Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Solidification and Solid-State Transformations of Metals and Alloys
Solidification and Solid-State Transformations of Metals and Alloys
Solidification and Solid-State Transformations of Metals and Alloys
Ebook697 pages6 hours

Solidification and Solid-State Transformations of Metals and Alloys

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Solidification and Solid-State Transformations of Metals and Alloys describes solidification and the industrial problems presented when manufacturing structural parts by casting, or semi-products for forging, in order to obtain large, flat or specifically shaped parts. Solidification follows the nucleation and growth model, which will also be applied in solid-state transformations, such as those taking place because of changes in solubility and allotropy or changes produced by recrystallization. It also explains the heat treatments that, through controlled heating, holding and cooling, allow the metals to have specific structures and properties. It also describes the correct interpretation of phase diagrams so the reader can comprehend the behaviour of iron, aluminium, copper, lead, tin, nickel, titanium, etc. and the alloys between them or with other metallic or metalloid elements.

This book can be used by graduate and undergraduate students, as well as physicists, chemists and engineers who wish to study the subject of Metallic Materials and Physical Metallurgy, specifically industrial applications where casting of metals and alloys, as well as heat treatments are relevant to the quality assurance of manufacturing processes. It will be especially useful for readers with little to no knowledge on the subject, and who are looking for a book that addresses the fundamentals of manufacturing, treatment and properties of metals and alloys.

  • Uses theoretical formulas to obtain realistic data from industrial operations
  • Includes detailed explanations of chemical, physical and thermodynamic phenomena to allow for a more accessible approach that will appeal to a wider audience
  • Utilizes micrographs to illustrate and demonstrate different solidification and transformation processes
LanguageEnglish
Release dateMar 16, 2017
ISBN9780128126080
Solidification and Solid-State Transformations of Metals and Alloys
Author

Maria Jose Quintana Hernandez

María José Quintana, PhD has an European Doctorate in Science and Technology of Materials from the University of Oviedo (Spain), and is a professor and researcher in the Materials group in the Panamerican University (Mexico). Dr. Quintana works on subjects of molding, thermomechanical treatments of steels, superplasticity of metals, microscopy and mechanical properties of materials, among others.

Related to Solidification and Solid-State Transformations of Metals and Alloys

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Solidification and Solid-State Transformations of Metals and Alloys

Rating: 5 out of 5 stars
5/5

2 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Solidification and Solid-State Transformations of Metals and Alloys - Maria Jose Quintana Hernandez

    text.

    Chapter 1

    Solidification of Metals

    Abstract

    The characteristics of metallic atoms are explained together with their behavior from the gaseous to the solid states, from a thermodynamic point of view. Crystalline systems and calculations related to packaging, volume change, and lattice during solidification are analyzed. The phenomenon of solidification of a metal is explained using thermodynamics of bonding during cooling and its resulting crystalline structure and grain microstructure are presented in a general fashion. Finally, the stiffness of the atomic bond is analyzed as a function of temperature and the relationship between melting point, heat treatments, and deformation temperatures are explained.

    Keywords

    Metallic bond; Crystalline systems; Melting temperature; Packaging; Solidification free energy

    1.1 Metals

    The main difference between metals and nonmetals lies in the number of electrons in the external orbit of the atoms: metals have a lower number of electrons which are easily released in order to form complete and stable orbits.

    Table 1.1 shows the electronegativity of metals, calculated as the energy necessary for an atom to attract an electron using 3.98 as the base value (Pauling criterion) assigned to Fluorine (the most electronegative element). Corresponding values for some nonmetals are: Boron (2.04), Phosphorous (2.19), Hydrogen (2.20), Carbon (2.55), Sulfur (2.58), Iodine (2.66), Bromine (2.96), Nitrogen (3.04), Chlorine (3.16), and Oxygen (3.44).

    Table 1.1

    Electronegativity (in Increasing Order) of Metals According to the Pauling Criteria

    When the difference in electronegativity between two metals is considerable, the bond between them will be of the ionic type. Other possible atomic or molecular bonds are covalent, coordinate covalent, polar covalent, and metallic.

    In the case of a metallic solid, each atom loses peripheral electrons to an electron cloud and this, due to its electronegativity, brings positively charged atoms together. This bond between atoms of the metallic crystal or grain is only observed in metals and therefore called metallic bond. A characteristic of this structure is the anonymity of the bond between atoms where each atom is not specifically connected to any other atom, which is in contrast from other types of chemical bonds. Another difference is the mobility of the cloud formed by valence electrons; this easiness in their displacement results in high thermal and electric conductivities.

    Another property of metals related to their bond is the amount of deformation before rupture, which compared to nonmetallic materials, such as ceramics, glasses, ionic solids, etc., is considerably larger. Once deformation reaches the yield stress value, the metallic bond does not break as the atoms can slide over each other, which is translated into plastic deformation (at atomic, microscopic, and macroscopic levels). On the other hand when nonmetallic materials reach elastic separation energy between the atoms of a molecule, the bond breaks and the material fractures.

    The metallic bond also results in the ability of metals to form alloys either by substitution or insertion of foreign atoms: due to the anonymity of the bond in the crystalline lattice (solvent) some of its atoms can be substituted by other metals (solute). The substitution of one kind of atom by another can be complete if atomic radius, electronegativity, valence (number of electrons offered to the electronic cloud), and crystalline lattice are similar to the ones of the solvent; for example, copper and nickel can form alloys in any proportion, such as Monel (75% Ni and 25% Cu), Constantan (45% Ni and 55% Cu), or Cupronickels (75% Cu and 25% Ni).

    1.2 From the Gaseous State to the Crystalline State

    Metals, like all other elements, will be in a solid, liquid, or gaseous state depending on the combination of temperature and pressure. As an example, Fig. 1.1 shows the equilibrium temperature-time curve during cooling for aluminum (Al) at a pressure of 1 atm: pure Al is solid below 660°C, liquid between 660°C and 2450°C, and gas above 2450°C. The gasified metal inside a closed environment behaves according to the kinetic model of a group of atoms with Brownian movement: atoms move in a disordered fashion, gravity has no influence in these movements and collisions follow, almost exactly, the laws for conservation of momentum and energy. Due to this high kinetic energy, it is practically impossible for the formation of any group of atoms.

    Fig. 1.1 Cooling curve of a metal from gas to solid.

    The total kinetic energy of the atoms can be evaluated by the pressure of the gas against the walls, while the temperature of the gas (in Kelvin degrees) measures the mean kinetic energy, though not all of them have the same kinetic value: some will have higher or lower values compared to the mean one and the true value of the kinetic energy for each atom, and therefore its speed, follows at any instant in time the Maxwell-Boltzmann law (mean speed and its standard deviation are directly proportional to the temperature of the gas). The most probable speed, according to the Maxwell-Boltzmann statistic distribution, at a temperature T is determined by:

       (1.1)

    where m is the mass of the atom and k ). Fig. 1.1 also shows the statistical distribution and the dn/N ratio of atoms with a certain speed v:

       (1.2)

    EXERCISE 1.1

    For Al gas at a temperature of 3000 K, calculate the most probable speed for one of its atoms.

    Solution

    As Eq. (1.1) requires the mass of the atom, it can be obtained by dividing the atomic weight by the Avogadro number (NA):

    And using this value in Eq. (1.1):

    As an example, at T1g, the curve follows the shape of the Maxwell-Boltzmann distribution with the most probable speed value v1. When temperature decreases to T2g (gaseous state), the new curve shows a smaller dispersion in the speed of the atoms and the mean speed v2 is lower than v1: a drop in speed produces the mean displacement of the atoms at T2g, lower than at T1g. As temperature decreases, a threshold temperature is reached where the attraction forces between the atoms of the gas start to balance the repulsion effect of kinetic energy. These attraction forces (caused by the metallic bond) are the reason why atoms tend to form groups and eventually the characteristic lattice of their crystalline state. Furthermore, metallic atoms are subjected to two types of interatomic forces:

    • Repulsion forces: caused by the kinetic energy of each atom which promotes dispersion (influenced by the collisions with the walls of the container and with each other).

    • Attraction forces: caused by the tendency to form metallic bonds.

    If temperature is lowered, attraction forces start to manifest and, therefore, begin to neutralize the repulsion ones (e.g., for Al at 2450°C). The metal changes from the gaseous state to the liquid one, by a mechanism of nonstop formation and breaking of bonds. In the liquid state, at temperatures close to evaporation (or overheated liquids), atoms have a freedom of movement similar to the one they have at the gaseous state. Each atom is surrounded by a defined number of anonymous neighbor atoms, without the tendency to adopt a crystalline structure, and with a behavior similar to a compressed gas. Yet, the difference with a gaseous state is that groups of atoms are formed, even though their lifetimes are short and their bonds easily disappear.

    If the temperature of the liquid metal drops even further, the groups last longer until they become permanent, and solid state characteristics are reached: the liquid presents a certain crystalline structure (i.e., the vibration of particles around certain equilibrium positions) even though it is a transitory state and can easily be dissolved. The atoms can move short distances and the thermal movement of each of them is constituted by vibrations with a defined mean frequency but with continuously changing direction and amplitude. The true values for the kinetic energy at any moment are also distributed according to the Maxwell-Boltzmann law. As a specific temperature is reached (660°C for Al), the liquid metal turns to solid, in the form of crystalline aggregates; each of them formed by atoms that vibrate and are positioned at locations of the typical crystalline lattice of each metal.

    If a crystalline solid is cooled even further, the amount of vibration also diminishes and the atoms become closer to one another until reaching a new equilibrium between attraction forces (metallic bond) and repulsion ones (addition of vibration energy and electrostatic rejection between the positive charges of the nuclei). The equilibrium distance between the atoms at the interior of each grain in solid state can increase if energy is added to the metal by heat, or can decrease if energy is removed by cooling: the dimensions of the crystalline lattice are reduced, which at a macroscopic scale results in contraction of the material.

    1.3 Crystalline Systems for Metals

    When metallic sample is properly polished, etched with an adequate chemical reagent, and observed through an optical microscope, a cellular structure similar to the one presented in Fig. 1.2 is revealed, where each cell is a crystalline grain or crystal. If X-ray diffraction techniques are used, it can be proved that inside each grain, the metallic atoms are positioned in a regular manner; repeated systematically in the three spatial directions. The fundamental principle of crystal structures is that the atoms are set in space either on the points of a Bravais lattice or in some fixed relation to those points. Therefore, the fundamental stacking unit is known as unit cell. Though it has an ideal or abstract character, crystals are formed by groups of these cells where an atom may be shared by several of these ideally sectioned units.

    Fig. 1.2 Optical micrographs of equiaxed grain structure in a metal: (A) hot-rolled and tensile deformed at high temperatures for shipbuilding steel, (B) as-extruded tellurium-copper alloy for electrical applications, and (C) hot-rolled and recrystallized titanium alloy for heat exchanger applications.

    Most materials used in industrial applications have, at room temperature, the following crystalline systems: face-centered cubic (e.g., γ-Fe, Al, Cu, Pb, Ni, Ag, Pt, and Au), body-centered cubic (e.g., α-Fe, V, Cr, Nb, Mo, Ta, and W), or hexagonal (e.g., Mg, Zn, Zr, Ti, Be, and Co). Fig. 1.3 shows the unit cell of the face-centered cubic system, where the centers of the atoms are located in the corners of the cube and at the center of the faces (each cell contains the mass of four atoms). Fig. 1.4 shows the unit cell for the body-centered cubic (bcc) system (with a total of two atoms per unit cell), with atoms located at the corners of the cube and at the center of the cell.

    Fig. 1.3 Face-centered cubic unit cell.

    Fig. 1.4 Body-centered cubic unit cell.

    Furthermore, the compact hexagonal system has as unit cell a prism with a rhombus as a base (of sides a and a) and height c; each corner is associated with two atoms: one occupies the (0,0,0) position and the other occupies the (2/3a,1/3a,c/2) position as indicated in Fig. 1.5. Joining two rhombic prisms and two semiprisms together results in what is erroneously, but commonly, known as a hexagonal unit cell. Fig. 1.5 also shows the base of the hexagonal cell: a hexagon formed by three rhombs. As the height of the cell is c, the centers of the atoms are located at the corners of the hexagonal prism, at the centers of the hexagonal bases, and at the middle plane of the prism (in total, there are six atoms per unit cell).

    Fig. 1.5 Hexagonal compact packing unit cell.

    The hexagonal system behaves as a compact one when the atoms at the middle plane are tangent to the base atoms. In this case, each atom is tangent to 12 others: 6 of its own plane, 3 of the upper plane, and 3 of the lower plane. If c is the height of the prism and a is the side of the hexagon (or the rhomb for the unit cell), the c/a ratio in the hexagonal compact system equals 1.633, which may be confirmed by the geometric considerations shown in Fig. 1.6, where

    Fig. 1.6 Compact condition in the hexagonal system.

       (1.3)

       (1.4)

       (1.5)

    Hexagonal compact packing (hcp) and face-centered cubic (fcc) systems are the only crystalline arrangements in which maximum tangency of an atom to the surrounding ones is achieved. Each atom is in contact with 12 others, so there is an analogy between both systems; in fcc, the compact packing of atoms corresponds to diagonal planes that are crystallographically known as (111); while in hcp, the compact packing of atoms corresponds to basal planes that are crystallographically known as (0001).

    Other metals adopt more complex structures. For example, tin crystallizes in the body-centered tetragonal system and uranium in the orthorhombic one. atoms).

    Table 1.2

    Crystalline System and Density of Metals

    Enjoying the preview?
    Page 1 of 1