100 -100-0 -200--100 -300--200 -400--300 -500--400

Shear stress xy around an edge dislocation with b = bx (arbitrary units)

Hydrostatic stress

(xx + yy + zz) /3

around an edge dislocation with b = bx (arbitrary units)

Steve Roberts Microplasticity

S.G. Roberts

1: 16

Menu

Strain energy of a screw dislocation

e c or
r

Elastic energy per unit volume = 2 1 (zz + zz ) = Gb 2 2 8 r Volume of shell, thickness r = 2r.r Elastic energy of shell =

In the shell shown,

Gb2 r 4 r

z = z =

b 4 r

R
Total elastic energy

Gb z = z = 2 r
All other stresses and strains are zero.

r0 =

4r dr

Gb2

Gb 2 R ln 4 r0
per unit length of dislocation line

Steve Roberts Microplasticity

S.G. Roberts

1: 17

Strain energy of an edge dislocation


Now imagine making the dislocation by cutting on green plane and displacing the two sides by relative b b

e c or
y

Gb 2(1 )x

xy x 0 with dislocation in place,

displacement

Work done at x

xy = yx at y = 0 : xy =

Gb x2 y2 = x 2(1 ) x 2 + y 2 2

1 Gb b 2 2(1 )x Total work done between r0 and R = R = = r0

Gb 2(1 )x

4(1 )x dx
This must equal the elastic energy per unit length of the dislocation.
1: 18

Gb 2

R Gb 2 ln 4(1 ) r0

Steve Roberts Microplasticity

S.G. Roberts

Menu

Core energy & total energy of dislocations


We can estimate the core energy as equivalent to about one broken bond per atom spacing along the core. This will be about (vacancy formation energy)/(co-ordn. no.) : typically (1 - 3 eV) / (4 12) per 0.2 0.3 nm So range is likely to be 0.5 5 eV nm-1 . = 0.1 1 nJ m-1 This energy is for the relaxed state. If we try to move the dislocation, the core energy will fluctate as the dislocation goes from one relaxed state to the next one. This is the origin of the Peierls Nabarro stress the minimum stress to move the dislocation line. (see later) To estimate the elastic energy

Eel =

Gb 2 R ln 4( ) r0

We need values for R and r0: reasonable values are R ~ 1 m, r0 ~ 1nm:

R 1 ln 0. 5 4( ) r0 Eel Gb 2 2

Typically G = 60 120 GPa, b = 0.2 0.3 nm: Eel will be in range 1 4 nJ m-1. So total energy of dislocation is likely to be (just) dominated by the diffuse elastic energy but energy fluctuations will depend on the core term.

Steve Roberts Microplasticity

S.G. Roberts

1: 19

Force on an (edge) dislocation


If there is a high enough shear stress component in the direction of b, then the dislocation can glide. If there is a dilatational stress parallel to b, then if temperature is high enough for vacancies to diffuse to or from the dislocation, it will climb. In both cases the applied stress does work in changing the shape of the crystal. This is achieved by the motion of the dislocation, which responds as if the stress is applying a force to it. This configurational force is defined by the rate of change of energy of the system as the dislocation moves. conservative motion
Steve Roberts Microplasticity S.G. Roberts

Edge: glide

Edge: climb

non-conservative motion
1: 20

Menu

Force on a dislocation
If section of dislocation line moves by s, then within area sl, all atoms above the slip plane are displaced by b w.r.t. all those below.

Work done, W = force.displacement = stress.displacement x area = .b sl l

l s

Force on dislocation (the Peach-Koehler force) is defined as work done per unit length of dislocation line when it glides unit distance. Force, F = work done / line length / glide distance = W / (s l) = .b

Dislocation line can only meaningfully move normal to the line vector l. F must be perpendicular to l

If the full stress tensor is used, then:

F = (b) l

Fx xx Fy = yx F z zx
S.G. Roberts

xy yy zy

xz b x l x yz b y l y zz bz l z
1: 21

Steve Roberts Microplasticity

Forces on dislocations
Dislocation motion only has meaning normal to the line vector. Forces on dislocations can only act normal to the line vector.

Screw

Mixed

Loop

F
b b
Steve Roberts Microplasticity S.G. Roberts 1: 22

Menu

Forces between dislocations


Using arguments based on dislocation energetics, it is easy to see that, for dislocations of same line vector on the same slip plane, dislocations with opposite b attract each other, dislocations with identical b repel.

Eel 2

Gb 2 2
b

Eel 2

Gb 2 2

-b r

b The elastic energies of the separated dislocations given apply strictly only as r . As r decreases, energy changes towards the overlap values. (see question sheet 1).

2b

Eel = 0
F= dEel <0 dr

Eel

G( 2b)2 = 2Gb 2 2
dE el >0 dr

This is one view of the reason for the forces between dislocations. For more complicated situations, it is easier to see the force as being due to the local stress at each dislocation.
S.G. Roberts 1: 23

F=

Steve Roberts Microplasticity

Forces between screw dislocations


y x z y b r b Alternatively, we can use the stress field expressed in Cartesian co-ordinates: 2 1 Force on dislocation 2 from dislocation 1, resolved onto the glide plane is:

Fres =

Gb 2 cos 2r

x Dislocation 2 feels the stress field of dislocation 1 (and vice versa).

xx = yy = zz = xy = yx = 0 xz = yz = Gb Gb sin y = 2 x 2 + y 2 2 r

z = z =

Gb 2r

Gb Gb cos x = 2 x 2 + y 2 2 r

So force on dislocation 2 from dislocation 1 is:

F=

Gb 2 2r

Note that the shear stress acting to shear atoms parallel to b above and below the glide plane is yz.

but this force acts in the radial direction.


Steve Roberts Microplasticity

Fres = yzb =
S.G. Roberts

Gb 2 Gb 2 x cos = 2r 2 x 2 + y 2
1: 24

Menu

Forces between edge dislocations


y x z b 1 x r b 2

xx = D y

(x

3x 2 + y 2
2

+ y

2 2

with : D =

Gb 2(1 )

xy = yx = D x

(x

x 2 y 2
2

+ y 2

So glide force, resolved onto the slip plane, is:

Dislocation 2 feels the stress field of dislocation 1 (and vice versa). The important components of the stress field are:

Fglide =
0.3 Fglide
Gb 2 2(1 )y

Gb2 x( x 2 y 2 ) 2(1 ) x 2 + y 2 2

xy produces glide force on disln 2; xx produces climb force on disln 2.

4 x / y

-0.3
Steve Roberts Microplasticity S.G. Roberts 1: 25

Stable arrangements for edge dislocations


0.3 Fglide
Gb 2 2(1 )y

For like Burgers vectors: x = y: unstable equilibrium x = 0 : stable equilibrium 2 4 x / y 6 8 For opposite Burgers vectors: x = y: stable equilibrium x = 0 : unstable equilibrium

-0.3 For a set of opposite Burgers vectors: There are a large number of possible stable arrangements. For like Burgers vectors: Stable array is a planar stack A low angle tilt boundary. This arrangement has a strong long-range stress field. Taylor lattice Dipole dispersion

These stable arrangements have minimal longrange stress fields.


S.G. Roberts 1: 26

Steve Roberts Microplasticity

Menu

Dislocations near free surfaces: image forces

y x

xx around edge dislocation If dislocation is near free surface, it produces stresses in the surface plane that the surface cannot support. Stress field must be modified by presence of free surface.

The trick is to introduce a virtual dislocation (of opposite b) which is the mirror image in the free surface of the real one. The combined stress fields of the real and image dislocations reproduce the solution for the stress field of the dislocation near the free surface. The real dislocation sees the stress field of the image dislocation and is attracted to the free surface.

Steve Roberts Microplasticity

S.G. Roberts

1: 27

Dislocation line tension


The energy per unit length of the dislocation line is equivalent to a line tension: Units: J m-1 = (N m) m-1 = N Strength Gb2/2 A straight dislocation line will exert an inward force equal to the (line energy / length) on any pinning points. A curved dislocation line will have a tendency to straighten itself between any two points, as this minimises line length and therefore energy. F Gb2/2 .b = Gb2 / 2R b l F Gb2/2
Steve Roberts Microplasticity S.G. Roberts 1: 28

On small dislocation line segment, length l Line tension exerts force: 2 (Gb2/2 sin(/2) ) = Gb2/2 sin = Gb2/2 Applied shear stress exerts force .b l = .b R These must balance: Gb2/2 = .b R

R /2 /2 R

or if we only take component of in direction of b: = Gb / 2R Stress required to move a dislocation line increases as its radius of curvature decreases.

Menu

Dislocation sources
The Frank-Read source is the most commonly illustrated type of dislocation source. F F F F b Apply a shear stress which has a component in the direction of the dislocations Burgers vector b. Dislocation line will be subjected to the Peach-Koehler force F. Imagine a segment of dislocation line pinned at two ends probably at large kinks (see later).

If F is greater than the minimum level needed to make the dislocation glide plus the effects of line tension, the dislocation will glide, in the direction of F.

Steve Roberts Microplasticity

S.G. Roberts

1: 29

Dislocation sources
Increasing is needed as the dislocation bows out, because the line curvature increases. At this stage, when the loop shape is a half-circle, dislocation curvature is a maximum. Now no increase in stress is needed to keep the dislocation moving the loop is unstable. b b l l The dislocation line is now expanding behind its original position. These segments are of opposite sign. They attract one another.

The two rear segments approach each other rapidly

Steve Roberts Microplasticity

S.G. Roberts

1: 30

Menu

Dislocation sources
Increasing is needed as the dislocation bows out, because the line curvature increases. At this stage, when the loop shape is a half-circle, dislocation curvature is a maximum. Now no increase in stress is needed to keep the dislocation moving the loop is unstable. The dislocation line is now expanding behind its original position. b These segments are of opposite sign. They attract one another.

The two rear segments approach each other rapidly and mutually annihilate.

Steve Roberts Microplasticity

S.G. Roberts

1: 31

Stress to operate dislocation sources


Increasing is needed as the dislocation bows out, because the line curvature increases. At this stage, when the loop shape is a half-circle, dislocation curvature is a maximum. .b = Gb2 / 2R 2R or if we only take component of in direction of b: = Gb / 2R

Steve Roberts Microplasticity

S.G. Roberts

1: 32

Menu

Real dislocation sources

The classic Frank-Read source. Silicon. (C. Dash, 1957 )

Half sources where a single pinning point is near a grain boundary, can act as dislocation mills. Molybdenum (bcc). (I. Robertson, 1999)

Note the alignment of dislocation segments along <110> directions. Anisotropy of Peierls stress controls loop shape. The dislocation lines leaving the slip planes at the pinning points are also visible.
Steve Roberts Microplasticity

Note the very different mobilities of the screw segments (slow) and edge segments (very fast) of the same loop. Loop shape is controlled by this mobility difference.
1: 33

S.G. Roberts

Widening of slip band by multiple cross-slip

Screw dislocation can have high enough resolved shear stress for glide on more than one slip plane. Cross-slip can occur. But this leaves some segments of dislocation on the original (primary) slip plane. Dislocation can cross-slip back on to a parallel primary slip plane, where it forms a new dislocation source. and the process can repeat.

Steve Roberts Microplasticity

S.G. Roberts

1: 34

Menu

Widening of slip band by multiple cross-slip


Glide band widening in LiF

original dislocations

Glide band width after first loading Glide band width after second loading

Steve Roberts Microplasticity

S.G. Roberts

1: 35

Strain rate from motion of dislocations


d w

If a dislocation moved the whole length of the crystal d, it would contribute b to the displacement D. If each dislocation moves an amount xi (less than d), then each will contribute ( xi / d ).b to D.

D=
Shear strain is:

b xi d i = D b = xi h dh i x= 1 xi N i

b Define average distance moved by each dislocation: Density of mobile dislocations is:

m =
&=

xi Strain rate:

N hd

d dx = b m = bm v dt dt

- where v is the average dislocation glide velocity.


Steve Roberts Microplasticity S.G. Roberts 1: 36

Menu

Dislocation motion The Peierls-Nabarro Stress


Elasticity theory tells us a lot about dislocations stresses, energetics & interactions, but to understand dislocation mobility we need to look at the core structure and how it changes as the dislocation moves. 1. Can we define the width of a dislocation core and what does it depend on? What happens as the dislocation moves on the glide plane? Why do dislocations choose particular crystallographic planes within a crystal on which to glide? In principle slip could take place on any plane. Why are the Burgers vectors of perfect dislocations usually the shortest available crystal lattice vectors? In principle they could be any crystal lattice vector. (E = Gb2/2 is one factor)
S.G. Roberts

w?

2. 3.

N.B. the P-N approach only applies to dislocations with planar cores. (e.g. edge dislocations, dissociated screw dislocations) Structure of pure screw dislocation cores (e.g. in bcc metals) needs direct atomistic simulation.
1: 37

4.

Steve Roberts Microplasticity

Core width bubble raft

Steve Roberts Microplasticity

S.G. Roberts

1: 38

Menu

Dislocation core width

Steve Roberts Microplasticity

S.G. Roberts

1: 39

Definition of core width


y b uB A a B uA u / b x Consider displacements of lattice with dislocation present w.r.t. lattice with dislocation absent. Above and below the slip plane these displacements u will be different at the same distance x from the centre line of the dislocation. The disregistry u is the difference between these displacements.
+0.5 +0.25 -0.25 -0.5

u = uB - uA Width w of core can now be defined as the width of the region within which u is greater than half its maximum value (0.5b)

Steve Roberts Microplasticity

S.G. Roberts

1: 40

Menu

Core types

wide

narrow

dissociated

0.5 u / b -0.5 f Area = b x

0.5 u / b -0.5 f Area = b

0.5 u / b -0.5 f Area = b/2

f(x) =

d u - the distrubution of Burgers vector along the slip plane. dx


S.G. Roberts 1: 41

Steve Roberts Microplasticity

Peierls energy and Peierls stress


b y Find that energy fluctuates, with period b/2: uB a uA Assume that 1) The atoms in plane A interact with the atoms in plane B via a simple sinusoidally-varying force law. 2) Disregistry forces are balanced by elastic stresses from material above and below the slip plane . gives an analytical solution for w: A B x

E = E0 + Ep =

Ep 2

4 x sin b
Peierls Energy

Gb 2 2w exp (1 ) b

Maximum force needed is maximum slope of energy / displacement curve:

Fmax =

2Gb 2w exp (1 ) b

w=

a (edge ) (1 ) = a (screw)

Stress to move dislocation = Fmax / b:

p =

then calculate the dislocation energy as sum of disregistry energy and elastic energy, as function of core position x w.r.t. lattice.
Steve Roberts Microplasticity S.G. Roberts

2G 2w exp (1 ) b

Peierls Stress

1: 42

Menu

Peierls valleys
Peierls Stress Consequences: Dislocations with wide, planar cores are easier to move than ones with narrow (or even nonplanar) cores.
edge dislocations will have lower p than screw dislocations (esp. in bcc see later)

p =

2G 2w exp (1 ) b

The Peierls energy varies with dislocation orientation w.r.t. the lattice. Dislocations tend to lie along directions where Ep is lowest Peierls valleys
(esp. in strongly-bonded materials)

[110]

[101] Dislocations with widely-spaced slip planes will have wider cores
hcp metals with high c/a will have low p

[011] Assumption of Peierls model is that dislocations move wholesale from one Peierls valley to the next. Also no thermal activation. p = stress to move at absolute zero.

Dislocations with small b will have low p. These effects are strong, as they are via an exponential relationship.

Steve Roberts Microplasticity

S.G. Roberts

1: 43

Peierls valleys, kinks and dislocation glide


At real temperatures, dislocation lines have a population of kinks in thermal equilibrium. Dislocation lines can move under stress by nucleation and sideways propagation of double kinks

Ep
Steve Roberts Microplasticity

Kink formation is opposed by line tension and the mutual attraction of kinks of opposite sign. Thermal activation is needed.
1: 44

S.G. Roberts

Menu

Crystallography of kinks
Kink also have structure they will have lowest energy if aligned along secondary Peierls valleys. (1) However, this gives maximum line length maximum line energy. Minimum line energy configuration is to have dislcoation line and kinks ignoring the primary and secondary Peierls valleys. (2) Real dislocation structure will be a compromise, giving minimum total energy (3). Balance will depend on bonding strength and temperature. Stress to produce glide at given rate will then depend on: 1. 2. Stress to produce double kinks; Stress to move kinks

These will depend on the details of the kink structure (material & temperature dependent) and the amount of thermal activation available (temperature dependent).
Theories have been worked out in detail (see Hirth and Lothe), but are beyond the scope of this course. We will look at simplified theories which ignore the kink mechanics. S.G. Roberts 1: 45

Steve Roberts Microplasticity

Kinks from dislocation intersections

The intersection of these two dislocations dislocations leads to the formation of a kink on one dislocation. The kink is effectively an atomic-length section of edge dislocation: - on the same crystal glide plane as the rest of the dislocation; - acts as a secondary dislocation - sideways motion of the kink can act as mechanism for glide of the main dislocation
Steve Roberts Microplasticity S.G. Roberts 1: 46

Menu

Jogs from dislocation intersections

The intersection of these two screw dislocations leads to the formation of a jog on each one. The jog is effectively an atomic-length section of edge dislocation: - probably on a non-glide crystal plane; - can only move with dislocation to which it is attached by climb.
Steve Roberts Microplasticity S.G. Roberts 1: 47

Jogs from dislocation intersections

The intersection of these two dislocations dislocations leads to the formation of a jog on one dislocation. The jog is effectively an atomic-length section of edge dislocation: - in this case on a crystal glide plane; - can move with dislocation to which it is attached by glide.

Steve Roberts Microplasticity

S.G. Roberts

1: 48

Menu

Climb and jogs

xx

xx

Mechanical driving force: work is done by the applied stress as the dislocation climbs. F = b xx (or full P-K for more complex stress states)

Vacancy diffusion to or from edge dislocations will cause them to climb . In the absence of a driving force, there will be equal diffusion rates to and from the dislocations core no net climb.

Chemical driving force: excess vacancy concentration in crystal. Dislocation core can act as vacancy sink.
1: 49

Steve Roberts Microplasticity

S.G. Roberts

Climb and Jogs

xx
etc.

xx

Climb will not occur uniformly along the dislocation line. Each vacancy that lands creates a pair of unit-height jogs. Line tension forces and mutual elastic attraction of the posistive and negative jogs will provide forces tending to remove them (removal still needs vacancy absorption or emission) The climb process can be visualised in terms of the creation, sideways diffusion and recombination of jogs. Jogs can also be created by dislocation intersections, and can be strong pinning points opposing dislocation glide.
Steve Roberts Microplasticity S.G. Roberts 1: 50

Menu

Dislocations in cubic close-packed metals


Close packed cubic pure metals: aluminium, copper, nickel, lead, silver, gold, palladium Alloy systems with close packed cubic phase: aluminium alloys, -brass, (some) stainless steels, nickel-based alloys

N.B. these are often called f.c.c. metals: while this is a valid description (the lattice type is facecentred cubic) it can lead to confusion. Cubic closed packed metals (facecentred cubic) have intersecting close-packed {111} glide planes. Close packed directions are <110>; Burgers vectors a/2 <110> Peierls valleys (weak) along <110> Other materials have a f.c.c. lattice but are not close packed. (e.g. Si, Ge, C). There are several similarities in dislocation structure and behaviour between the f.c.c. metals and the covalent f.c.c. materials, but also many differences.

Steve Roberts Microplasticity

S.G. Roberts

3:1

Cubic close-packed metals


[001] [001] [010] Unit cell, looking down onto the (111) plane.

[100]

[010]

[100]

Steve Roberts Microplasticity

S.G. Roberts

3:2

Menu

Cubic close-packed metals


[001] Unit cell, looking down onto the (111) plane. <110> slip directions on this plane.

[110]

[011]
[100]

[101]
[010] (111)

Steve Roberts Microplasticity

S.G. Roberts

3:3

Cubic close-packed metals


[001] Unit cell, looking down onto the (111) plane. <110> slip directions on this plane. a/2 <110> perfect Burgers vectors on this plane. a 2 [110]

a [011] 2

a [101] 2

[100]

[010]

Steve Roberts Microplasticity

S.G. Roberts

3:4

Menu

Cubic close-packed metals


[001] Unit cell, looking down onto the (111) plane. <110> slip directions on this plane. a/2 <110> perfect Burgers vectors on this plane. a 2 [110] a/6 <211> partial Burgers vectors on this plane.

a [011] 2

a [101] 2

a 2 [110] a [121] 6 a [211] 6

[100]

[010]

a [011] 2

a [112] 6

a [101] 2

Steve Roberts Microplasticity

S.G. Roberts

3:5

Dissociation of dislocations in ccp metals

The perfect a/2<110> Burgers vector moves atoms in the upper layer to positions occupied by other atoms in the same layer The crystal structure is unchanged by the passage of a dislocation with this Burgers vector; the dislocation displaces atoms by a whole lattice vector.

The partial a/6<211> Burgers vector moves atoms in the upper layer to positions not normally occupied by atoms in either of the layers shown. The crystal structure is changed by the passage of a dislocation with this Burgers vector, which is not a whole lattice vector. A Stacking Fault is produced. Another partial dislocation is needed to remove the stacking fault.

Steve Roberts Microplasticity

S.G. Roberts

3:6

Menu

Dissociation of dislocations stacking faults

Perfect a/2<110> Burgers vector: The green layer has been displaced by the passage of a perfect dislocation, carrying with it the orange layer above it . Note the 3-layer stacking repeat (ABC.), normal for the ccp structure. No energy change.
Steve Roberts Microplasticity

Partial a/6<211> Burgers vector: The green layer has been displaced by the passage of a perfect dislocation, carrying with it the orange layer above it . Note the local 2-layer stacking repeat (ABA), characteristic of the hcp structure. Higher energy.
S.G. Roberts 3:7

Motion of partials & stacking fault

The picture shows just the two layers either side of the active slip plane. In both the animation and the picture, all the distortions are drawn confined to one atomic plane. In reality, the distortions will be (nearly) equally distributed around the dislocation lines, and will decrease with distance from the dislocations. These distortions are the dislocations strain fields.

The animation shows just the two layers either side of the active slip plane. A pair of partial dislocations pass between these two planes. The first partial moves atoms into faulted positions; the second moves the atoms into normal positions (but translated from their original positions by a/2<110>)
Steve Roberts Microplasticity

S.G. Roberts

3:8

Menu

Energetics of partials & stacking fault


(111) The stacking fault has a higher energy than the normal crystal structure so why is is there at all ? The partial dislocations bounding it have a combined energy lower than that of the perfect dislocation.

Gb 2 2

Partial dislocations:

b= a [2 1 1] 6

b= a [1 2 1] 6 b= a [1 1 0] 2

G a 2 E 2 [12 + 22 + 12 ] 2 6 = Ga 2 6
Ga 2 2 [1 + 1 ] 2 2 Ga 2 4
S.G. Roberts
2

Perfect dislocation:

The energy calculations for the partial dislocations assume that the partials are completely separated (I.e. stacking fault is infinitely wide) As stacking fault width increases (and stacking fault energy per unit length of line increases), dislocation energy decreases from Ga2/4 to Ga2/6.

E =

Steve Roberts Microplasticity

3:9

Energetics of partials & stacking fault


Energy / length Criteria for dissociation:
2

Ga 4

r ne le a ot

gy

1. 2.
y

Ga 2 6

a stacking fault structure of low energy partial dislocations bounding the stacking fault which are of lower combined line energy than the perfect dislocation.

dislo catio n

energ

rg ne lt e u fa

In ccp metals both these criteria are met.

d Separation of Partials There is a minimum in the energy of the system at a non-zero fault width, d. This is because the stacking faults between a/6<211> dislocations in ccp metals are of relatively low energy. (e.g. Cu: 70 mJ m-2) This is because the ccp and hcp structures are very similar, and because bonding in close packed metals is largely non-directional.
Steve Roberts Microplasticity

Since the geometry of the close-packed plane means that the angle between the partials must be 120, we can see by inspection that b2 for the partial dislocations must be less than b2 for the perfect dislocations.

b= a [2 1 1] 6
120

b= a [1 2 1] 6 b= a [1 1 0] 2

S.G. Roberts

3:10

Menu

Stacking faults and partial dislocations: forces


(111) At equilibrium spacing, d, forces must balance: 1. 2. attractive force, = (stacking fault energy / unit area) repulsive force F between dislocations (varies with d)

Approximate treatment:

b= a [2 1 1] 6

b= a [12 1] 6

Gb 2 4d

(treats partials as screws)

At equilibrium:

Gb 2 = 4d 4 d= Gb 2

For copper, G = 48 GPa, = 70 mJm-2,, b = 0.26 nm, d giving d 3.7 nm. (See Qsheet 2 for more exact treatment)
Steve Roberts Microplasticity S.G. Roberts 3:11

Partial dislocations and stacking faults: images

Bubble raft

500 nm Cu 7% Al HREM (Si)

Triple junction in AlN 250 nm


Steve Roberts Microplasticity

[N.B. - most of these images are not from ccp metals!]


S.G. Roberts 3:12

Menu

Building Thompsons Tetrahedron


[001] Unit cell, looking down onto the (111) plane. <110> slip directions on this plane. a/2 <110> perfect Burgers vectors on this plane. a 2 [110] a/6 <211> partial Burgers vectors on this plane.

a [011] 2

a [101] 2

a 2 [110] a [121] 6 a [211] 6

[100]

[010]

a [011] 2

a [112] 6

a [101] 2

Steve Roberts Microplasticity

S.G. Roberts

3:13

Thompsons tetrahedron
[001] All possible slip planes and dislocations can be conveniently represented on a tetrahedron. Faces (a,b,c,d): Edges (AB, etc.): (111): c A (111): b a [011] 2 [100] D a [101] 2 [010] a [110] 2 C (111): d a [011] 2 a 2 [110] B slip planes; perfect Burgers vectors Peierls valleys Vertex-centre (A, B, etc): Shockley partial dislocations

a [101] 2

(111): a a [011] 2 A

a 2 [110] a [121] 6 a [211] 6

a [112] 6

a [101] 2

C
Steve Roberts Microplasticity S.G. Roberts 3:14

Menu

Make your own Thompsons tetrahedron

Puzzle: one direction is wrongly indexed ! C


D
C C

(by Hirth & Lothe)

Steve Roberts Microplasticity

S.G. Roberts

3:15

Partial dislocations and the Thompson Tetrahedron


b? B A

The partials have to pass through the structure in the correct order to create this intrinsic stacking fault. (If the partials act in the other order, a highenergy fault would result.) This will be important if we are considering dislocation reactions we have to know which partial dislocations will meet of the 4 possible combinations.
Steve Roberts Microplasticity S.G. Roberts

D C If we have an edge dislocation on the ABC plane with l in the C direction, is b = AB or BA ?

3:16

Menu

FSRH Convention and dissociation order


l
FSRH Convention B A b We use the FSRH Convention, so if the dislocation has its extra half-plane outside the ABC (111) plane, then b = BA. This will split into B and A.

The dissociation order rule is then: Imagine you are standing on the outside of the tetrahedron looking down the dislocation line in the sense of the line vector. The Greek-Roman partial is on your left, and the Roman-Greek on your right. So, in this case:

A BA D C C
Steve Roberts Microplasticity S.G. Roberts

3:17

Lomer-Cottrell lock
Two dislocations with the same line vector CD, but each gliding on its own plane. B A Dislocation 2: Dislocation 1: l = CD, b1 = BC Splits into B and C l = CD, b1 = CA Splits into C and A

Use the ordering rule. b2 D C

b1

Steve Roberts Microplasticity

S.G. Roberts

3:18

Menu

Lomer-Cottrell lock
Two dislocations with the same line vector CD, but each gliding on its own plane. B A Dislocation 2: Dislocation 1: l = CD, b1 = BC Splits into B and C l = CD, b1 = CA Splits into C and A

Use the ordering rule. C C B D A

Steve Roberts Microplasticity

S.G. Roberts

3:19

Lomer-Cottrell lock
Two dislocations with the same line vector CD, but each gliding on its own plane. B A Dislocation 2: Dislocation 1: l = CD, b1 = BC Splits into B and C l = CD, b1 = CA Splits into C and A

Use the ordering rule. C C D A + B C - The stable product is an immobile dislocation Reaction will give A + B = BA along CD. This is a low energy dislocation (b = a[110]). - The reaction will give rise to a stable product. There is no glide plane containing BA and CD.

Steve Roberts Microplasticity

S.G. Roberts

3:20

10

Menu

Cross-slip
B A A screw dislocation on plane BCD (a plane)

b = DC

Steve Roberts Microplasticity

S.G. Roberts

3:21

Cross-slip
B A Splits into two partials: DC D + C A screw dislocation on plane BCD (a plane)

Steve Roberts Microplasticity

S.G. Roberts

3:22

11

Menu

Cross-slip
B A Splits into two partials: DC D + C Following the order of partials rule, they must be in the order as shown. But this dislocation is now a planar object it can only exist in this form in the BCD plane. The cross-slip plane is ADC. C C A screw dislocation on plane BCD (a plane)

Steve Roberts Microplasticity

S.G. Roberts

3:23

Cross-slip
B A Splits into two partials: DC D + C Following the order of partials rule, they must be in the order as shown. But this dislocation is now a planar object it can only exist in this form in the BCD plane. The cross-slip plane is ADC. DC D C C To cross-slip, the partials must recombine to create a pure screw dislocation. A screw dislocation on plane BCD (a plane)

Steve Roberts Microplasticity

S.G. Roberts

3:24

12

Menu

Cross-slip
B A Splits into two partials: DC D + C Following the order of partials rule, they must be in the order as shown. But this dislocation is now a planar object it can only exist in this form in the BCD plane. The cross-slip plane is ADC. C C D To cross-slip, the partials must recombine to create a pure screw dislocation. This can then glide on the ADC plane and it will immediately dissociate: DC D + C A screw dislocation on plane BCD (a plane)

Steve Roberts Microplasticity

S.G. Roberts

3:25

Cross-slip
B A Splits into two partials: DC D + C Following the order of partials rule, they must be in the order as shown. But this dislocation is now a planar object it can only exist in this form in the BCD plane. The cross-slip plane is ADC. C D C C To cross-slip, the partials must recombine to create a pure screw dislocation. This can then glide on the ADC plane and it will immediately dissociate: DC D + C A screw dislocation on plane BCD (a plane)

Steve Roberts Microplasticity

S.G. Roberts

3:26

13

Menu

Cross-slip
B A Splits into two partials: DC D + C Following the order of partials rule, they must be in the order as shown. But this dislocation is now a planar object it can only exist in this form in the BCD plane. D D C The cross-slip plane is ADC. C C To cross-slip, the partials must recombine to create a pure screw dislocation. This can then glide on the ADC plane and it will immediately dissociate: DC D + C The cross-slip continues by sideways propagation of the constricted nodes.
Steve Roberts Microplasticity S.G. Roberts 3:27

A screw dislocation on plane BCD (a plane)

Cross-slip
B A To allow cross-slip to happen, this constriction of the dissociated dislocation must form. The partials which mutually repel each other must be brought together. This requires a local stress.

D DC D C C

Steve Roberts Microplasticity

S.G. Roberts

3:28

14

Menu

Cross-slip
B A To allow cross-slip to happen, this constriction of the dissociated dislocation must form. The partials which mutually repel each other must be brought together. This requires a local stress. This may happen if there is an obstacle to glide, and possibly also a pile up. D C C D The stacking fault energy will assist the constriction force. Materials with a high stacking fault energy will have narrow stacking faults that constrict relatively easily. Easy cross-slip. Materials with a low stacking fault energy will have wide stacking faults that constrict with difficulty. Cross-slip more difficult.

Steve Roberts Microplasticity

S.G. Roberts

3:29

Cross-slip so what ?
Ease of cross-slip (or the lack of it) has strong effects on many aspects of mechanical behaviour. Monotonic loading A: early stages of workhardening - glide band spreading - formation of new dislocation sources - by-passing precipitates B: later stages of workhardening - as (A) and - bypassing of locks and dislocation tangles Strain Fatigue: - irreversible slip in cyclic loading - stability of persistent slip band structures

Stress

Steve Roberts Microplasticity

S.G. Roberts

3:30

15

Menu

Frank loops & dislocations


C B A C B A C

CCP layers seen edge on, with a random arrangement of vacancies. (more vacancies than are in thermal equilibrium)

C B A C B A C C B A C B A C

Redrawn showing just the vacancies.

Vacancies can lower their energy by forming a disc-shaped planar cluster.

C B A C B A C A

C B B A C

The adjacent planes collapse into the vacancy sheet. An intrinsic stacking fault is formed, bounded by a pure edge dislocation loop.
S.G. Roberts 3:31

Steve Roberts Microplasticity

Frank loops & Thompsons tetrahedron


Because the dislocation loop surrounds a stacking fault, it must be a partial dislocation. The Burgers vector is the same all way round the loop. b {111} b The Burgers vector is normal to a {111} plane, and is of length equal to one {111} interplanar spacing.

b=

a < 111 > 3

b B This is represented on the Thomson tetrahedron by a vector of type D.

D =

a [111] 3

The faulted loop would be on the (111) - ABC plane. D C


Steve Roberts Microplasticity S.G. Roberts 3:32

16

Menu

Frank loops - removal


b= a < 111 > 3
{111} b The dislocation is sessile. b The loop can only shrink or grow by removal or addition of vacancies i.e. by climb. The Burgers vector of the dislocation bounding the stacking fault does not lie on any {111} glide plane.

Another way the stacking fault can be removed is by a dislocation reaction. The boundary dislocation has Burgers vector (say) D. If this combines with a Shockley partial dislocation of Burgers vector (say) A, a perfect dislocation is formed. D + A AD B
S.G. Roberts 3:33

A
Steve Roberts Microplasticity

Frank loops - conversion


D Frank loop, b= D, with stacking fault.

Steve Roberts Microplasticity

S.G. Roberts

3:34

17

Menu

Frank loops - conversion


D Frank loop, b= D, with stacking fault. Shockley partial loop, b= A, nucleates within the Frank loop; B D The Shockley partial can glide in the plane of the loop.

AD

Steve Roberts Microplasticity

S.G. Roberts

3:35

Frank loops - conversion


D Frank loop, b= D, with stacking fault. Shockley partial loop, b= A, nucleates within the Frank loop; B The Shockley partial can glide in the plane of the loop. The Shockley partial traverses the Frank loop

AD

AD

AD

Steve Roberts Microplasticity

S.G. Roberts

3:36

18

Menu

Frank loops - conversion


D Frank loop, b= D, with stacking fault. Shockley partial loop, b= A, nucleates within the Frank loop; B The Shockley partial can glide in the plane of the loop. The Shockley partial traverses the Frank loop, and completely coneverts it to a perfect dislocation loop, b= AD. Note that the perfect lattice dislocation can glide but only in the ADC or ADB planes. It cannot glide in its own loop plane. This is called a prismatic dislocation loop.

AD

AD

AD

AD

Steve Roberts Microplasticity

S.G. Roberts

3:37

Dislocations in hexagonal close-packed metals


c/a ratio of ideal structure = (8/3) = 1.633 Metals with c/a 1.6333 Basal slip is favoured. Slip systems are (0001)<1120> Metal c/a CRSS on (0001)<1120> 0.18 MPa 0.43 MPa 0.57 MPa Metals with c/a < 1.6333 Prism plane slip is favoured. Slip systems are {1100}<1120> Metal c/a CRSS on CRSS on (0001)<1120> {1100}<1120> 110 MPa ? 49 MPa 6 MPa

Zn Mg Cd

1.856 1.624 1.886

Ti Zr

1.587 1.593

Can also get slip on {1101}<1120>

(0001)

(1010) [1210] [1210] [2110]


S.G. Roberts

(1101) [1210] [2110] [1120]


3:38

[2110]
Steve Roberts Microplasticity

[1120]

[1120]

19

Menu

Basal slip in hcp metals

Basal slip is very similar to {111}<110> slip in ccp metals. The Burgers vector is that of closest atomic spacing in the close-packed plane. Perfect dislocations will split into partial dislocations.

Action of a single partial dislocation produces a stacking fault. The stacking fault locally has a three-fold layer repeat a few layers of ccp structure.

a a a [11 2 0] [10 1 0] + [01 1 0] 3 3 3


Steve Roberts Microplasticity S.G. Roberts 3:39

Non-basal-plane slip in hcp metals

(1010) [1210] [2110] [1120]

There is no possible stacking fault arrangement on these crystal planes. However, a low energy fault is possible on the basal plane. Even though dislocations glide on the prism or pyramidal plane, the relaxed core structure is extended on the basal plane. The dislocations have non-planar cores, which must rearrange before glide can happen.

(1101) [1210] [2110] [1120]

This is why non-basal glide requires a high critical resolved shear stress.

Steve Roberts Microplasticity

S.G. Roberts

3:40

20

Menu

Dislocations in body-centred cubic metals


[001] BCC metals include: Fe, Mo, W, Ta, V, Cr, Nb, Na, K Slip always has Burgers vector b =

a 111 2

Possible slip planes are {110}, {112}, {123}. Any given Burgers vector lies in: 3 {110} planes 3 {112} planes 6 {123} planes

[100]

(101)

b=

a [111] 2

There is no low-energy stacking fault structure on any of these planes. Dislocations do not dissociate.

[010]

Steve Roberts Microplasticity

S.G. Roberts

3:41

Screw dislocations in bcc metals cross slip

Cross-slip of screw dislocations is very easy: - there are no possible stacking faults, so the screw dislocations are not dissociated; - there are very many possible slip planes for a given dislocation.
Video shows cross-slip of screw dislocations in Mo Steve Roberts Microplasticity

This ready cross-slip means that screw dislocations will follow the easiest path through a bcc crystal, changing frequently between slip lanes to avoid obstacles. On the optical microscope scale, this appears as wavy slip.

S.G. Roberts

3:42

21

Menu

Screw dislocations in bcc metals core structure


(011) [121] [211] Screw dislocations in bcc metals have non-planar cores. Illustrations show atomistic simulations of two possible core structures. The whole dislocation has b = a [111] (normal to plane of the page) Arrows at each atom position indicate displacement from that position when screw dislocation is present at the centre. The shifts are concentrated equally along three of the potential slip directions. This is not dissociation there are no well separated dislocations bounding stacking faults. The core is just delocalised.

(101)

[112]

(111) plane (110)

Steve Roberts Microplasticity

S.G. Roberts

3:43

Screw dislocations in bcc metals glide

The screw dislocation core is spread over three potential glide planes. If a shear stress is applied, before the dislocation can glide, it has to re-arrange itself to be nearly planar before glide can occur. Pictures show atomistic simulations of changes in core structure under stress.

Results: Glide of screw dislocations is difficult. Thermal activation will assist core re-arrangements and hence critical resolved shear stress for glide. Schmids law is not obeyed as: Glide in one sense is easier than glide in the other (the tension-compression asymmetry) Non-shear stresses can affect core rearrangements and hence glide stress.

Steve Roberts Microplasticity

S.G. Roberts

3:44

22

Menu

Dislocation loops in bcc metals


Edge dislocations in bcc metals have undissociated but planar cores. They can move much more readily than screw dislocations. Consequences: Dislocation loops tend to have long, straight, slow moving screw segments and short, rapidly moving edge segments. Operation of dislocation sources is be restricted by the rate at which screw dislocations can get out of the way.
Videos: Upper: Molybdenum - movement of fast edge dislocation trailing long screw dislocation; Lower: Molybdenum - operation of half-Frank-Read dislocation source
Courtesy of I. Robertson et al., University of Illinois at Urbana-Champaign

Steve Roberts Microplasticity

S.G. Roberts

3:45

Dislocations in bcc metals - general


CRSS (MPa) 30 0.01% C 20

Iron

Dislocation mobility in bcc metals varies strongly with impurity content: - even a few ppm of interstitial impurities give Cottrell atmospheres. - stress fields around interstitial solutes in bcc have strong shear as well as dilatational components interact strongly with both edge and screw dislocations. - thermal activation is needed to pull dislocation cores off these solutes.

0.001% C "pure"

10

0 0 50 100 150 200 250 300 Temperature (K)

Dislocation mobility in bcc metals varies strongly with temperature (especially at low temperatures): - fairly directional bonding means moderate activation energy for glide even of edge dislocations - thermal activation is needed to put screw dislocations into glide configuration
Steve Roberts Microplasticity

S.G. Roberts

3:46

23

Menu

Intermetallic compounds: structures


[001] L12 structure. Cu3Au, Ni3Al, etc. [001] L10 structure. TiAl

[100]

[010]

[100]

[010]

B2 structure. CuZn

There are very many more crystal structures of a huge variety of intermetallic materials. Some are closely related to the simple structures exhibited by elemental metals (superlattices). Others are very complex.

Steve Roberts Microplasticity

S.G. Roberts

3:47

Ordered intermetallics: superlattice dislocations

The slip planes in the L12 structure (e.g. Ni3Al) are {111} just like in ccp metals. But the lattice is now simple cubic the shortest lattice vector in this plane is a<110>. The normal a/2 <110> vector moves an (e.g.) Al atom on the site of a (e.g.) Ni atom. These a/2 <110> dislocations can also split into a/6 <211> partial dislocations.
Steve Roberts Microplasticity S.G. Roberts

Al Ni

3:48

24

Menu

Dislocation structure in L12 intermetallics


Stacking fault Anti-phase boundary Stacking fault

Al Ni

The line energy = Gb2/2 criterion favours splitting of the lattice dislocation into the four a/6 <211) partials.

A complicated structure results the partials repel each other elastically but are held together by the energies of the resulting stacking faults. The central fault, where Al is swapped onto on Ni site, is called an Anti-phase boundary (APB).

Steve Roberts Microplasticity

S.G. Roberts

3:49

APB energies and the yield stress anomaly


In some L12 materials, notably Ni3Al, the APB has a lower energy on the {100} planes than on the {111} planes. Anti-phase boundary on (111) So, given some thermal activation to pinch the leading stacking fault,

Steve Roberts Microplasticity

S.G. Roberts

3:50

25

Menu

APB energies and the yield stress anomaly


In some L12 materials, notably Ni3Al, the APB has a lower energy on the {100} planes than on the {111} planes. Anti-phase boundary on (111) So, given some thermal activation to pinch the leading stacking fault,

Steve Roberts Microplasticity

S.G. Roberts

3:51

APB energies and the yield stress anomaly


Anti-phase boundary on (100) In some L12 materials, notably Ni3Al, the APB has a lower energy on the {100} planes than on the {111} planes. So, given some thermal activation to pinch the leading stacking fault, a screw dislocation can cross slip onto a {100} plane.

Anti-phase boundary on (111)

Steve Roberts Microplasticity

S.G. Roberts

3:52

26

Menu

APB energies and the yield stress anomaly


Anti-phase boundary on (100) In some L12 materials, notably Ni3Al, the APB has a lower energy on the {100} planes than on the {111} planes. So, given some thermal activation to pinch the leading stacking fault, a screw dislocation can cross slip onto a {100} plane.

Steve Roberts Microplasticity

S.G. Roberts

3:53

APB energies and the yield stress anomaly


Anti-phase boundary on (100) In some L12 materials, notably Ni3Al, the APB has a lower energy on the {100} planes than on the {111} planes. So, given some thermal activation to pinch the leading stacking fault, a screw dislocation can cross slip onto a {100} plane. It will finally end up as shown Two screw dislocations dissociated on the glide plane, but connected by an low-energy APB on a non-glide plane. This dislocation can only move by recross-slipping to the higher-energy planar glide configuration.

Stacking fault on (111)

Steve Roberts Microplasticity

S.G. Roberts

3:54

27

Menu

APB energies and the yield stress anomaly


In some L12 materials, notably Ni3Al, the APB has a lower energy on the {100} planes than on the {111} planes. So, given some thermal activation to pinch the leading stacking fault, a screw dislocation can cross slip onto a {100} plane. It will finally end up as shown Two screw dislocations dissociated on the glide plane, but connected by an low-energy APB on a non-glide plane. This dislocation can only move by recross-slipping to the higher-energy planar glide configuration. Locked dislocation with kinks At higher temperatures, more segments of dislocations transfer to the non-glide arrangement. The Yield stress rises.
Steve Roberts Microplasticity S.G. Roberts 3:55

Dislocations in ionic crystals

Bonding is non-directional, Shear displacements of atoms when dislocation cores move is easy except that strong fluctuations of energy will occur as ions of like or opposite charge move closer to each other. These energy fluctuations will determine flow stresses and even slip systems.

O2-, Cl-, F- Mg2+, Na+, Li+,

Steve Roberts Microplasticity

S.G. Roberts

3:56

28

Menu

Dislocations in ionic crystals

In the rock-salt structure, {111} planes contain ions all of the same charge. Dislocations gliding on {111} would have very strong charge interactions. Slip on {111} isnt observed other systems are much easier to operate.

O2-, Cl-, F- Mg2+, Na+, Li+,

Steve Roberts Microplasticity

S.G. Roberts

3:57

Dislocations in ionic crystals

In the rock-salt structure, {110} planes are charge-neutral, and contain the shortest lattice vector.

b=

a 110 2

Motion of dislocations with this Burgers vector do not bring like charges too close together. Slip on this system is fairly easy.

O2-, Cl-, F- Mg2+, Na+, Li+,

Steve Roberts Microplasticity

S.G. Roberts

3:58

29

Menu

Dislocations in ionic crystals

(100) In the rock-salt structure, {110} planes are charge-neutral, and contain the shortest lattice vector.

b=

a 110 2
[011] [011] Edge dislocation in rock-salt structure, on {110}<110> slip system. (alternate layers have atoms in same positions, but with opposite charges)

Motion of dislocations with this Burgers vector do not bring like charges too close together. Slip on this system is fairly easy.

Steve Roberts Microplasticity

S.G. Roberts

3:59

Dislocations in ionic crystals

In the rock-salt structure, {100} planes are charge-neutral, and contain the shortest lattice vector.

b=

a 110 2

Motion of dislocations with this Burgers vector do not bring like charges very close together, but planes are more closely spaced than {110} planes. Charge interactions are stronger than for {110} slip. Slip on this system is possible, only at high stresses.

O2-, Cl-, F- Mg2+, Na+, Li+,

Steve Roberts Microplasticity

S.G. Roberts

3:60

30

Menu

Dislocations in ionic crystals


C.R.S.S. (MPa) 500 400 300 200 100 <110>{110} 0 0 500 1000 Temperature (K) 1500
-

MgO is typical of a wide variety of ionic materials: Slip systems determined by strength of ionic effects on dislocation motion as much as by straight crystallography. Ionic repulsion as dislocation moves makes even easy slip systems difficult to operate - thermal activation can make slip easier Limited number of active slip systems at low temperatures - may not meet Von Mises criterion Charged impurities (e.g. Fe3+, Cr3+) have very large effects on flow stress even at ppm levels.

<110>{001}
-

Steve Roberts Microplasticity

S.G. Roberts

3:61

Dislocations in covalent crystals


Covalent bonds have well-defined lengths and interbond angles: For materials with the FCC diamond cubic structure: C: Si: Ge: SiC: GaAs: : 0.154 nm, 0.235 nm, 0.245 nm, 0.307 nm, 0.400 nm, 109.4 109.4 109.4 109.4 109.4

Steve Roberts Microplasticity

S.G. Roberts

3:62

31

Menu

Dislocations in covalent crystals


Covalent bonds have well-defined lengths and interbond angles. For materials with the FCC diamond cubic structure: C: Si: Ge: SiC: GaAs: : 0.154 nm, 0.235 nm, 0.245 nm, 0.307 nm, 0.400 nm, 109.4 109.4 109.4 109.4 109.4

As in the cubic close-packed metals, Glide planes: Burgers vectors: {111} a/2 <110>

(can still use Thompsons tetrahedron)

But these are not close-packed planes or directions.

Steve Roberts Microplasticity

S.G. Roberts

3:63

Peierls valleys and dislocation types in diamond structure


Dislocation cores are highly structured: Core energy depends strongly on direction. Dislocations lie along <110>.

[ Electron Beam Induced Conductivity (EBIC) image in SEM. Demonstrates dislocations are electrically active. Important in semiconductor devices.] P.R. Wilshaw 1989

60

screw

60 Dislocations thus are likely to be either pure screw or 60 type. What will their core structure be ?

60

screw

60

Steve Roberts Microplasticity

S.G. Roberts

3:64

32

Menu

Peierls valleys and dislocation types in diamond structure


60 <110> screw 60 As in ccp metals, dislocations in diamond structure are always dissociated e.g.:

a a a [1 1 0](111) [2 1 1](111) + [1 2 1](111) 6 6 2


60 screw 60 Dissociation width depends on shear modulus G and stacking fault energy in the usual way.

30 90

The (a/6)<211> partial Burgers vectors are at 30 to the perfect (a/2)<110> vector. Dissociation thus produces two types of dislocation: 30 Burgers vector at 30 to line 90 pure edge. So what structure do these have ?

Steve Roberts Microplasticity

S.G. Roberts

3:65

Dislocation core structure in diamond structure


Looking down the line direction of a dissociated 60 dislocation. Stacking fault is visible as two sets of ABA bonds.

FCC structural unit Stacking fault structural unit High Resolution Electron Microscopy (HREM) image of this type of stacking fault.

Steve Roberts Microplasticity

S.G. Roberts

3:66

33

Menu

The Dislocation in Literature


Only infinite jumble and mess and dislocation, Backed by a solemn appeal, 'For God's sake, do not stir, there!'
Arthur Hugh Clough, The Bothie of Tober-na-vuolich: a Long-vacation Pastoral.

Finest dislocations of the avant-garde turn out to be the simplest captions of disease.
Peter Porter, The Worst Inn's Worst Room (1994)

In Pope, there are no disharmonies and dislocations ever and anon to grate and interrupt; all in him is perfect.
James Stirling, Burns in drama (1878)

Dislocation, perhaps!" muttered the Doctor: "let us hope there is no worse injury done.
Charlotte Bront, Villette (1853)

"Any dislocation, and we are thrown out again! We must hold together if this riddle is ever to be read.
George Meredith, The Egoist (1879)

Every half step he attempted was like a dislocation. His groans and grunts were frightful.
George Meredith, The Ordeal of Richard Feverel (1859)

Thomson went up to assist Mr. Morgan in the reduction of the dislocation. --- When this was successfully performed, they wished me joy
Tobias Smollett, Roderick Random (1748)

One of the terrible dislocations of our habits of mind.


Joseph Sheridan Le Fanu, Uncle Silas (1864).

Steve Roberts Microplasticity

S.G. Roberts

3:67

34

Menu

Why do materials have any strength ?


Calculations show that it should only need a few Nm-2 to move dislocations about in metals - and experiments on pure metals show that this is the case. Strength restricted dislocation motion.

Effects of bonding type on ease of dislocation motion. choice of basic material Grain boundaries as blocks to dislocation motion grain refinement hardening (and toughening) Internal stress fields from dissolved atoms interact with dislocation stress fields solute hardening Internal stress fields from precipitates interact with dislocation stress fields precipation hardening (age hardening) Precipitates with hard crystal structures act as local blocks to dislocations precipation hardening (age hardening) Immobile dislocations block mobile dislocations work hardening Use phase chages in basic material to produce fine, strained microstructures quenching
S.G. Roberts 5: 1

Microplasticity 5 Strengthening

Effects of bonding type on yield strength


Yield Strength - resistance to slip by
dislocation motion increases as the directionality and rigidity of interatomic bonding increases. Increasing directionality of bonding: increasing yield stress and hardness Close-packed metals e.g.: copper, aluminium Nearly nondirectional bonding. Dislocation motion easy Other metals iron, tungsten Some directionality in bonding. Dislocation motion fairly easy (varies with temperature) Ionic compounds NaCl, MgO Bonding by electrostatic attraction: not directional but slip moves strongly charged ions past each other. Dislocation motion fairly difficult Intermetallic compounds CuAl2, TiAl, Ni3Al Ordered structures with fairly directional bonding. Dislocation motion difficult. Can use as fine dispersions to harden metallic alloys Covalent solids diamond, SiC, Al2O3 Ceramics: Very strongly directional bonding. Dislocation motion very difficult. Can use in composites

Microplasticity 5 Strengthening

S.G. Roberts

5: 2

Menu

Effects of Bonding type Isomechanical Groups


Hardness (MPa) 104
Ge Si

Hardness / Youngs Modulus 10-1

103
W

Si Ge

10-2

102
Pb Al

Fe

Mo Al Pb

Cu

10
0 400 800 1200 1600 2000

10-3

W Mo 0 0.2 0.4 Fe 0.6 Cu 0.8

Temperature (K)

T / Tmelt

Microplasticity 5 Strengthening

S.G. Roberts

5: 3

Hardening by substitutional solutes


Possible interactions between Dislocations and Solutes Elastic
Stress field of dislocation interacts with stress field around solute

Modulus
Dislocation energy varies where solute locally changes bond stiffness

Stacking fault
Stacking fault energy varies if solute present

Electrical
Dislocations with charged cores (ionic materials) will have strong interactions with charged point defects

Short-range order & Long-range order


Motion of dislocation / stacking fault will change local ordering, hence energy.

In all cases, any interaction between dislocation and solute will harden the material: - attractive dislocations pinned near solute atoms - repulsive dislocations have to be pushed past solute atoms
Microplasticity 5 Strengthening S.G. Roberts 5: 4

Menu

Solution hardening
Interaction energies between stress fields of the moving dislocation and the static misfitting solute atoms give rise to widely varying dislocation enrgy with position. Since

Energy

Dislocation Position

F=

dE dx

This implies a high stress to move dislocations through such an array of obstacles. [Any array of small regions with associated stress fields will give similar effects]

Microplasticity 5 Strengthening

S.G. Roberts

5: 5

Substitutional solid solution hardening in 3D


Case shown is for solute atoms smaller than the parent atoms. C T
Solutes below the dislocations slip plane Solutes above the dislocations slip plane (N.B. only one sort of solute atom colour is just to indicate position.)

2 1

Minimum energy configurations of dislocations are a compromise between line tension (energy / length) and attraction to / repulsion from solute atoms.

Microplasticity 5 Strengthening

S.G. Roberts

5: 6

Menu

Strong diffuse obstacles interacting via long range stresses


Mean stress on dislocations (the internal stress) = i

Gb 2R Gb R= 2 i i =

Dislocation line tends to bypass obstacles one at a time.

If R< , the dislocation line can bend tightly enough for it to find a minimum energy shape through the array of stress centres. Strong Interactions control the shape of the dislocation line. Flow stress is that to overcome the mean internal stress

flow =

2 i

Microplasticity 5 Strengthening

S.G. Roberts

5: 7

Weak diffuse obstacles interacting via long range stresses


Mean stress on dislocations (the internal stress) = i

R=

Gb 2 i

If R > , the dislocation line cant bend tightly enough for it to weave between the obstacles. Line Tension is a strong influence on the shape of the dislocation line. L

Moving segments of the dislocation line are much longer (L) than the obstacle spacing . Number of obstacles in length L, n:

n=

In a random arrangement, there will be an excess of obstacles of one sense over the other = n
Microplasticity 5 Strengthening S.G. Roberts 5: 8

Menu

Weak diffuse obstacles interacting via long range stresses


Mean stress on dislocations (the internal stress) = i Over length L, we have n active obstacles, each acting over length

flow bL = flow

2 n ib
1

2 2 = i L

Estimate L by assuming advance distance = l, dislocation line curvature controlled by applied stress: L

= R R 2 (L / 2)2 2 L2 L2 flow = 4R 2Gb


R This gives:
1

flow
L
Microplasticity 5 Strengthening S.G. Roberts

2 i 3 = 4 i 3 Gb
5: 9

Weak diffuse obstacles interacting via long range stresses


0.50 50

0.40

40

flow i

0.30

30

0.20

20

= 10nm G = 40 GPa b = 0.3 nm

0.10

10

0 0 100

i
1

200

300

400

500

(MPa )

flow

2 i 3 = 4 i 3 Gb

For typical values of , G, b, i,

flow ~ 0.1 - 0.3 i


S.G. Roberts 5: 10

Microplasticity 5 Strengthening

Menu

Solute Dislocation Elastic Interactions: the maths from dislocation

hydrostatic stress H
rM rS
y

rF

r x

rM = size of matrix atom ( size of hole) rS = size of solute atom rF = final size

hydrostatic misfit stress & strain

Microplasticity 5 Strengthening

S.G. Roberts

5: 11

Solute Dislocation Elastic Interactions: the maths


Model: solute atom is elastic sphere of same E, G, as matrix.
rM rS rF

Radius before insertion = rS = (1 + ) rM. Radius after insertion = rF = (1 + ) rM.

Vmis = VS VM
2 = 4rM (rM ) 3 = 4rM

(1 + ) 3(1 )

V =

V = VF VM
3 = 4rM

(1 + ) 3 4rM 3(1 ) (1 + ) = Vmis 3(1 )

(
Microplasticity 5 Strengthening

Proof of this is an exercise for the student)


5: 12

S.G. Roberts

Menu

Solute Dislocation Elastic Interactions: the maths


Interaction Energy, Ei = P.V

y P r x

Ei =

Gb y (1 + ) 3 2(1 + ) 4rM 3 2(1 ) x 2 + y 2 3(1 )

However, this is for an infinite matrix: for a finite matrix, there is a correction factor of : 3(1 ) (1 + ) -effectively V = VH, = 4ra2 , and so
3 Ei = 4rM

Hydrostatic Stress from disln, P:


1 P=3 ( xx + yy + zz )

2(1 + ) Gb y 3 2(1 ) x 2 + y 2

2(1 + ) Gb y 3 2(1 ) x 2 + y 2

3 4Gb rM (1 + ) sin 3 (1 ) r

Microplasticity 5 Strengthening

S.G. Roberts

5: 13

Solute Dislocation Elastic Interactions: the maths


The glide force on the dislocation is:

Fx =
y P r x

d Ei dx 3 (1 + ) d y 4Gb rM 2 = 2 3 (1 ) dx x +y
3 8Gb rM xy (1 + ) 3 (1 ) x 2 + y 2

= =

3 8Gb rM (1 + ) sin cos 3 (1 ) r2

Ei = =

3 (1 + ) sin 4Gb rM 3 (1 ) r 3 4Gb rM (1 + ) y 3 (1 ) x 2 + y 2

The maximum glide force is at = 60:

Fx

MAX

=
(

3 (1 + ) 1 2Gb rM 2 3 (1 ) r

Proof of this is an exercise for the student)


5: 14

Microplasticity 5 Strengthening

S.G. Roberts

Menu

Solute Dislocation Elastic Interactions: the maths


y P r = 60 x

The maximum interaction energy is when r = b and = 90:


3 4G rM (1 + ) 3 (1 ) G M

Ei

MAX

Fx

MAX

3 (1 + ) 1 2Gb rM = 2 3 (1 ) r

The absolute maximum glide force exerted by the misfitting solute on the dislocation is when r = b and = 60:

These rough theoretical values are an overestimate, but are within an order of magnitude of real values. Typical values: EiMAX = 3 eV for C.P. metals EiMAX = 20 eV for covalent solids

Fx

MAX

3 2G rM (1 + ) 3 b (1 ) GM b

Microplasticity 5 Strengthening

S.G. Roberts

5: 15

Strength of solution hardening


Flow Stress (MPa) 6 30 4 20 2 0 10 0 Flow Stress (MPa)

Ag r = 0.2889 nm

Au r = 0.2884 nm

Al

3 %Cu

r = 0.2863 nm

r = 0.2556 nm

misfit, = ~ 0.17 %

misfit , = ~ 8 % Strength of hardening effect is (to a good approx.) proportional to: 1. amount of solute 2. solute misfit, 3. elastic modulus of matrix Effects of multiple solutes are roughly additive.
Atomic size data: www.webelements.com

Ei

MAX MAX

G M G M b

Fx

Microplasticity 5 Strengthening

S.G. Roberts

5: 16

Menu

Solution hardening interstitial solutes


A limited number of cases, but of great importance: C, N in Iron (body centred cubic) (and other b.c.c. metals) O in Titanium (close packed hexagonal) A
c

The atoms here are shown smaller than they actually are relative to the b.c.c. unit cell. The A-A spacing is smaller than the B-B spacing, and determines the size of interstitial atom that will fit.

rint. < 0.29 rhost


Actual radius ratios: B

rnitrogen riron

I B A
b

= 0.44

rcarbon = 0.46 riron

Result: each interstitial atom has a large and non-spherically symmetric stress field around it: Shears as well as dilatations.

Misfit , = ~ 10 % to 100%

Microplasticity 5 Strengthening

S.G. Roberts

5: 17

Interstitial solutes dislocation pinning


Three features of interstitial solutes in bcc metals make them powerful hardening agents: They have very large stress fields around them
so they can lower energy by sitting at dislocation cores

The stress fields have both shear and dilatational components


so they pin both edge and screw dislocations

They can diffuse rapidly, even at room temperature so they can find the dislocations in hours or days. High Stress applied

Solute diffusion to static dislocation

Dislocation pinned

Dislocation unpinned and mobile at lower stress

Microplasticity 5 Strengthening

S.G. Roberts

5: 18

Menu

Interstitial solutes and dislocations: the maths


y x Elastic Interaction Energy: r

EMAX i c core = c bulk exp kT

Ei =

3 (1 + ) sin 4Gb rM 3 (1 ) r

Take cbulk = 0.1 at%


100%

Ei

MAX

G M

80% 60%

Ei = 0.2eV

0.3eV 0.4eV 0.5eV

(theoretical value overestimate)

For interstitial solutes,


Misfit , = ~ 10 % to 100%

ccore
40% 20% 0

(substnl. solutes: 0% to ~14%)


Typical binding energy = 0.2 to 0.5 eV
Microplasticity 5 Strengthening

200

400

600

800

1000

Temperature (K)

S.G. Roberts

5: 19

Yield point and strain ageing


Stress (MPa) 500hr 300 4hr 0.5hr Lower Yield Stress U.Y.S. Upper Yield Stress

200 unload age reload 0.03%C steel, 60C 0 5 Strain (%)


Microplasticity 5 Strengthening S.G. Roberts 5: 20

Ageing

L.Y.S.

100

10

Menu

Solute stacking fault interactions


Hardening if solute atoms have different solubility in stacking fault than in matrix. With solute atoms with higher solubility in stacking fault:

c0 c0

A: (Non equilibrium) Dislocations and stacking fault. Solute atoms at concentration c0 throughout B: (Equilibrium) Solute atoms at concentration c0 in bulk, c1 ( > c0) in stacking fault. Stacking fault energy per unit area is lowered. Stacking fault widens; Dislocation stacking fault energy per unit length Dislocation elastic energy per unit length Overall dislocation energy To move dislocation, must leave solute cloud behind + go back to higher energy configuration B. Similar (though weaker) effects to Cottrell atmospheres

c1 c0

Suzuki effect

Microplasticity 5 Strengthening

S.G. Roberts

5: 21

Solutes modulus interactions


Point defects can change the local elastic modulus of the material
defect - matrix bonds may be softer or harder than those of pure matrix visible as a macroscopic change in modulus regions close to solute will have large change in modulus dislocation stress field energy locally changed elastic interaction between dislocation and point defect

Soft defects (e.g. vacancy) (modulus locally decreased)


attractive interaction with dislocations

Hard defects (modulus locally increased)


repulsive interaction with dislocations

[ Compare with image forces ]

Microplasticity 5 Strengthening

S.G. Roberts

5: 22

Menu

Modulus interactions the maths


Elastic energy E of any small volume element V Simplest case is for screw dislocations, with shear modulus locally changed at a radius r from G to G`:

E =

1 2

( )V
ij ij

( )
ij ij

ij

= ( ) = (G G)b b 2r 2r

Assume that the strain field around the dislocation stays the same, but that the local stress changes in response to modulus changes. Change in energy of this volume element::

(E) =

1 2

( ( ) )V
ij ij ij 4 ij 3

So that:

E = G
G =

Gb 2 rs3 6 r 2

Assume single solute atom, changing modulus over radius rs:

E =

1 2

( ( ) )
ij ij

where

rs3
S.G. Roberts

G G G

Microplasticity 5 Strengthening

5: 23

Modulus interactions the maths


y r x rs

Fx =

dE dx Gb 2 x rs3 = G 3 ( x 2 + y 2 )2

Interaction energy between dislocation and modulus-changed region around solute atom:

MAX Fx

Gb 2 rs3 cos 3 r 3 Gb 2 = G when r = rs and = 0 3 = G

Gb 2 rs3 E = G 6 r 2 Gb 2 rs3 = G 6 ( x 2 + y 2 )
Microplasticity 5 Strengthening

These effects, for both screw and edge dislocations, are generally weaker than elastic interactions. E decays as 1/r2 (for elastic interactions decays as 1/r). Can be important for chemically different solute atoms with small size mismatches.
S.G. Roberts 5: 24

Menu

Electrical solute dislocation interactions


(100) Non isovalent impurities create charged point defects: e.g. Cr3+ in Mg2+O2Mg O [011] [011] O Edge dislocation in rock-salt structure, e.g. MgO: {110}<110> slip system. (alternate layers have atoms in same positions, but with opposite charges) JOGS have a non-cancelled charge. Mg O Mg O Mg Mg O Cr O Mg O Mg O O Mg O Cr O

Mg O

For every two Cr3+ there must be one Mg2+ vacancy. These charged defects interact very strongly with the kinks on the dislocations: A few 100s of ppm charged impurity can double the hardness of MgO
S.G. Roberts 5: 25

Microplasticity 5 Strengthening

Hardening by precipitates
1. 2. 3. 4. 5. 6. 7. 8. Coherency strains Long range stress fields around coherent precipitates Semi-coherent precipitates Long and short-range interactions with interfacial dislocations Interfacial energy and morphology Cutting of precipitates increases their surface area Lattice friction stress Precipitates have higher (or lower) y than matrix Stacking fault energy Dissociated dislocations have different SFE in precipitates Ordered precipitates Burgers vector multiplies and dissociation become more complex in precipitates Modulus effects Dislocations have different line energy in precipitates Bowing round precipitates. (Orowan mechanism) Dislocation curvature increases on bypassing precipitates
Long range interactions
Microplasticity 5 Strengthening

Cutting interactions
S.G. Roberts

By pass interaction
5: 26

Menu

Hardening by precipitates
Hardening by strain fields

Strain fields around GuinierPreston zones of copper in aluminium

Small precipitates may be "coherent" with the matrix. All lattice planes, including slip planes are continuous as they pass through the precipitates; the change in lattice parameters in the precipitate causes each one to be surrounded by a (relatively long range) stress field. This interacts with the stress field of dislocations.

~ 2 G f
= misfit strain, f = volume fraction of ppts
Microplasticity 5 Strengthening S.G. Roberts 5: 27

Hardening by precipitates
Hardening by interaction with interfacial dislocations

Semicoherent
S` phase in CuA -Mg

` precipitates in Al - Cu

2 m Medium-sized precipitates, if too big to be fully coherent may be partially coherent" with the matrix. Most lattice planes, including slip planes are continuous as they pass through the precipitates; mistfit strain is taken up by interfacial dislocations. Stresses from these interact with the stress field of the mobile dislocations.
S.G. Roberts 5: 28

Incoherent
S phase Al2CuMg

Charai et al., Acta mater. 48 (2000) 2751

Microplasticity 5 Strengthening

Menu

Hardening by precipitates
Hardening by interaction with interfacial dislocations

xy

(generated using the dislnwall.xls spreadsheet)

Microplasticity 5 Strengthening

S.G. Roberts

5: 29

Dislocations and weak obstacles

Precipitates on dislocations slip plane

Dislocation line

Microplasticity 5 Strengthening

S.G. Roberts

5: 30

Menu

Dislocations and weak obstacles

Microplasticity 5 Strengthening

S.G. Roberts

5: 31

Dislocations and weak obstacles

Dislocations can cut through the precipitates (and their stress fields): but the bigger and more compound-like they are, the more difficult this is. Note that the precipitates get chopped up with increasing strain.

Microplasticity 5 Strengthening

S.G. Roberts

5: 32

Menu

Interfacial energy and morphology


coherent or semi-coherent ppt r Surface energy = 4r2

Ag G.P. zones in Cu

Surface energy (4r2 + rb )

Fobst =

dE E 1 = r dx 2b 2
5: 33

Microplasticity 5 Strengthening

S.G. Roberts

Stacking fault energy in precipitates

SF
coherent or semi-coherent ppt

SF

Stacking fault energy per unit area is lowered in ppt. Stacking fault widens; Dislocation stacking fault energy per unit length Dislocation elastic energy per unit length Overall dislocation energy in pptt

SF
coherent or semi-coherent ppt

SF

Stacking fault energy per unit area is increased in ppt. Stacking fault contracts; Dislocation stacking fault energy per unit length Dislocation elastic energy per unit length

Overall dislocation energy in ppt

Either way, the ppts harden the material.

Microplasticity 5 Strengthening

S.G. Roberts

5: 34

Menu

Ordered precipitates
Ni

Ni

0.2 m

Al Ni,Fe

Superalloys are the classic example Key component of the microstructure is precipitates of (Ni, Fe)3Al: `. A modern superalloy might be 60 - 85% ` Nickel is effectively a glue holding the `together: flow is restricted by Hall-Petch effect in narrow channels.

Microplasticity 5 Strengthening

S.G. Roberts

5: 35

Ordered precipitates

APB

coherent ordered ppt

1)

Single matrix dislocations will reverse the ordering, producing an Anti-phase boundary. Matrix dislocations must pair up to go through. Energy of the dislocation pair is increased by the APB.

2) 3)

Ordered structure means higher flow stress in precipitate. APB may be more stable on non-glide plane dislocation can flip into non-glide configuration.

Microplasticity 5 Strengthening

S.G. Roberts

5: 36

Menu

Ordered precipitates

Anti-phase boundary on (100) Stacking fault on (111)

Yield stress (MPa)


600 20% ` 500 40% 400 300 200 100 0 0 200 400 600 800 1000 1200 60% 80% 100% `

Ni / Ni3Al

Temperature (C) Transition of APB from {111} glide plane to {100} non-glide plane is thermally activated
Microplasticity 5 Strengthening

The yield stress of `increases with increasing temperature (up to about 700C).
S.G. Roberts 5: 37

Localised Obstacles
Fobst Localised Obstacles no long-range forces on dislocations

Fobst 2 Gb2 R Fobst

Gb 2 Fobst = 2 cos 2 = Gb 2 cos


Obstacle spacing =

and

R=

Gb 2

When dislocations break away from obstacles, Fobst is as high as the obstacle can extert, and is a critical (minimum) value c. .b = Fobst

Gb 2 cos c b Gb = cos c

Microplasticity 5 Strengthening

Obstacle strength is characterised by this critical angle for breakaway, c

S.G. Roberts

5: 38

Menu

Strong and Weak local obstacles


2c

b R

Weak obstacles:

c / 2

2c
b R Strong obstacles:

c 0

Microplasticity 5 Strengthening

S.G. Roberts

5: 39

Strengthening from local obstacles


What is the obstacle spacing? D A: If dislocation frees from one obstacle, then goes to same R, does not encounter another obstacle; continues to unzip tends towards B B: If dislocation frees from one obstacle, then goes to same R, doesnt quite encounter another obstacle - tends towards C . C,D: Dislocation already hitting multiple obstacles: c wont be reached; tends towards A. What is the critical condition? : mean particle spacing; A etc: effective particle spacings for these dislocations
Microplasticity 5 Strengthening S.G. Roberts 5: 40

(Weak obstacles: c / 2)

D C C

Menu

Strengthening from weak local obstacles


When dislocation unpins from one obstacle, it can move forward until its curvature rises again to that before unpinning as it encounters another obstacle. In Steady state: is constant (= flow stress) R is constant A* = area swept per obstacle passed = 2 A* = (area of segment, chord 2 ) 2.(area of segment, chord ) A*

A * = 2 = 2 = 3 Gb

(2 )3 3 3 2 = 12R 12R 2R
=

b but

Gb cos c
3 Gb (cos c )2

(cos c )

Microplasticity 5 Strengthening

S.G. Roberts

5: 41

Strong obstacles - hard precipitates


Precipitates are very effective blocks to dislocation motion, if: They are ordered intermetallic compounds strong directional bonding (e.g. CuAl2 in Al alloys, Ni3Al in Ni-based superalloys, Fe7Mo6 in maraging steels) They are [near] covalent compounds: e.g. Fe3C, WC, AlN in steels, MoSi2 in Al alloys They have a distinct interface with the matrix metal They have a strong associated stress field (from partial coherency or simple misfit) They are large, and thus difficult to cut through. However, Iarger means fewer and more widely spaced.

Microplasticity 5 Strengthening

S.G. Roberts

5: 42

Menu

Dislocations and strong obstacles

Precipitates on dislocations slip plane

Dislocation line

Microplasticity 5 Strengthening

S.G. Roberts

5: 43

Dislocations and strong obstacles

Microplasticity 5 Strengthening

S.G. Roberts

5: 44

Menu

Dislocations and strong obstacles

The bigger the spacing between the obstacles, the easier it is for dislocations to squeeze through the gaps. Each bypass event leaves a dislocation behind narrowing the gaps. Affects work-hardening rate.
Microplasticity 5 Strengthening S.G. Roberts 5: 45

Precipitation hardening strong obstacles


If the the obstacles do not allow dislocations to pass, breakaway angle approaches zero. b R Loop shape can appraoach to a halfcircle; radius of dislocation curvature is a minimum for this obstacle spacing.

.b = Gb2 / 2R = Gb2 / or if we only take component of in direction of b: = Gb /

(m) Typically (Steels), G = ~ 90 GPa 0.1 b = ~ 0.25 nm 1 10


Microplasticity 5 Strengthening S.G. Roberts

(MPa) 2250 225 22.5 2.25 feasible? useful Not a useful strength increase
5: 46

0.01

Menu

Precipitate dislocation interactions in age-hardening


0.05 m 0.5 m 2 m 10 m

Yield Stress
(MPa) 400 coherent G.P. zones semi-coherent semi-coherent

200

Cu in solution

interfacial dislns, hard ppts

Bowing
incoherent

Coherency Stress

Solute

ppts locally block dislocations


"base level" - pure Al (~50hrs at 190C)

Ageing Time

(scale varies with ageing temperature)


Microplasticity 5 Strengthening S.G. Roberts 5: 47

Stopping dislocations Grain boundaries


z z y x y x

Slip planes in both grains have common line in g.b. (may happen at twin boundaries) Burgers vector lies in g.b. plane. Still may be problems if dislocation is dissociated.

z y x

x z

b 1

b2

Most general case - most likely case ! Direct transmission of dislocations almost impossible.

Microplasticity 5 Strengthening

S.G. Roberts

5: 48

Menu

Stopping dislocations Grain boundaries


Pile up in Cu 4.5% Al. Every dislocation in the pile up exerts stress back on the source - tending to stop it operating. Slip must also be transmitted from one grain to the next . This is assisted by a stress concentration at the head of the pile-up.

Shear Stress

app

dislocation source dislocation source


ba ses ck-stres

app

blocked dislocation

Microplasticity 5 Strengthening

S.G. Roberts

5: 49

Stopping dislocations Grain boundaries


app

Stress felt by the possible new dislocation source is proportional to the stress at the head of the pile up, g.b.:

g.b. = Ndis dis


Ndis is the number of dislocations in the pile-up; dis is the stress each "passes on" to the next one in the array.

dislocation source

d
app

g.b.

dis = app - frict


frict is the minimum stress to move dislocations within the grain. The number of dislocations in the pile up is:

For slip to be passed from one grain to the next, the source must be activated at some critical source. Applied stress needed:

* app

Ndis = (/Gb) d` dis


where d' is equal to d, or d/2 - exact value not important. so:

* 2 source = frict + d

The Hall-Petch equation.

g.b. = d (app frict)2

y = i + k d
S.G. Roberts

1 2

Microplasticity 5 Strengthening

5: 50

Menu

Hall-Petch Effect
Lower Yield Stress (MPa) 400

Effect of grain size on yield strength of mild steel. Note: range of grain sizes typical for an engineering alloy
(martensites may have interfaces spaced by only a few 10s of nm) (in pearlite, interfaces may be a few m apart).

300

200

100 1mm 0 2 100m 4 6 8 10m 10

~3x range in yield strength over this grain size range variation of strength with temperature relatively high i: due to strengthening mechanisms other than grain size.

d-1/2 (mm-1/2)

Microplasticity 5 Strengthening

S.G. Roberts

5: 51

Steels
0.15%C

10m 50m Pearlite Iron-Carbon martensites: many interfaces interiors of laths heavily twinned or dislocated high and strongly varying local stresses around each locked-in C between laths between colonies of different orientation All these give very high resistance to dislocation motion.

1-10m Pearlite (Fe/Fe3C): Flow in Fe lamallae restricted by Hall-Petch Flow in Fe3C lamallae restricted by its high yield stress

Microplasticity 5 Strengthening

S.G. Roberts

5: 52

Menu

Work Hardening
nin rde a H ork =W
stic Pla

ate gr

iron = 20% 1012

Sl o

pe

Stress

109

Disln. density (cm-2)

Elastic

= 0%

106

Strain
Most dislocations are not mobile they are just obstacles to the few that are.
Microplasticity 5 Strengthening S.G. Roberts

stress

1 disln. spacing

= (disln. density)
5: 53

Simualtions of interactions between dislocations

When the misorientation between the two dislocation glide planes is high and the applied stress is high, the two interacting lengths are short.

When the misorientation between the two dislocation glide planes is small and the applied stress is small, the dislocations react over a longer length and form junctions which are strong obstacles to plastic flow.

http://zig.onera.fr/%7Edevincre/DisGallery/index.html
Microplasticity 5 Strengthening S.G. Roberts 5: 54

Menu

Dislocation debris formation and Forest interactions

Menu

Formation of a small dipolar loop by double cross-slip (centre) of a mobile dislocation (cyan). The cross-slipped portions are shown in pink.

Interaction of one single dislocation line with the forest. The specimen is oriented for slip on a single slip system (cyan). Dislocations on other systems are below the CRSS, and act as local obstacles to slip.

http://zig.onera.fr/%7Edevincre/DisGallery/index.html
Microplasticity 5 Strengthening S.G. Roberts 5: 55

Dislocation dynamics simulation of work-hardening

Deformation of a fcc single crystal (Cu) of linear dimension 15 m. The stress tensile axis is [100], the imposed strain rate is 50 per second and the plastic strain reached at the end of this sequence is 0.1%.

http://zig.onera.fr/%7Edevincre/DisGallery/index.html
Microplasticity 5 Strengthening S.G. Roberts 5: 56

Menu

Dislocation dynamics simulation of work-hardening

Stress (MPa)

Strain (%) Long range (elastic) interactions only Short range (jogs, locks) interactions only

Copper work hardened to 20% strain

Materials Science and Engineering A 309-310, (2001), 211-219

Microplasticity 5 Strengthening

S.G. Roberts

5: 57

Work-hardening: single- vs. poly- crystals


Applied tensile stress , tensile strain .

Stress (MPa) 120 100 80 60 40 20 0

Pure Aluminium

For single crystals, on slip system with resolved shear stress , shear strain :

Polycrystal

= cos cos =
At Yield:

y = M y
d = 1 d M

For polycrystals: y = M y Single Crystal

where M is averaged over all grains.


0.8 1

0.2

0.4 0.6 Strain (%)

d d = M2 d d
For ccp metals, the Taylor factor, M 3.1

Microplasticity 5 Strengthening

S.G. Roberts

5: 58

Menu

Workhardening and ductility


A A B C D B C D

: e ru T
E

,
e

, , s ring e e n ngi

Stress

Critical point at C is when

d = d

Strain

This is where the workhardening rate is no longer high enough to offset increase in stress in an incipient neck. Workhardening rate controls ductility.

Microplasticity 5 Strengthening

S.G. Roberts

5: 59

Ideal microstructure for strength ?


Cant use a non-metal as base of alloy
could get high yield strength this way but toughness would be far too low. effectiveness very limited for materials with yield stress increased close to UTS by other means reduces toughness (by reducing work-hardening capacity) will anneal out at moderate temperatures relatively limited (but useful) strength improvement (a few hundred MPa) fine grain sizes no good at high temperatures grain size will grow grain boundaries are fast diffusion paths creep strength will be poor in bcc, interstitial solutes give a limited but useful increase in strength (+ 50 to 80 MPa) (most steels are well over the solubility limit for C anyway) can quench in excess C, N into solution in martensite. (up to + 1000 MPa) but this form of martensite is very brittle (all dislocations are locked) substitutional solutes can give moderate strength increases (a few hundred MPa) limited by solubility solutes with big misfits will have low solubilities need to achieve a very fine dispersion of very hard precipitates (up to + 1000 MPa) or a very fine dispersion of weaker precipitates with strong associated strain fields (not quite so effective)

Can use work-hardening to improve strength, but:

Can use fine grain sizes to improve strength (and toughness!)


Can use solution strengthening

Can use precipitation hardening


Microplasticity 5 Strengthening

S.G. Roberts

5: 60

Menu

Ideal Microstructures
Use precipitation hardening
need as fine a dispersion as possible of particles as hard as possible. Solutes may be present anyway for a variety of reasons: to scavenge undesirable impurities (e.g. Mn, Si, Al in steels) to slow down diffusional reactions and thus promote martensite to confer corrosion resistance (e.g. Cr in steels) to react with other elements present and produce precipitates Generally a good idea, as toughness is improved

Topped up with solution strengthening

Why not have a fine grain size anyway ?

(Can we make spatial variations: e.g. surface different from bulk?) Extra constraints if high temperature use is likely:
base material must have a high melting point (!) any precipitates must be stable (there is a Tdynamic driving force to reduce number and increase size) reduce driving force (interface energy) slow down coarsening (reduce diffusion rate) fine grain size may not be good grain boundaries are fast diffusion paths so can get high creep rates grain size will tend to grow anyway beware (or use!) further reactions between elements present.

Microplasticity 5 Strengthening

S.G. Roberts

5: 61

What makes a metal suitable as a base for useful alloys ?


Readily available & cheap (bearing in mind the application) Cubic crystal structure
large number of slip systems for good ductility allows strong solution hardening allows use of precipitation reactions to form strong alloys precipitation of these gives strong hardening can cycle through phase change to produce physically complex microstructures can use partitioning of solutes between the phases to produce chemically complex microstructures can get martensitic (quenched) microstructures for some applications, anyway. first step in base for high temperature alloys either in itself, or via a strong, tenacious oxide film poisonous, explosive, inflammable, radioactive.

Readily dissolves a wide range of substitutional and interstial solutes Solubility of some solutes varies strongly with temperature Forms hard compounds with a range of elements Undergoes an allotropic phase change (e.g. Fe, Ti)

Low density

High melting point Chemically fairly stable Not hazardous

Microplasticity 5 Strengthening

S.G. Roberts

5: 62

Menu

The big 3
Iron
Available? Cheap ? ( / tonne 1997) M.P. (C ) Density (kg m-3) Allotropy Plentiful Extraction easy with C. ~150 1538 7874 bcc fcc bcc with rising T good solvent for many elements. Solubilty depends on allotropic form. can be strong and affected by allotropic changes forms compounds with many elements Oxidises steadily. Can be improved by solutes. No.

Aluminium
Useful ores (bauxite) widespread Needs cheap electricity to extract. ~1000 660 2700 fcc only moderate solvent for many elements strongly for many elements forms compounds with many elements Oxidises readily to form stable thin film. No. S.G. Roberts

Copper
Useful ores fairly rare Ores tend to be thin. Refining needed. ~1500 1084 8920 fcc only

Dissolves things ?

good solvent for many elements not strongly for most elements forms compounds with many elements Fairly inert. No. 5: 63

T-Variable solubilty Compounds Stability Hazardous ?

Microplasticity 5 Strengthening

The other 3
Nickel
Available? Cheap ? ( / tonne 1997) M.P. (C ) Density (kg m-3) Allotropy Thin ores in Canada and New Caledonia. Complicated, energy intensive extraction. ~4700 1455 8908 fcc only good solvent for many elements. can be strong forms compounds with many elements Fairly inert. No.

Titanium
Useful ores not plentiful Needs cheap electricity to extract. ~10000 1668 4507 cph bcc with rising T good solvent for many elements. Solubilty depends on allotropic form. can be strong and affected by allotropic changes forms compounds with many elements Oxidises readily to form stable thin film. No. S.G. Roberts

Magnesium
Useful ores fairly common, also in seawater. Needs electrolysis. ~1500 650 1738 cph only moderate solvent for some elements can be strong forms compounds with some elements Oxidises readily to form fairly stable film. Can burn. No. 5: 64

Dissolves things ?

T-Variable solubilty Compounds Stability Hazardous ?

Microplasticity 5 Strengthening

Menu

Microplasticity (2nd year MSoM / MEM) Steve Roberts Michaelmas term Question sheet 1 Concepts
Most of the answers to these questions are (or should be) short, but to answer the questions properly you will have to think about, use and extend material in the lectures, not just reproduce it! You may also have to look up some basic data. 1) Basics. Give a brief (max. 100 words, ideally less, one or two diagrams) definition of the following: a. A dislocation b. A slip plane c. A (glissile) dislocation loop d. A force on a dislocation e. A stacking fault formed by shear 2) More basics. Give a brief (max. 100 words, one or two diagrams) explanation of the following: a. How do we know dislocations really exist? b. How does dislocation glide give rise to plastic flow? c. Why can dislocations in some types of pure material glide at fairly low stresses? d. Why cant dislocations in other types of pure material glide at low stresses? e. What might make two dislocations attract each other? 3) Nodes. Assume that all three dislocations are pure edge in the diagram below:
b1 b3 b1 1 b2 2 3

a. b. c.

Sketch an atomic plane diagram around dislocations 2 and 3 to demonstrate that b1 = b2 + b3. If the line vector of dislocation 2 were redefined to be in the opposite sense, how would this affect: (i) the definition of b2; (ii) the summing of the Burgers vectors? Why would a node of such a type be unlikely to exist in reality?

4) Loops. Consider the dislocation loop below; what will the crystal look just after the has loop emerged:

b l b b b b l b

a. b.

Only in the middle of the left and right faces; Only in the middle of the near and far faces.

Menu

5) Loops. Is it possible for a dislocation loop to exist for which all parts of the loop are: a. Pure screw dislocations? b. Pure edge dislocations? In each case, either sketch the configuration of the loop or explain why it cannot exist. 6) Forces. Use the same methodology as for the derivation of the glide force on a dislocation to give an equation for the climb force on an edge dislocation in a crystal subjected to a tensile or compressive stress. 7) Forces / FSRH convention. On the diagram below, showing screw, mixed and looped dislocations, the line vectors are not indicated. Indicate the correct sense of the line vectors to give glide in the direction shown with the indicated shear stress (darker green indicates shear stress directionon top to the crystal).

8) Peach-Koehler Formula. ( F = (b) l ). Show that: a. For an edge dislocation with b in the x-direction and l in the z-direction, xy produces a glide force (in which direction?) and xx produces a climb force (in which direction?), but other stress components produce no force on the dislocation. For a screw dislocation with b and l in the x-direction, xy and xz produce glide forces (in which directions?), but other stress components produce no force on the dislocation. For an edge dislocation with b in the [110] direction and l in the [110] direction, xz produces a glide force (how large, and in which direction?), but xy produces no glide force.

b. c.

9) Dislocation energies and interaction stresses. Download the spreadsheet dislnpair.xls from my website at http://users.ox.ac.uk/~roberts/sgrgroup/index.htm. Use this to plot the energy of a pair of like (or unlike) dislocations at various separations on the same slip plane. Show that the energy of the system varies in the expected way with separation, giving rise to a force between the dislocations that varies inversely with their separation. Experiment with different configurations (dislocations separated in the direction of b, normal to b, moving on parallel planes, etc.) (NB the energy is given in arbitrary units, and only sums up the energy within the area plotted on the graphs. Dislocations close to the edge of the plot area will lose energy). 10) Dislocation interaction stresses. Consider two parallel edge dislocations with identical Burgers vector b on glide planes separated vertically by 10b. a. Plot the force acting on one of these dislocations as a function of horizontal separation ranging from 50b to +50b. b. What equilibrium separations are possible? c. What equilibrium separations are possible if the dislocations have parallel Burgers vectors of opposite sign? d. Compare the way these stresses vary with position with a plot of dislocation energy vs mutual position for this configuration, using the spreadsheet introduced in Q9.

Menu

11) Image forces. a. Pure copper has a minimum shear stress to move dislocations of ~10 MPa. If a thin foil of copper is examined in a TEM, how thick a specimen is needed to have a reasonable chance that it contains any dislocations? b. If the foil were of a single crystal of magnesium (HCP), how could you align the foil to prevent the dislocations leaving it due to image forces? c. What forces will a dislocation feel if it is close to a surface covered in a thick layer of oxide, which has a much higher modulus than the bulk metal? 12) Peierls stress. Explain in qualitative terms (no equations!) why dislocations with wide cores need less applied stress to move them. 13) Thompsons tetrahedron: . First make your tetrahedron.. a. To move the dislocations so as to form the Lomer-Cottrell lock as illustrated in the lecture, in what direction would one have to apply (i) a compressive stress (ii) a tensile stress ? b. There are three other possibilities similar to the Lomer-Cottrell lock for combinations of positive and negative b1 and b2. Use the Thompsons tetrahedron notation to follow the course of these reactions, and comment on the stability or otherwise of the reaction products. c. If the stacking fault energy were so high as to effectively prevent dissociation, what possible reactions could occur ? Is the most stable product more or less stable than for the most stable dissociated product? Is it mobile (glissile) or immobile (sessile)? 14) Frank loop conversion and prismatic loops: a. Use a side view of the {111} planes, labelling the ABC stacking, to show what happens to the stacking during the conversion of a Frank loop to a prismatic loop by a Shockley partial (3 drawings will be needed: before, half-way and after). b. What are the energetics of the conversion? c. Explain qualitatively why the conversion can only happen in the loop is above a certain critical size. (A quantitative calculation about this will be found on Q-sheet 2) d. If the prismatic loop is subjected to a high enough stress to make the dislocation glide, what will it do ? 15) Dislocations in bcc metals: a. For a dislocation with b = a [111], list all the possible slip planes. b. On a drawing of one or more bcc unit cells, draw in one possible slip plane of each type ({110}, {112} and {123}) for b = a [111]. c. Explain why slip lines in bcc metals are wavy (why arent they in ccp metals?) d. Which type of dislocation (edge or screw) controls the stress-strain response of a pure bcc metal? Why ? 16) Dislocations in ordered intermetallics: TiAl is an up & coming lightweight structural material. It has the L10 crystal structure (tetragonal, c/a =1.02). It slips on {111}<110> systems. a. Sketch the atomic arrangement on the slip plane. b. What kinds of dislocations and associated stacking faults might exist in this structure? 17) Dislocations in ionic materials: a. The lecture handout shows the lowest energy atomic arrangement at the core of an {110}<110>edge dislocation in the rock-salt structure. Sketch likely atomic positions for 2 or 3 intermediate positions as the dislocation glides to the next lowest energy position. b. How many {110}<110> slip systems are there ? How many are likely to be independent? What effect will this have on the deformation behaviour of (say) MgO?

Menu

18) Dislocations in covalent materials: a. A dissociated dislocation loop in silicon is shown in the lecture notes. What will be different about it if a similar loop is formed in gallium arsenide? (Hint: try drawing a section of the loop as seen edge on down <110>, labelling the atoms Ga and As carefully.) b. Silicon is (relatively!) ductile only above about 700C. What might the lowest temperatures for some ductility be in (i) germanium and (ii) diamond? c. The lecture notes show dislocation velocity / stress data for silicon. How might these have been obtained?

Menu

Microplasticity (2nd year MSoM / MEM) Steve Roberts Michaelmas term Strength of materials: Concepts
Most of the answers to these questions are (or should be) short, but to answer the questions properly you will have to think about, use and extend material in the lectures, not just reproduce it! You may also have to look up some basic data. 1) Yield stress and bonding. At roughly what temperature would you expect: a. Diamond to have a hardness of 20 GPa. b. Nickel to have a yield stress of 100 MPa. c. Molybdenum to have a yield stress of 250 MPa. (Diamond: room temperature hardness, 45 GPa. For metals, hardness = ~ 3x yield stress. Other relevant information can be found at www.webelements.com and in the lecture handouts.) 2) Samples of a well-annealed pure cubic-close packed metal are taken to 70% of its melting point and cooled rapidly by quenching. Tensile tests give yield strength values that decrease over several days, reaching a stable value a little below that for un-heat-treated samples. Explain what is happening, and what hardening mechanisms operate. 3) Diffuse obstacles. Describe and explain the differences in the ways dislocations interact with Weak diffuse obstacles and Strong diffuse obstacles. 4) Modulus effects: Explain why a material is hardened by solute atoms that produce a local change in modulus, irrespective of whether the modulus is increased or decreased. 5) Strain ageing: a. b. c. Tensile testing can be done in either load control (load is increased steadily, extension is measured) or displacement control (extension is increased steadily, load is measured). Which mode should be used to investigate strain ageing, and why? What is happening in the near-horizontal part of the 20c stress-strain curve for a plain carbon steel shown below? As temperature rises, stress-strain curves in iron-carbon alloys pass can become serrated (see below, 100C - 200C). Explain why this happens, and why the serrations disappear as the temperature rises further (250C).

Menu

6) Coherency stress hardening. Starting with aluminium with 4 wt% copper dissolved as a random solid solution, plot the expected hardening due to coherency stresses as the solutes form Guiner-Preston zones. 7) Local Obstacles. Describe and explain the diffenences in the ways dislocations interact with a. Diffuse obstacles and local obstacles b. Weak local obstacles and Strong local obstacles 8) Precipitation hardening. Describe and explain the differences in the ways dislocations interact with coherent precipitates, semi-coherent precipitates and incoherent precipitates. 9) Precipitation hardening. In Al-Cu alloys,, what interactions between dislocations and precipitates contribute to the hardening effects produced by a. b. c. d. Guinier-Preston zones `` precipitates ``precipitates precipitates

10) Work-hardening: Compare and contrast the likely work-hardening behaviour of polycrystalline specimens of: a. b. c. d. e. f. g. magnesium aluminium aluminium 4% Cu, quenched and aged at 180C for 10 hours. aluminium 4% Cu, quenched and aged at 180C for 500 hours. iron iron 0.2% carbon, normalised iron 0.2% carbon, quenched and tempered

11) Alloy design for strength: The nickel-based superalloy MAR M200 has the following composition: 9% Cr, 10% Co, 1.5% Ti, 5.5% Al, 0.15% C, 0.05% Zr, 0.015% B, 10% W,

Menu

2.5% Ta, 1.5% Hf, balance Ni. Describe the rle(s) of the underlined elements in strengthening the alloy.

Menu

Microplasticity (2nd year MSoM / MEM) Steve Roberts Michaelmas term Question sheet 2 Problems

1) Pure aluminium has a Youngs modulus of 70 GPa, a Poissons ratio of 0.24, a lattice parameter of 0.404 nm and a critical resolved shear stress to move dislocations of 10 MPa.

(a) What is diameter of the smallest dislocation loop, with Burgers vector in the plane of the loop, that can exist in aluminium?

(b)For such a loop of twice this minimum diameter, what level of resolved shear stress would be needed: (i) (ii) to make it expand to make it contract?

For each of b(i) and b(ii), what will happen once dislocation motion starts if the stress is kept at the same level?

You can try to solve part (a) either roughly, by assuming that the line tension of the dislocation line in a loop, radius R, is the same as that for an infinite straight dislocation, or in a more sophisticated way, by taking: E = [Gb2 / (4)]. ln(R/b). . ( = 1 or (1-) use either)

The latter approach may need use of numerical methods.

Menu

2) For a pure edge dislocation, with line vector along the z-direction, show (by transformation of the stress field as expressed in Cartesian co-ordinates) that:

Gb sin 2( 1 ) r Gb cos r = 2( 1 ) r Gb sin zz = ( 1 ) r Gb( 1 + ) sin P= 3( 1 ) r rr = =


3) Show that the non-core line energy per unit length E of a dislocation with Burgers vector b inclined at an angle to the dislocation line is:

Gb 2 ( 1 cos 2 ) R E= ln r 4( 1 ) 0
(you may assume the standard results for pure screw and pure edge dislocations).

4) An ionic solid made up of equal amounts of two elements has the NaCl crystal structure. The solid melts at 850C. At room temperature it can deform on {001}<110> slip systems with a critical resolved shear stress of 1050 MPa, and on {110}<1 1 0> slip systems with a critical resolved shear stress of 120 MPa.

(a) Why does slip not occur on {111} planes in this crystal structure?

[3]

(b) Why do the critical resolved shear stresses for the {001}<110> and {110}<1 1 0> slip systems differ by so much? [3]

(c) At what value of applied stress will a single crystal of the ionic solid start to deform if it is loaded in compression: (i) (ii) along the [111] direction along the [100] direction? [4] [4]

Menu

(d) What is likely to happen if a polycrystalline specimen of the ionic solid is tested in compression to a plastic strain of 5%: (i) (ii) at room temperature at 650C? [3] [3] [EMS finals 2000]

5. (a) Two straight edge dislocations with parallel Burgers vectors lie on parallel slip planes 30 nm apart. If the shear modulus G is 60 GPa, the magnitude of the Burgers vector b is 0.3 nm, and the Poisson ratio is 0.3, calculate the minimum resolved shear stress on the slip plane required to force the dislocations to glide past each other. The following information may be of use: The stress field of a straight edge dislocation along the z-axis, with a Burgers vector in the x-direction, has non-zero components: [12]

xx =

Dy 3 x 2 + y 2

(x

+ y2

),

xy =

Dx x 2 y 2

(x

+y

2 2

), ),
)

yy =

Dy x 2 y 2

(x
(

+y

2 2

zz = xx + yy ,
where D =

Gb 2 p(1 n )

(b) Giving reasons, describe what would happen if: (i) (ii) one of the edge dislocations were replaced by a screw dislocation both dislocations were replaced by screw dislocations. [4] [4]

[MSoM finals 2000]

Menu

6) Two edge dislocations on parallel planes, with mutually perpendicular Burgers vectors of the same magnitude b, pass close to each other as shown:

Show that the maximum climb force exerted between the two dislocations is:
Fmax = Gb 2 , ( 1 )d

where d is the separation of their slip planes. At what points on the dislocations is this climb force exerted? At what points on the dislocation is the climb force zero? What will happen if the temperature is raised enough for climb to occur?

[NB: Error in previous versions, with a rogue factor of 4 in the maximum climb force. Thanks to RIT for pointing this out. SGR 23/10/07]

7) By resolving the partial dislocations into their screw and edge components, find the equilibrium stacking fault width in copper for (a) a dissociated screw dislocation (b) a dissociated edge dislocation. How do these results compare with the rough value of 3.7 nm derived in the lectures? For copper, G = 48 GPa, = 70 mJ m-2, b = 0.256 nm.

Menu

8) [An expanded variant of the question set in finals in 2005]

y x

a ) The diagram shows an edge dislocation (with Burgers vector in the x direction and line vector along z) close to two substitutional solute atoms in positions A and B.

In of the following two cases, outline the mechanism of interaction (if any) between the dislocation and the solute atom at A, and the effect on the materials flow stress:

(i)

the solute atoms are smaller than the matrix atoms, but do not change the matrixs elastic moduli;

(ii)

the solute atoms are the same size as the matrix atoms, but locally reduce the matrixs elastic moduli.

b) How would the interactions for cases (i) and (ii) above change if the solute atom were at position B rather than position A?

c) Unobtanium (Ut) is a cubic-close packed metal with atomic diameter d = 0.537 nm and Youngs modulus 180 GPa. The following solutes are available:

Solute (X)

Atomic diameter (nm)

Youngs Modulus of Ut-10% X (GPa) 180 172 172

Solubility Limit in Ut (at% X) 10 35 18

Dullium (Du) Petitbonum (Pe) Chewingum (Cg)

0.472 0.537 0.541

Menu

Pure unobtanium has yield stress of 25 MPa. A solid solution of Ut 10%Du has a yield stress of 70 MPa. A solid solution of Ut 30%Pe has a yield stress of 36 MPa. (all the above materials have the same grain size).

What is the maximum increase in yield stress for a material of this grain size that could be produced by adding chewingum in solid solution to unobtanium?

Menu

Microplasticity Past Finals (General paper) Questions


(NB the course has changed somewhat over the years. Nonetheless, even the earlier questions are good exercises in the topics covered in the SGR 2007/8 variant of this course) 2007 1. a) Give brief explanations for two of the following experimental observations: (i) Slip in many body-centred cubic metals does not follow Schmids law. (ii) The room temperature ductility of copper is increased by the addition of 30wt% zinc. (iii) The addition of a few tens of parts-per-million of Cr2O3 impurity to MgO considerably increases its flow stress. (iv) The flow stress of the ordered intermetallic compound Ni3Al increases from ~80MPa at room temperature to ~350Mpa at 600C. [6] b) A sample of gold is quenched to room temperature from 90% of its melting point, and then examined by transmission electron microscopy. Small discs are seen, with contrast indicating that each is a stacking fault bounded by a dislocation. The specimen is examined several weeks later, and many of the discs seen previously now appear as dislocation loops with no stacking fault. Explain these observations. [6] c) Careful analysis of the specimen observed in (b) showed that the dislocation loops that had lost their stacking faults were all greater than 48nm in diameter. Explain why dislocation loops smaller than this are stable against loss of their stacking faults, and hence calculate a value for the stacking fault energy of gold. [8] [Gold: shear modulus = 27GPa; lattice parameter = 408pm]

2.

Menu

The micrograph shows a cross-section of an indentation made with a spherical indenter in a specimen of Fe-0.02%C; the section was made normal to the indented surface and cuts mid-way through the indentation impression. The specimen was annealed at 700C for 40 minutes, then polished and lightly etched. A series of hardness tests was made along the line A-B. Results are shown in the graph below.
500

Hardness (MPa)

450 400 350 300 250 0

10

12

14

16

18

Distance along line (mm)

a) Give an outline explanation for the hardness results. [4] b) The grain size at a position on the line 9mm from position A was measured as 150 m and the grain size 3 mm from position A was 1500 m. Estimate the grain size at position A. [8] c) Sets of specimens of an Al-4%Cu material were heat-treated by being held at 550C for 30 minutes, and quenched into water. They were then aged at 180C for either 10 hours or for 50 hours. Tensile test results after ageing were: Material Yield stress (MPa) Al-4% Cu (10 hours) 270 Al-4% Cu (50 hours) 425

Ultimate Tensile stress (MPa) 420 475

Strain at fracture (%) 30 12

Explain the different behaviour of these different sets of specimens in terms of the strengthening and work-hardening mechanisms operating. [8]

Menu

2006 1 a) Explain why interstitial solutes such as carbon and nitrogen can give substantial strengthening if dissolved in body centred cubic metals, but give a lower degree of strengthening if dissolved in cubic close packed metals. [4] b) Describe how you would expect the yield strength of a metal to vary with grain size. [2] c) Describe how you would expect the yield strength of a metal to vary with precipitate size, if the precipitates are incoherent and formed by reaction between the matrix metal and a solute that has negligible room temperature solubility. [4] d) The table below gives results of microstructural examination and tensile testing on an alloy aged for the given times. The alloy contains intragranular precipitates as a result of the aging. Aging time (hr) 2 10 20 50 75 100 500 1000 Grain size (m) 25 43 54 74 83 93 159 200 Precipitate size (nm) 13 22 27 37 41 46 79 100 Precipitate type coherent semi- coherent semi- coherent incoherent incoherent incoherent incoherent incoherent Yield Strength (MPa) 231 269 315 430 A 367 266 235

Assuming the effects of grain size and of precipitates on yield strength are simply additive, calculate the unknown yield strength A, and the contribution of each type of strengthening. Give your answer to 3 significant figures. [10]

Menu

2. a) Explain briefly why: i) screw dislocations in body centred cubic metals cross-slip more easily than those in cubic close packed metals ii) in a body centred cubic metal, screw dislocations have lower mobility than edge dislocations. 2 x [3] b) Two parallel edge dislocations, of Burgers vector b, lie on parallel glide planes that are separated by a fixed distance y, as shown.

i)

ii)

iii)

Sketch a graph showing the variation of interaction glide force, F, between the two dislocations for -5y < x < 5y. State, with reasons, which positions of equilibrium are stable and which are unstable. (Note equations for the stress field from an edge dislocation given below). [8] Explain what happens to F if the sign of the Burgers vector of the upper dislocation is reversed. [3] What would be the effect on F of replacing the upper edge dislocation with a screw dislocation? Why? [3]

Stress field components at position x, y from an edge dislocation situated at x = 0, y = 0 with line direction along z and Burgers vector b along x. G is the shear modulus and is Poissons ratio.

Menu

Menu

2005 1. The diagram below shows an edge dislocation (with Burgers vector in the x direction and line vector perpendicular to the page) close to a substitutional solute atom in either position A or position B.

a) In each of the following two cases, outline the mechanism of interaction (if any) between the dislocation and a solute atom at A, and the effect on the stress required to move the dislocation past the atom by glide: i) the solute atom is smaller than the matrix atoms, but does not change the elastic constants of the matrix; ii) the solute atom is the same size as the matrix atoms, but locally reduces the stiffness of the matrix. [8] b) How would the interactions for cases (i) and (ii) above change if the solute atom were at position B rather than position A? [5] c) The following solutes are available, neither of which changes the elastic modulus significantly:

The Um-rich ends of the corresponding Um-X binary phase diagrams are as follows:

Menu

Which solute is able to give the greatest amount of solid solution strengthening at room temperature, and by what factor does this amount exceed that obtainable using the other solute (answer to 2 significant figures)? [5] d) Briefly discuss one additional strengthening mechanism that may be obtainable in UmDu alloys, and suggest a suitable alloy composition with which to investigate this possibility. [2]

Menu

2004 1. a) Precipitates can increase the flow stress of a material, even when dislocations can pass through the precipitates. Outline three distinct mechanisms by which this can occur. [6] b) Explain how the strength of interaction between precipitates and dislocations can be characterised by a critical angle c and show that the stress required to move a dislocation through an array of such obstacles with spacing is given by:

Gb cos c

where G is the shear modulus of the material and b is the magnitude of Burgers vector of the dislocation. [6] c) A material with G = 70GPa and b = 0.25 nm contains an array of coherent precipitates with the same lattice parameter as the matrix. At an early stage of ageing when the precipitate diameter is 20 nm and the mean particle spacing is 100 nm, the critical angle c = 35. (i) Calculate the flow stress inside the precipitates. [5] (ii) The precipitates coarsen with further ageing; calculate the maximum size of precipitates that can be cut by dislocations. [3] 2. a) Explain why dislocations in cubic close packed metals are often dissociated, while those in body centred metals are normally undissociated. [6] b) A transmission electron microscope is used to study dislocations in a pure single crystal cubic close packed metal with lattice parameter a = 0.388 nm, shear modulus G = 46GPa and Poissons ratio = 0.31. The plane of the thin foil specimen is (111). A dissociated screw dislocation with line vector along [110] on a (111) plane is observed to have a stacking fault width of 3.2 nm. By resolving the partial dislocations bounding the stacking fault into edge and screw components, calculate the stacking fault energy of the material. [10] c) The foil is subjected to a shear stress in its plane and parallel to the [121] direction. How will the dislocation in (b) respond? [4]

Menu

Note that the force per unit length F between two parallel dislocations of Burgers vector of magnitude b, separated by a distance d, in a material with shear modulus G is given by:

Gb 2 F= 2d (1 x )
where x = 0 for a screw dislocation and x = Poissons ratio for an edge dislocation.

Menu

2003 1. a) In each of the following cases, describe without the use of an equation how the energy of a pair of parallel dislocations on a single, common glide plane varies with separation and interpret the variation in energy in terms of forces on the dislocations if: i) the dislocations are identical and of edge character ii) one dislocation is pure edge and the other pure screw. [4] b) Two parallel screw dislocations of opposite sign are lying on parallel glide planes as shown in the diagram.. The glide planes are a distance d apart and the separation of the dislocations, as measured in the glide planes, is z. Derive an expression for the component of force in the glide plane on unit length of dislocation.

[8] At what values of z is the force in the glide plane (i) zero (ii) a maximum? [4] c) What will be the likely outcome if initially, x >> d and the glide component exceeds the frictional force, and if: i) the material has a close-packed hexagonal crystal structure with a high c/a ratio [2] ii) the material has a body-centred cubic crystal structure? [2] 2. a) Discuss the interactions between dislocations and solute strains, and the consequential effects on yield strength. Distinguish carefully the differences between interactions with edge dislocations and interactions with screw dislocations for certain point defects. [10]

Menu

b) A face-centred cubic metal has a shear modulus of 210 GPa and lattice parameter 352 pm. Various solutes are added to strengthen the material by solid solution and the following data pertain: Solute A E G J L Atomic radius, pm 124 125 143 145 91 Solubility limit, at% 13 19 4 3 0.08

Giving reasons for your choice; which alloying element is likely to give the greatest increase in strength? Justify any assumptions made. [10]

Menu

2002 1. a) Explain why dislocations in aluminium dissociate into partial dislocations. [5] b) In an intermetallic compound with a cubic structure, the dislocations dissociate into partials separated by an anti-phase boundary (APB) according to: a[110] a/2 [110] + a/2 [110] Calculate the separation of the partials when a screw dislocation dissociates if the APB energy is 40 mJm-2, the shear modulus of the crystal is 52 GPa and its lattice parameter is 0.4 nm. [3] c) With reference to the operative dislocation mechanisms, discuss the variation with temperature of the flow stress of: i pure iron ii pure nickel iii Ni3Al [6] 5. a) What is meant by work hardening in a ductile, crystalline material? [3] b) Write briefly on the work hardening processes involved when glide dislocations interact with the following: (i) other dislocations (ii) incoherent particles [7] c) A single crystal of a fcc alloy with stress axis along [123] is deformed in tension. The yield stress is 400 MPa and electron microscopy shows that the crystal contains a uniform dispersion of incoherent spherical particles 20 nm diameter. The shear modulus of the crystal is 70 GPa and the lattice parameter, 0.36 nm; calculate the spacing of the particles. [10]

Menu

2001 1. a) Show that in a work-hardened polycrystalline sample, the yield stress, y is given by:

y = i + k

where, i is the lattice resistance stress, k is a constant and is the dislocation density. [8] b) Discuss what happens (i) when two screw dislocations, each with line direction perpendicular to the slip plane of the other, intersect on perpendicular slip planes; [4] (ii) when two 600 mixed dislocations, each with line direction perpendicular to the slip plane of the other, intersect on perpendicular slip planes. [5] c) Which, if either, of the interactions in (b) can give rise to hardening? [3] 2. a) Explain carefully what is meant by force on a dislocation when it is placed in uniform stress field. Does a vacancy experience is such a force in uniform stress field? Explain your answer. [6] b) Consider a dislocation loop on a particular slip plane. Let be the shear stress resolved on the slip plane in the direction of the Burgers vector. Show that the force exerted on the dislocation is everywhere normal to the dislocation line and that the force per unit length dislocation is b where b is the magnitude of the Burgers vector. [9] c) A straight dislocation has a Burgers vector (bx, by, bz) = (b,0,0). The dislocation is subjected to the uniaxial stress xx. For each of the following line directions state whether a force is exerted on the dislocation, and if so whether it would move by glide or climb: i) (0,0,1) ii) (1,0,0) [5]

Menu

2000 1. a) Describe four mechanisms by which interactions between dislocations and particles control the yield stress of an age-hardening alloy. [4 x 3] b) Show that the resolved shear stress required to sustain a dislocation with a radius of curvature R is of order Gb/R, where b is the magnitude of the Burgers vector and G is the shear modulus. Hence explain the decrease of the yield stress during overageing. [8] 2. a) Two straight edge dislocations with parallel Burgers vectors lie on parallel slip planes 30 nm apart. If the shear modulus G is 60 GPa, the magnitude of the Burgers vector b is 0.3 nm, and the Poisson ratio is 0.3, calculate the minimum resolved shear stress on the slip plane required to force the dislocations to glide past each other. [12] The following information may be of use: The stress field of a straight edge dislocation along the z-axis, with a Burgers vector in the x-direction, has non-zero components:

xx =

Dy 3 x 2 + y 2

zz = v xx + yy

(x

+y

2 2

+y Gb , where D = . 2(1 v )
2

),

xy

Dx x 2 y 2

(x

2 2

),

yy

Dy x 2 y 2

(x

+y

2 2

),

b) Giving reasons, describe what would happen if i) one of the edge dislocations were replaced by a screw dislocation [4] ii) both dislocations were replaced by screw dislocations. [4]

Menu

1999 1. a) What are the similarities and differences between the stress fields around screw and edge dislocations? A screw dislocation lying along the z-axis has a stress component y = Gb / 2x a distance x from its core, where G is the shear modulus and b the Burgers vector. By evaluating the work done in the xy plane forming the dislocation, show that the elastic energy per unit length between radii R and r0 is given by

Gb R = ln 4 r0
b) Explain how this leads to a dislocation line tension T = Gb2, where is a constant. c) A certain metal has a dislocation friction stress F. Find the minimum size of dislocation loop that can be stable if G = 50 GPa, b = 3 x 10-10 m and F = 107 Pa. 2. a) Two edge dislocations with Burgers vector b of opposite sign move freely on parallel planes. If the position of one relative to the other is given in polar coordinates (r, ), find the value of at which stable equilibrium will occur. [10] b) What will happen if screw dislocations replace the edge dislocations? [4] c) If an f.c.c. metal is alloyed with a small percentage of a substitutional element, what differences would this make to the behaviour of (i) edge and (ii) screw dislocations? [6] [In polar coordinates, the components of force between two edge dislocations of Burgers vector b in a material of shear modulus G and Poisson ratio are

Fr =

Gb 2 Gb 2 sin 2 ; F = 2r (1 v ) 2r (1 v )

where the angle is measured from a direction parallel to the Burgers vector.]

Menu

1998 1. a) Show that the theoretical shear strength of th of a crystal is given by:

th

1 a 2

where is the shear modulus, a the spacing of the crystal lattice and l the periodicity of the lattice in the sliding direction. [8] b) Consider a single crystal cube of side R subjected to a shear stress . Show that, for this cube to deform by a dislocation mechanism, the stress must reach a value of approximately:

b R

where b is the magnitude of the Burgers vector of the dislocation. State any assumptions you make. [9] c) Show that there is a lower limit on the size of the cube below which the cube will not deform by a dislocation mechanism. [3]

Menu

1997 1. a) Write down the components of stress that tend to make: i) an edge dislocation glide, ii) an edge dislocation climb and, iii) a screw dislocation cross slip. [6] b) Discuss three mechanisms by which particles may interact with dislocations to produce strengthening of a metallic alloy. Under what conditions does each mechanism apply? [10] c) Calculate the mean particle separation in an overaged Al-Cu alloy when the yield stress is 500MPa. [4] (You may assume that the shear modulus is 40 GPa and the lattice parameter is 0.356 nm.) 2. a) Derive the displacement, strain and stress fields of a screw dislocation under isotropic elastic conditions. Hence, obtain an expression for the energy per unit length of the dislocation. [8] b) Screw dislocations in b.c.c. iron tend not to interact with substitutional atoms such as sulphur. However, they do interact with carbon interstitials. Explain these observations. [7] c) Would you expect the observations in (b) above to apply to edge dislocations in b.c.c. iron? If so, why? [5]

Menu

1996 1. Describe the dislocation reactions which occur during the formation of a LomerCottrell lock in an f.c.c. crystal undergoing plastic deformation. [8] Discuss the transition from stage I to stage II in the work hardening behaviour of an f.c.c. metal single crystal during uniaxial tensile deformation and explain why the onset of the transition is dependent on the initial orientation of the tensile axis [8] Why is stage II not reached before fracture occurs for some h.c.p. metal single crystals? [4] 2. What are the essential assumptions in the Friedel statistical model of strengthening caused by small obstacles? [4] Assuming a distribution of obstacles as a square array on the slip plane which is subjected to a shear stress t, explain qualitatively how the spacing between obstacles pinning a dislocation depends on the obstacle strength, F. [6] For spherical particles explain the physical mechanisms causing the force F for two of the following: i) Coherency (strain) hardening. ii) Order hardening. iii) Stacking fault hardening. iv) Modulus hardening. [5+5] 3. What are the physical origins of the interactions between point defects and dislocations in metals and in ionic crystals? [4] Explain how these interactions give rise to the following: i) Cottrell atmospheres. [4] ii) iii) iv) The yield drop. [4] Strain ageing. [4] The Portevin-Le Chatelier effect. [4]

Menu

4. a) Derive the displacement, strain and stress fields of a screw dislocation under isotropic elastic conditions. Hence, obtain an expression for the energy per unit length of the dislocation. [8] b) Screw dislocations in b.c.c. iron tend not to interact with substitutional atoms such as sulphur. However, they do interact with carbon interstitials. Explain these observations. [7] c) Would you expect the observations in (b) above to apply to edge dislocations in b.c.c. iron? If so, why? [5]

Menu

1995 1. Write short notes on four of the following: a) the strain-rate dependence of the flow-stress in b.c.c. metals; b) stage I work-hardening in copper; c) Cottrell atmospheres; d) twinning in hexagonal metals; e) the strength of cold worked -brass after low-temperature annealing. [5] x 4 2. a) Some metals contain small bubbles of gas or voids and the tensile stress is known to increase with increasing density of bubbles within the grain. If dislocations cut through the bubbles discuss the strengthening mechanism and derive an expression for the interaction force. [8] b) A bubble density of 2 x 1010 mm-3 is found to increase the yield stress by 100 MPa; what bubble density is needed to achieve an increase of 200 MPa if Friedels weak pointobstacle criterion is applicable? [12] 3. Explain how vacancies in a fcc metal can condense to form Frank dislocation loops, and how the Frank loop can be converted into a prismatic loop with unit Burgers vector. [8] Find the critical size of a circular Frank loop in copper at which conversion to a prismatic loop is energetically favourable. [10] How may such loops be introduced into metals in practice? [2] [Assume that the elastic strain energy per unit length of dislocation line is given by

b 2 r ln where is the shear modulus (45 GPa for copper), r is the radius of the loop 4 b
and b the magnitude of the Burgers vector. The lattice parameter of copper is 360 pm and the stacking fault energy is 40 mJ m-2].

Menu

1994 1. Write notes on four of the following: a) line tension of a dislocation b) stacking faults in an fcc structure c) Frank-Read sources d) dislocation climb e) Lomer-Cottrell dislocations. 2. Compare and contrast the Peierls stress in body-centred cubic and in close-packed hexagonal metals. Discuss mechanical twinning in these structures. 3. Derive an expression for the energy per unit length of a screw dislocation lying along the axis of a long, hollow cylinder of elastically isotropic material with inner and outer radii r and R respectively. State the equivalent expression for an edge dislocation, and hence derive an expression for the energy per unit length of a mixed-character dislocation with Burgers vector b at an angle to the line direction. For a dislocation in a real crystal, discuss the factors determining the magnitudes of the parameters r and R in practice. When a sample of cold-worked copper is recrystallised, 1 MJ m-3 of energy is released; estimate the dislocation density in the sample. (For copper, the shear modulus is 40 GPa and the lattice parameter is 360 pm.)

Menu

1993 1. Sketch tensile stress-strain curves to illustrate the effect of the following dislocation/solute atom interactions: i) dislocation locking, ii) friction stress effects. In (i) explain why an initial decrease in load occurs at the onset of plasticity. In (ii) discuss the factors which may determine the force exerted upon a dislocation by an individual solute atom, illustrating your answer with examples from copper-based solid solutions. 2. Explain Thompsons notation for describing the slip planes and Burgers vectors of whole and partial dislocations in the fcc structure. Thompsons tetrahedron is chosen so that face is (111), face is (111) and BC = [110]. A dislocation moving on plane with line direction BC and Burgers vector BD encounters a parallel dislocation moving on plane with Burgers vector AB. Discuss the interactions which occur when the dislocations meet along the line of intersection of the slip planes (a) when they are undissociated, and (b) when they are dissociated into Shockley partial dislocations separated by a ribbon of intrinsic stacking fault. 3. Draw a graph illustrating the work hardening behaviour of a single crystal of an fcc metal deformed in tension. Describe in outline how the form of the graph may be accounted for in terms of the dislocation interactions taking place. Discuss the effect of the following upon the form of the graph: i) crystal orientation, ii) stacking-fault energy, iii) temperature of deformation.

Menu

1991 1. Explain what is meant by (a) Schmids critical resolved shear-stress law, and the Schmid factor. In what circumstances might you expect Schmids law to be obeyed? The structure of corundum, Al2O3, is hexagonal with oxygen atoms in hexagonal close packing and the aluminium atoms filling two thirds of the octahedral interstices. Slip occurs on the basal plane in a <10 1 0 > direction. How does this slip system compare with that in zinc? Give reasons for any differences. The a[10 1 0 ] unit dislocation in corundum dissociates into partials

a[10 1 0]

a [10 1 0] + a [10 1 0] + a [10 1 0] 3 3 3

Calculate the separation of the partials when the dislocation dissociating is a screw. [For corundum, the stacking-fault energy is 0.32 J m-2, the lattice parameter, a, is 0.476 nm, and the shear modulus is 156 GPa.]

Menu

1990 1. Compare the typical forms of the tensile stress-strain curves of single crystals and polycrystals of fcc metals and account briefly for the main features of the curves. Discuss the effect of grain size on the stress-strain curve of the polycrystal. Fluorite (fcc Bravais lattice) slips only on the systems {001} <1 1 0 > at low temperatures. Discuss the probable behaviour in a tensile test of: a) a single crystal with [100] tensile axis b) a single crystal with [110] tensile axis c) polycrystal

Menu

CM MOK cmmok128@yahoo.com

Online reading materials: Ductile Iron Tutorial http://www.ductile.org/didata/Section2/2intro.htm Alloys and Stainless Steels http://www.roymech.co.uk/Useful_Tables/Matter/Alloy_Steels.html Metallurgical Engineering http://www.sut.ac.th/Engineering/Metal/course.html Matters http://www.matter.org.uk/steelmatter/metallurgy/default.htm

Appendix D. Page 116 of 116.

You might also like