You are on page 1of 17

Some Properties of Engineering Materials

Notes for ME 083 Dr. Stefan Zauscher Duke University, Pratt School of Engineering Department of Mechanical Engineering and Materials Science Spring Semester 2003

Contents
1 Introduction 2 Equilibrium Properties 2.1 2.2 2.3 2.4 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mechanical Properties (Chapter 6, Callister) . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thermal Properties (Chapter 19, Callister) . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electric and Magnetic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.1 2.4.2 Electric Properties (Chapter 18, Callister) . . . . . . . . . . . . . . . . . . . . . . . . . Magnetic Properties (Chapter 20, Callister) . . . . . . . . . . . . . . . . . . . . . . . . 1 1 1 2 5 5 5 7 8 8 8 8 9 9 9 9 10 10 10 10 12 13 14

3 Transport Properties 3.1 Heat Transfer (Chapter 19, Callister) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 3.1.2 3.2 3.3 3.4 3.2.1 3.3.1 3.4.1 Thermal Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thermal Diusivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Diusivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electrical Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Mass Transfer (Chapter 5, Callister) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Momentum Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Charge Transfer (Chapter 18, Callister) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4 Other Properties 4.1 4.2 4.3 Radiation properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chemical Properties (Chapter 17, Callister) . . . . . . . . . . . . . . . . . . . . . . . . . . . . Interaction Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5 References

List of Figures
1 Typical engineering stressstrain behavior to fracture, point F. The tensile strength is indicated at point M. The circular insets represent the geometry of the deformed specimen at various points along the curve (Fig. 6.11, Callister). 2 . . . . . . . . . . . . . . . . . . . . . . . 2 (a) The dependence of repulsive, attractive, and net force on interatomic separation for two isolated atoms. (b) The dependence of repulsive, attractive, and net potential energies on interatomic separation for two isolated atoms. (Fig. 2.8, Callister). . . . . . . . . . . . . . . . 3 4 Relationships between hardness and tensile strength for steel, brass, and cast iron. (Fig. 6.19, Callister). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (a) Plot of potential energy versus interatomic distance, demonstrating the increase in interatomic separation with rising temperature. With heating, the interatomic separation increases from r0 to r1 to r2 , and so on. (b) For a symmetric potential energyversusinteratomic distance curve, there is no increase in interatomic separation with temperature, i.e., r 0 = r1 = r2 . (Fig. 20.3, Callister). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 The spectrum of electromagnetic radiation, including wavelength ranges for the various colors in the visible spectrum. (Fig. 22.2, Callister). . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 6 4 3

Supplement for ME 83: Some Properties of Engineering Materials

Introduction

Many properties of materials can be described in terms of the relationship between a geometry-independent eld intensity (what one does to a material) and a geometry independent response (how the material responds). Examples of eld intensities are mechanical stress, temperature, and electric eld strength. Examples of response variables are mechanical strain, heat transfer per unit mass, and electric displacement. If the behavior is linear, the property can be described completely by two numbers the slope of the eld intensity vs. response curve, and the maximum value of eld intensity which the material can withstand. If the behavior is nonlinear, the entire eld intensity vs. response curve provides the most information, but specic details of this curve may be treated as specic properties (e.g., yield strength, ultimate tensile strength, and elongation from a stress-strain curve). Note that when the behavior is nonlinear, the slope must be dened in dierential terms. Other properties cannot be described quantitatively as relationships between responses and eld intensities. In such cases, behavior may be described qualitatively (e.g., attack by solvents), or materials may be rank-ordered according to some measure of their behavior under similar conditions (e.g., the galvanic series, which ranks corrosion tendency in sea water). Those properties which are most frequently encountered by engineers working with materials are listed below. The answer to the question, Do I have to memorize all those equations? is No, but you will be expected to know how every variable and property is dened and how to use them to solve simple engineering problems. For some of you, this expectation may mean memorizing equations; for others, alternative approaches to learning will suggest themselves.

Equilibrium Properties

The relationships describing the response of a material (change in size or shape, heat absorption, charge accumulation, ...) to an applied eld (gravity, temperature, electric eld, stress, magnetic eld,...) utilize a group of quantities known as material properties. If these properties do not vary with time, they are designated equilibrium properties.

2.1

Density

Density v can be described as a gravitational property; it is dened as mass per unit volume (Eq. 1), dm = v dV. The specic volume is dened as (Eq. 2), v= 1 . v (2) (1)

Densities of solid materials range from < 0.1 g/cm3 for light polymer foams to > 20 g/cm3 for the platinumgroup metals.

Supplement for ME 83: Some Properties of Engineering Materials

Figure 1: Typical engineering stressstrain behavior to fracture, point F. The tensile strength is indicated at point M. The circular insets represent the geometry of the deformed specimen at various points along the curve (Fig. 6.11, Callister).

2.2

Mechanical Properties (Chapter 6, Callister)


F . A

Stress is dened as the force, F , per unit area, A (Eq. 3), = (3)

If the cross-sectional area over which the stress is distributed is the initial area (A 0 ), rather than the instantaneous area A which changes with deformation, the stress is called engineering stress, e = F . A0 (4)

Strain is dened, in one-dimensional deformation, as the relative change in a materials dimension when compared to its original length, L0 , (Eq. 5)
L

=
L0

dL L = ln . L L0

(5)

If the strain is approximated by (Eq. 6), L (6) L0 the strain is called engineering strain. The elastic modulus E is the slope of the stress (eld intensity) e

strain

(response variable) curve for an elastic or elastic-plastic material (Eq. 7)(Fig. 1), d = Ed . (7)

Supplement for ME 83: Some Properties of Engineering Materials

Figure 2: (a) The dependence of repulsive, attractive, and net force on interatomic separation for two isolated atoms. (b) The dependence of repulsive, attractive, and net potential energies on interatomic separation for two isolated atoms. (Fig. 2.8, Callister). Elastic moduli range from 100 N/m2 (< 200 psi) for soft vinyl plastics to > 0.6 1012 N/m2 (93 Mpsi) for tungsten carbide. On an atomic scale, macroscopic elastic strain is manifested as small changes in the interatomic spacing, and stretching of interatomic bonds. For covalently bonded materials the elastic modulus is therefore a measure of the resistance to bond stretching from the equilibrium separation r 0 . This modulus is proportional to the slope of the interatomic forceseparation curve (Fig. 2) at the equilibrium spacing, E dF dr .
r0

(8)

The smaller the radius of curvature in the bond energy well, i.e., the steeper the walls of the potential energy well, the larger is the amount of energy to displace atoms from their equilibrium positions; the higher is the elastic modulus. If measured in shear, the modulus is called the shear modulus G (Eq. 9), xy = Gxy , (9)

where xy is the shear stress (force in x direction, applied to an area with its normal in the y direction) and xy is the shear strain (dx/dy ). Tensile strength of an elastic (brittle) material is the stress at which fracture occurs in tension. Tensile strengths vary from about 7 104 N/m2 (10 psi) for soft foam rubbers to nearly 3.5 109 N/m2 (5 105 psi)

Supplement for ME 83: Some Properties of Engineering Materials

Figure 3: Relationships between hardness and tensile strength for steel, brass, and cast iron. (Fig. 6.19, Callister). for asbestos bers. Plastic deformation involves a permanent change in shape and always results in a nonlinear stress-strain curve. The stress at which yielding begins in plastic deformation is called the yield strength; the maximum engineering stress the material can sustain in tension is called the ultimate tensile strength (UTS), and the maximum strain at tensile fracture is called the elongation. The area under the stress-strain curve is a measure of toughnessmechanical energy absorbed per unit volume V (Eq. 10), W =V d . (10)

Yield strengths of ductile materials vary from about 4 106 N/m2 (600 psi) for soft polyethylene foam to about 2.5e 109 N/m2 (350,000 psi) for austempered steels. Tensile strengths range up to 2.5 3.5 106 N/m2 (4-500,000 psi) for austempered and highly cold-drawn steels. Elongations vary from less than 1% for highly cold-drawn steels to > 800% for polyethylene lm. Hardness is a measure of resistance to penetration or scratching. Hardness may be described with respect to a rank-ordering scheme (Mohs scale), by the stress calculated for a xed indenter and known load (Brinnel, Knoop, DPH), or by the depth of penetration for a xed indenter and a known load (Rockwell). Hardness of materials range from less than 20 kg/mm2 (KHN) for talc and soft plastics to 7000 kg/mm2 for diamond. Figure 3 shows the relationship between hardness and tensile strength. Hardness tests are usually non-

Supplement for ME 83: Some Properties of Engineering Materials

destructive and much quicker and easier to perform than tensile tests, which makes them valuable tools in quality control and non-destructive materials testing. A few other mechanical properties are piezocaloric coecient (change in heat transfer per unit mass with stress), stress-corrosion (the enhancement of corrosion rate with stress), and piezoelectricity (change in dielectric displacement with stress).

2.3

Thermal Properties (Chapter 19, Callister)

The specic heat capacity c can be described as the relationship between heat transfer per unit mass Q/m (response) and temperature change T (eld intensity), in dierential form (Eq. 11), dQ = cmdT. (11)

Specic heats of solids at room temperature range from about 0.13 kJ/kg-K for uranium, gold, and the platinum metals to about 2-2.5 kJ/kg-K for a number of polymers. The melting point is a specic temperature, measured most precisely by determining the temperature at which a structure-dependent property like density or specic heat varies the most with temperature when the material undergoes a solid-liquid phase transformation. Melting points range from below room temperature for some elastomers to about 3410o C for Tungsten. The area under a curve of specic heat vs. temperature up to the melting point yields the energy per unit mass to melt a given material. Thermal expansion can be described as the relationship between strain (response) and temperature change T (eld intensity). The coecient of linear thermal expansion L is then dened by (Eq. 12), d = L dT. (12)

Coecients of linear thermal expansion for common engineering materials vary from less than 10 7 /K for thermal-shock-resistant ceramics such as Pyroceram (Corning Ware) to over 10 3 /K for some soft elastomers such as silicone rubber. Figure 4 shows the eect of temperature on interatomic position and shows that the thermal expansion coecient is directly related to the asymmetry and depth of the potential well. A few other thermal properties are pyroelectricity (change in dielectric displacement with temperature), Curie temperature (above which a ferromagnetic material becomes paramagnetic), and thermoelectricity (the change in current density produced by temperature at the junction of two conductors, the basis for temperature measurement by thermocouples).

2.4
2.4.1

Electric and Magnetic Properties


Electric Properties (Chapter 18, Callister)

For conductive properties see Section 3.4. Dielectric properties describe the electrical behavior of insulators. A material can be polarized and is able to hold charge; this polarization is most easily illustrated with a

Supplement for ME 83: Some Properties of Engineering Materials

Figure 4: (a) Plot of potential energy versus interatomic distance, demonstrating the increase in interatomic separation with rising temperature. With heating, the interatomic separation increases from r 0 to r1 to r2 , and so on. (b) For a symmetric potential energyversusinteratomic distance curve, there is no increase in interatomic separation with temperature, i.e., r0 = r1 = r2 . (Fig. 20.3, Callister). parallel plate capacitor. A capacitor can be envisioned as two, parallel conducting plates of equal area, separated by an insulator (dielectric). The electrical charge stored on either plate is directly proportional to the applied voltage (Eq. 13), Q = CV, (13)

where the proportionality constant, C , is dened as the capacitance [F = farad]. Capacitance depends on the geometrical properties of the capacitor and on the polarizability of the dielectric, capacitance is then dened as (Eq. 14), C= where
r r 0

A , d

(14)

is the relative dielectric constant,

is the permittivity of vacuum (8.85 1012 F/m), A is the area

of the capacitor, and d is the plate separation. The permittivity of a dielectric is the relationship between the electric eld strength E = dV /dx (voltage gradient) and electric displacement D = Q/A (charge per unit area of one electrode). For a linear dielectric, D = E, and for a ferroelectric (nonlinear, irreversible dielectric), dD =
(E ) dE.

(15)

(16)

The relative dielectric constant er is the most frequently quoted, and most easily measured, dielectric

Supplement for ME 83: Some Properties of Engineering Materials

property (Eq. 17),


r

=
0

(17)

Dielectric constants for electrical insulators vary from about 1.05 for foamed polystyrene to about 40 for glass-bonded micas. The value of electric eld strength at which breakdown occurs is called the dielectric strength. Values of dielectric strengths vary from about 2 kv/m for some silicone foams to > 80 kV/m for natural mica. Ferroelectrics are dielectrics which exhibit nonlinear, and irreversible polarization behavior. For a ferroelectric, the remanent polarization is the maximum polarization which such a material can maintain in zero eld; it is a rough measure of the maximum (saturation) polarization. A few other electrical properties are converse piezoelectricity (the change in strain with electric eld intensity), electrocaloric coecient (the change in heat transfer per unit mass with electric eld intensity), and the electro-optic coecient (change in birefringence with electrical eld intensity). 2.4.2 Magnetic Properties (Chapter 20, Callister)

The magnetic eld strength H (eld intensity variable) is dened in terms of an equivalent solenoid (Eq. 18), H= NI , l (18)

where N is the number of turns, I is the current through the coil, and l is the length of the coil. According to Faradays law of magnetic induction a search coil around a solenoid will generate a voltage whenever the magnetic eld cutting the coil changes. Therefore, V , the voltage induced in the coil is dened by the time-rate of change in the magnetic induction B (Eq. 19), V = 1 dB , A dt (19)

where A is the eective cross-sectional area of the coil. Linear (weak) magnetic materials can be described by the magnetic permeability, [H/m], dened by Equation 20, B = H. In analogy to the dielectric materials, the relative permeability is dened as (Eq. 21), r = , 0 (21) (20)

where 0 is the permeability of vacuum (1.257 106 H/m). If is positive (and small), the material is paramagnetic; if is negative (and small), the material is diamagnetic. Ferromagnetic materials are non-linear, with large positive values of , and permeability must be dened in dierential terms (eq. 22), dB = (H ) dH. (22)

Supplement for ME 83: Some Properties of Engineering Materials

Typically, ferromagnetic materials are described in terms of the relationship between the magnetization M and the magnetic eld strength H (Eqs. 23,24), M= and M = m H, (24) B H, 0 (23)

where m is called the magnetic susceptibility. If the curve of magnetic induction or magnetization vs. H is reversible, the material is a soft magneteasily magnetized, easily demagnetized. If B or M vs. H is irreversible, the material retains a considerable amount of magnetization at zero eld strength, and it is said to be a hard magnet. The magnetization retained at H = 0 is called the remanent magnetization. A couple of other magnetic properties are magnetostriction (change in strain with magnetic eld strength) and hall coecient (change in transverse current density with magnetic eld strength).

Transport Properties

Transport properties are used in relationships between eld intensities (thermal gradient, electric eld strength, concentration gradient, ...) and uxes, or ows, through the material (heat, charge, mass, ...). Properties are designated transport when the material is acting as a conductor.

3.1
3.1.1

Heat Transfer (Chapter 19, Callister)


Thermal Conductivity

Thermal conductivity, kT , can be described as the relationship between the heat ux Q/(At) [J/(s m 2 )] and the thermal gradient T /x [K/m] (Eq. 25), dT 1 dQ = kT , A dt dx or at steady-state, q = kT dT . dx (26) (25)

Values of thermal conductivity near room temperature range from about 0.2 10 4 J/s-m-K for common insulators (cork, plastic foams) to about 2.9 J/s-m-K for pure silver. 3.1.2 Thermal Diusivity

Thermal diusivity considers the nonsteady-state thermal conduction process that involves both changing temperature (specic heat) and heat transport (thermal conductivity). Thermal diusivity k/(c) can be

Supplement for ME 83: Some Properties of Engineering Materials

considered the relationship between the time-variation of temperature at a point and the spatial variation of the temperature at that same point (Eq. 27), k d2 T dT = , dt c dx2 (27)

derived by considering a dierential volume of material through which heat is passing. Part of the heat added to the material passes through the dierential volume (thermal conduction), and part of the heat goes into raising the temperature of the dierential volume (specic heat capacity).

3.2
3.2.1

Mass Transfer (Chapter 5, Callister)


Diusivity

Diusivity, Dm , also has a thermal origin (the thermal vibrations of the atoms), and can be described as the relationship between the mass ux of atoms m/(At) [kg/s-m2 ] and the concentration gradient C/x [kg/m4 ] of that atomic species in the material (Eq.28), dC 1 dm = Dm , A dt dx or at steady-state, J = Dm dC . dx (29) (28)

The minus sign indicates that diusion takes place from a high concentration to a low concentration. Diusivities vary from almost zero for incompatible elements to 108.7 m2 /s for hydrogen in iron.

3.3
3.3.1

Momentum Transfer
Viscosity

Viscosity, , is another transport property (i.e., it can be treated similarly to thermal conductivity and diusivity). In this case the ux is a transverse momentum ux, and the eld intensity is the shear-strain rate xy /t (Eq. 30), 1 dpy dxy = , Ax dt dt (30)

where py is the momentum mvy transverse to the x direction of uid ow and xy is the shear strain. At steady state we have F = dpy /dt and Equation 30 becomes, yx = where yx is the shear stress. dxy , dt (31)

Supplement for ME 83: Some Properties of Engineering Materials

10

3.4
3.4.1

Charge Transfer (Chapter 18, Callister)


Electrical Conductivity

Electrical conductivity can be described as the relationship between the electric ux q/(At) [C/m 2 ] (charge/unit time per cross-sectional area, or current density) and the electric eld strength E = V /x [V/m] (Eq. 32), 1 dq dVe = e , A dt dx current density Equation 32 becomes, dVe . (33) dx If E is constant, i.e., dVe /dx = Ve /l where l is the length of the conductor, then we recover Ohms law Je = e (Eq. 34), V = IR, by noting that Je = I/A and that e = l/(RA). (34) (32)

where e is the electrical conductivity, and its reciprocal e = 1/e is the electrical resistivity. For constant

4
4.1

Other Properties
Radiation properties

Two generally distinguishable forms of radiation exhibit signicant interactions with solids: electromagnetic (EM) radiation, e.g., radio waves, light, and ultraviolet radiation, and particle radiation, e.g., electrons, neutrons, alpha particles, etc. Very long wavelength EM radiation, e.g., electrical power transmission and long radio waves, are of low enough frequencies that their interactions with materials can be treated in quasistatic terms, e.g., by considering electrical conductivity, dielectric and magnetic susceptibilities in slowly varying elds. Note that radiation in this frequency range (102 1010 Hz) can produce internal heating in materials by electronic conduction in conductors, by dipole oscillation in dielectrics (the principle on which the microwave oven is based). Figure 5 shows the range of the EM spectrum. In the middle of the EM spectrum, infrared (IR), visible, and ultraviolet (UV) radiation typically are considered to interact with materials by being transmitted, reected or absorbed. Highly reective materials in the visible spectrum possess free electrons, to which a continuum of energy levels is available after excitation. Highly transparent materials possess only localized electrons, which together with their nucleiform induced dipoles as an EM wave passes. Bulk absorption of EM radiation occurs by a series of collisions, after each one of which the EM wave has a lower intensity I , so that intensity decreases with thickness x according to (Eq. 35), dI = dx, I where is called the linear absorption coecient. In integral form, the above equation is (Eq. 36), I = I0 exp x , (36) (35)

Supplement for ME 83: Some Properties of Engineering Materials

11

Figure 5: The spectrum of electromagnetic radiation, including wavelength ranges for the various colors in the visible spectrum. (Fig. 22.2, Callister). or (Eq 37), I = I0 expm x , where m = / is called the mass absorption coecient that is usually tabulated in the literature. In addition to bulk absorption, many discrete (in the vicinity of a specic frequency) absorption mechanisms exist: IR absorption by bond resonance in organic molecules (the basis for IR spectroscopy); light absorption by quantized d electrons (the basis for color in copper, gold, and many glasses); UV absorption by bond rupture in plastics and biological molecules (the mechanism of UV degradation and sun tans). In addition to the mass absorption coecient, the optical properties most often quoted in the literature are percent transmission (for a given thickness), index of refraction, and birefringence. The percent transmission could be derived from the mass absorption coecient, but it usually is determined experimentally; of most use are curves of percent transmission vs. wavelength from the IR region to the UV. The index of refraction n is simply the ratio of the speed of light in a vacuum to the speed of light in the material. Some materials, e.g., transparent plastics under stress, are birefringent; that is, they possess two dierent indices of refraction in dierent directions. The dierence between these two indices of refraction is called the birefringence and is responsible for the colors that transparent materials exhibit in polarized light. As the radiation frequency increases to the x-ray and gamma-ray regions of the EM spectrum, the probability of absorption by some type of discrete mechanism becomes much greater than that for bulk absorption. The mass absorption coecient varies as about the third power of the wavelength between (37)

Supplement for ME 83: Some Properties of Engineering Materials

12

wavelengths at which discrete absorption mechanisms operate, so high-frequency radiation experiences less bulk absorption than low-frequency radiation. Also, high-frequency EM radiation possesses more energy, according to Einsteins equation (Eq. 38), E = h, (38)

therefore high-frequency radiation has a higher probability of breaking atomic bonds and creating discrete imperfections. Electromagnetic radiation also can be scattered, the size of the scattering centers being related directly to the wavelength of the radiation. The absorption, transmission, and scattering of particles e.g., electrons, alpha particles, and ionized atoms, by a material are similar to the absorption and transmission of EM radiation. The only dierence is that particles have mass, therefore can be expected to exhibit higher absorption coecients (for the same energy) and to do more damage in the process of being absorbed.

4.2

Chemical Properties (Chapter 17, Callister)

Chemical properties describe those aspects of the behavior of materials which relate to atomic bonding and to rearrangements of atoms occasioned by a change in the chemical environment of a material. For example, the tendency of a material to oxidize is given by its free energy of oxide formation per oxygen atom, a thermodynamic chemical property. All metals except silver and gold have negative free energies of oxide formation, therefore will oxidize. All plastics will oxidize. Most glasses and ceramics already are oxides. Note that the rate of oxidation of any material will be governed by other factors, and is considered a kinetic chemical property, rather than a thermodynamic property. Corrosion is the electrochemical solution of a metals ions in an electrolyte. The tendency of a metals ions to go into solution can be measured by the potential of the half-cell reaction (Eq. 39), M e M e++ + 2e , (39)

compared to that of some standard electrode, usually a hydrogen electrode. Rank ordering metals by these potentials yields the Electromotive Force (EMF) Series. A more pragmatic approach is to rank metals and commercial alloys in a common electrolyte (usually sea water) according to their tendency to dissolve electrochemically, this results in the galvanic series. Solution involves the dissolving of a solute in a solvent. Most of the information on solvents for materials is qualitative, but the general rules are that non-polar liquids dissolve (or are absorbed by) nonpolar plastics; polar liquids dissolve (or are absorbed by) polar plastics; liquid metals are the best solvents for solid metals (phase diagrams can be used to predict extent of solution vs. temperature); molten glasses and ceramics are the best solvents for solid glasses and ceramics. For some materials, especially plastics and organic solvents, a solubility parameter has been measured; a liquid can be expected to be a good solvent for a solid if both have similar solubility parameters. The surface energy of a solid is another chemical property; it is considered due to the unsatised valence electrons at a surface. Since these electrons are not taking part in interatomic bonding, they are not at

Supplement for ME 83: Some Properties of Engineering Materials

13

their lowest possible energy levels, and this dierence in energy, divided by the area, is the surface energy . Metals have relatively high surface energies; oxides have lower surface energies; and plastics (especially nonpolar molecular solids) have the lowest surface energies. Surface energy determines how readily water or solder will wet a surface and how well an adhesive will stick to a surface. A surface will be wet by another material only if the wetting action lowers the total surface energy. Other chemical properties attempt to describe the tendency of two materials to react chemically, or for a material to react with its environment, producing new reaction products, e.g., suldation, eutectic melting, polymerization, and cross-linking.

4.3

Interaction Properties

A number of uncommon properties can be regarded as the change in one property when a second eld intensity is changed. Magnetoresistivity measures the change of electrical resistivity produced by a change in magnetic eld strength. Thermoresistivity measures the change in electrical resistivity produced by a change in temperature. Piezoresistivity measures the change in electrical resistivity produced by a change in stress. Thermoelasticity describes the change in elastic modulus with temperature. Radiation damage describes the structural changes produced by ionizing radiation-changes which aect almost all other properties. Many other interaction properties are possible some of which have not even been studied in sucient detail yet to accumulate much data.

Supplement for ME 83: Some Properties of Engineering Materials

14

References
1. W. D. Callister, Materials Science and Engineering and Introduction, 5th ed., Wiley (2000). 2. J. P. Schaer, A. Saxena, S. D. Antolovich, T,. H. Sanders, Jr., and S. B. Warner, The Science and Design of Engineering Materials, McGraw-Hill (2000). 3. J. F. Shackelford, Introduction to Materials Science for Engineers, 4th ed., Prentice Hall, (1996). 4. M. F. Ashby, and D. R. H. Jones, Engineering materials, vols. 1 and 2, Pergamon Press, (1981,1988). 5. W. G. Moatt, G. W. Pearsall, and J. Wul, The Structure and Properties of Materials, Vol. I, Structure, Wiley (1964). 6. J. H. Brophy, R. M. Rose, and J. Wul, The Structure and Properties of Materials, Vol. II, Thermodynamics of Structure, Wiley (1964). 7. H. W. Hayden, W. G. Moatt, and J. Wul, The Structure and Properties of Materials, Vol. III, Mechanical Behavior Wiley (1965). 8. R. M. Rose, L. A. Shepard, and J. Wul, The Structure and Properties of Materials, Vol. IV, Electronic Properties, Wiley (1966). 9. E. A. Mechtly, The International System of Units; Physical Constants and Conversion Factors, 2nd rev., Washington, Scientic and Technical Information O., National Aeronautics and Space Administration (1973).

10. F. Rodriguez, Principles of Polymer Systems, McGraw-Hill (1970).

You might also like