Uploaded by ce_amcastillo_82

asme materials

advertisement
Overview of Mechanical
Properties of Materials
for Engineers
Participant’s Workbook
CONTACT INFORMATION
ASME Headquarters
1-800-THE-ASME
ASME Professional Development
1-800-THE-ASME
Eastern Regional Office
8996 Burke Lake Road - Suite L102
Burke, VA 22015-1607
703-978-5000
800-221-5536
703-978-1157 (FAX)
Midwest Regional Office
1117 S. Milwaukee Ave.
Building B - Suite 13
Libertyville, IL 60048-5258
847-680-5493
800-628-6437
847-680-6012 (FAX)
Northeast Regional Office
326 Clock Tower Commons
Route 22
Brewster, NY 10509-9241
914-279-6200
800-628-5981
914-279-7765 (FAX)
International Regional Office
1-800-THE-ASME
Southern Regional Office
1950 Stemmons Freeway - Suite
5037C
Dallas, TX 75207-3109
214-746-4900
800-445-2388
214-746-4902 (FAX)
Western Regional Office
119-C Paul Drive
San Rafael, CA 94903-2022
415-499-1148
800-624-9002
415-499-1338 (FAX)
You can also find information on
these courses and all of ASME,
including ASME Professional
Development, the Vice President
of Professional Development,
and other contacts at the ASME
Web site...
http://www.asme.org
Overview of
Mechanical Properties of Materials
for Engineers
by
Professor Stephen Liu
Colorado School of Mines
Metallurgical and Materials Engineering
Golden CO 80401
Copyright © 1999 by
All Rights Reserved
TABLE OF CONTENTS
Part 1: PARTICIPANT NOTES……………………………………… 3
Part 2: BACKGROUND MATERIAL………………………………… 30
List of References
Introduction to Metallurgical Principles
Part 1
Participant Notes
3
Notes
Overview of Mechanical Properties
of Materials for Engineers
Prepared by:
Professor Stephen Liu
Ph.D., CEng, MIM, FAWS
Professor of Metallurgical Engineering
Colorado School of Mines
Golden, Colorado 80401
Notes
Overview of Mechanical Properties
of Materials for Engineers
• Objectives and Outline
– Elastic Behavior of Materials
– Plastic Behavior of Materials
– Impact and Fracture Toughness of Materials
– Fatigue of Materials
– Creep of Materials
– Development of Mechanical Properties
– Model Material Systems
– Summary
4
Notes
Elastic Behavior of Materials
Chart
Chartfor
for Elastic
Elastic
Design
Design
•Contours
•Contours show
show
Failure
FailureStrength
Strength
•Selection
•Selection of
ofMaterials
Materials
for:
for:
––Springs
Springs
––Elastic
ElasticHinges
Hinges
––Pivots
Pivots
––Elastic
Bearings
Elastic Bearings
––…
…
––Yield-beforeYield-beforeBuckling
BucklingDesign
Design
Ashby
Ashbyand
andJones
Jones1980
1980
Notes
Material
Testing
Smith
Smith1993
1993
• Tensile Testing
• Compression
Testing
• Fatigue Testing
• Fracture
Toughness Testing
• Specimen
Geometry
• Extensometer
5
Notes
Elastic Behavior of Materials
• Stress-Strain Definition
– Stress (Engineering Stress)
S=
F
A
e=
∆!
!o
– Strain (Engineering Strain)
• Normal Tensile Strain
• Normal Lateral Strain
– Poisson’s Ratio
ν =−
Lateral Strain
; ν ≈ 0.33
Tensile Strain
– True Stress
σ=
F
A
– True Strain
dε =
d! ε = Ln !
!o
!
Notes
Elastic Behavior of Materials
• Stress-Strain Relationship
Ashby
Ashbyand
andJones
Jones1980
1980
6
Notes
Elastic Behavior of Materials
• Stress-Strain Relationship
– Hooke’s Law
• Normal Tensile
σ = Eε n
• Simple Compression
τ = Eγ
• Volumetric Shrinkage
p = −κ∆
– E: Young’s Modulus
– G: Shear Modulus
– K: Bulk Modulus
u (E ) = u (σ ) e.g . psi, ksi, Nm −2
κ ≈ E ; G ≈ 23 E
Notes
Problem Solving - Stress/Strain Definition
• A low carbon structural steel sample of
dimensions 0.5 in. x 0.5 in. x 6 in. was pulled
along the long axis during tensile testing. The
2.00 in. gage marking originally notched on
the specimen measured 2.5 in. apart after the
test. Calculate the engineering strain and the
percent elongation that the sample
experienced.
(Ans. ε = 0.25, % Elongation = 25)
7
Notes
Problem Solving - Stress/Strain Definition
• Calculate the engineering stress and strain,
the true stress and stain using the tensile
testing data reported in the following:
– Load applied to specimen = 25,000 lb
– Initial specimen diameter = 0.50 in.
– Diameter of specimen under load = 0.45 in.
(Ans. Eng. Stress = 127,600 psi, Eng. Strain = 0.23
True Stress = 157,000 psi, True Strain = 0.21)
Notes
Deformation of Materials
• Stress-Strain Relationship
Ashby
Ashbyand
andJones
Jones1980
1980
8
Notes
Deformation of Materials
• Stress-Strain Relationship
Ashby
Ashbyand
andJones
Jones1980
1980
Notes
Deformation of Materials
• Stress-Strain Relationship
True Stress
σ=
F
A
True Strain
dε =
d!
!
ε = Ln
!
!o
Gladman
Gladman1997
1997
9
Notes
Plastic Behavior of Materials
• Yield Strength
– Lower and Upper Yield Point
– Lüders Band
(Hall 1970)
(Hall 1970)
Notes
Problem Solving - Plastic
Deformation
• A sheet of thermomechanical controlled
processing (TMCP) steel is cold rolled 20% to
a thickness of 5.00 mm The sheet is then
further cold rolled to 3.00 mm. What is the
total percent cold deformation?
(Ans. Total cold deformation = 52%)
10
Notes
Problem Solving - Plastic
Deformation
• Calculate the percent cold reduction when an
annealed eutectoid steel wire (with 0.8 wt.
pct. carbon) is cold drawn from a diameter of
2.5 mm to a diameter of 1.25 mm.
(Ans. Total cold reduction = 75%)
Notes
Problem Solving - Plastic
Deformation
• Calculate the percent cold reduction when an
annealed eutectoid steel wire (with 0.8 wt.
pct. carbon) is cold drawn from a diameter of
2.5 mm to a diameter of 1.25 mm.
(Ans. Total cold reduction = 75%)
11
Notes
Plastic Behavior of Materials
• Dislocation-Grain Boundary Interaction
• Strain Hardening
τ = αGb ρ
( )
1
2
Sub-Grain
Boundary
Dislocations pile-up against
grain and sub-grain boundary
Grain
Boundary
Notes
Fracture Appearance
• Ductile Fracture
(Slow Crack Growth)
– Cup and Cone Failure
– Fibrous Zone
– Shear Lip
• Brittle Fracture
(Fast Crack Growth)
– Radial Zone
MnS
Carbides/
Silicates
Shear
Lip
Metals
MetalsHandbook
HandbookV.10
V.10
Fibrous
Zone
Radial
Zone
12
Notes
Plastic Behavior of Materials
• Hardness Testing
– Not a Well-Defined Property
• Tests use Different Combinations of the Elastic,
Yielding and Work Hardening Characteristics
– Relates Better to Tensile Strength than
Yield Strength
σ UTS ≈ 3.2 H V
– Types of Measurement
• Indentation - Brinell, Vickers, Rockwell
• Height of rebound - Shore
Notes
Plastic Behavior of Materials
• Hardness Testing
– Measures Size of Indentation of Prescribed
Geometry Under a Known Load
• Brinell
–
–
–
–
Indenter made of Hardened Steel Sphere
10 mm Diameter
3000 kg Load (Steel)
Constant Load/(Diameter)2 Ratio
Load
Contact Surface Area of Indentation
P
HB =
u (H B ) = kg f mm − 2
πD
D − D2 − d 2
2
HB =
(
)
13
Notes
Plastic Behavior of Materials
• Hardness Testing
– Measures Size of Indentation of Prescribed
Geometry Under a Known Load
• Vickers
–
–
–
–
Diamond Square-Based Pyramid Indenter
Indenter Included Angle - 136o
Variable Load - 120 kg to 5 kg to grams
Geometrically Similar Impressions Under Different
Loads
136
2 P sin
P
2
HV =
HV = 1.8544 2
L2
L
u (HV ) = kg f mm 2
Notes
Plastic Behavior of Materials
• Hardness Testing
– Measures the Depth of the Indentation
• Rockwell C, A, D
–
–
–
–
Diamond Cone Indenter
Indenter Included Agnle - 120o
Pre-Load of 10 kgf
Variable Load - 150 kg f (C), 60 kgf (A), 100 kgf (D)
• Rockwell B, E, F, G, …
Smith
Smith1993
1993
– Indenters made of Hardened Steel Spheres
– Variable Diameters - 1.59 mm (B, F, and G), 3.17
mm (E)
– Variable Load - 100 kg f (B), 100 kg f (E), 60 kgf (F),
150 kgf (G)
14
Notes
Plastic Behavior of Materials
Hardness
Hardness
Testing
Testing
Hayden,
Hayden,Moffatt,
Moffatt,
Wulff
Wulff1965
1965
Notes
Plastic Behavior of Materials
• Hardness Testing
– Measures the Height of Rebound of an
Indenter
• Shore
– Diamond-Pointed Hammer Weighing 2.5 grams
– Hammer Falls from Standard Height Down a
Graduated Tube
– Index of Hardness is the Height of the First Rebound
15
Effect of Deformation Rate
Notes
• Deformation Rate (ε" )
– Static (10-2 - 10-5 s-1)
• Tensile Testing
• Fracture Toughness Testing
– Dynamic (100 - 10-2 s-1)
• Charpy Impact Testing
– Shock Loading (102 - 104 s-1)
• Explosive Forming
• Material Behavior at Different ε"
– σ increases with increasing ε"
– Toughness decreases with increasing
ε"
Toughness
Toughnessisisaameasure
measureofofthe
theamount
amountofofenergy
energy
that
thataamaterial
materialcan
canabsorb
absorbbefore
beforefracturing.
fracturing.ItIt
isisrelated
relatedtotocrack
crackpropagation.
propagation.
Notes
Charpy V-Notch Impact Testing
Smith
Smith1993
1993
16
Notes
Charpy V-Notch Impact Testing
Upper
Shelf
Energy
Lower
Shelf
Energy
Smith
Smith1993
1993
Ductile
DuctileBrittle
BrittleTransition
TransitionTemperature
Temperature- -DBTT
DBTT
Fracture
FractureAppearance
AppearanceTransition
TransitionTemperature
Temperature- -FATT
FATT
Notes
Fracture Toughness vs Strength
Chart
Chartfor
for Elastic
Elastic
Design
Design
•Contours
•Contours show
show
Process
ProcessZone
Zone(Plastic
(Plastic
Zone)
Diameter
Zone) Diameter
•Selection
•Selection of
ofMaterials
Materials
for:
for:
–– Yield-before-Break
Yield-before-Break
Design
Design
–– Leak-before-Break
Leak-before-Break
Design
Design
Ashby
Ashbyand
andJones
Jones1980
1980
17
Notes
Fracture Toughness of Materials
• Fracture Mechanics Approach
– Linear Elastic Fracture Mechanics (LFEM)
• Stress and Defect Dependent
K = Y EG = Yσ πa
• KI - Stress Intensity Factor (Mode 1)
• KIC - Fracture Toughness
K IC = Yσ f πa
σσ--Applied
AppliedNormal
NormalStress
Stress
aa--Edge
EdgeCrack
CrackLength
Lengthor
orHalf
Halfthe
theLength
Lengthofofan
anInternal
InternalThrough
ThroughCrack
Crack
YY--Dimensionless
DimensionlessGeometric
GeometricConstant
Constantofofthe
theOrder
Orderofof11
σσf --Fracture
FractureStress
Stress
f
Notes
Problem Solving - LEFM
• For a particular engineering application, a
2024 aluminum alloy plate must support 220
MPa in tension. Determine the largest internal
flaw size that this material can support.
– KIC=26.4 Mpa.m0.5
– Assume Y=1.
(Ans. 2a = 9.2 mm)
18
Notes
Fatigue of Materials
•Repeated Loading
Conditions
– Cyclic Loading
– S-N Curves
(Fati gue
Stren gth)
Smith
Smith1993
1993
Notes
Fatigue of Materials
σ + σ min
•Mean Stress σ m = max
2
•Range of Stress σ r = σ max − σ min
•Stress Amplitude σ a =
•Stress Ratio R =
σr
2
σ min
σ max
19
Notes
Fatigue of Materials
• S-N Curves
– Endurance Limit, Fatigue Strength
Smith
Smith1993
1993
Notes
Fatigue of Materials
• Micromechanism of Fatigue
– Striations
• Intrusion and Extrusion
Metals
MetalsHandbook
HandbookV.10
V.10
20
Notes
Fatigue of Materials
Hardened Steel Connecting Rod
Inclusion initiated Fatigue Fracture.
Metals Handbook V. 10
Notes
Fatigue of Materials
Hardened Steel Valve Spring
Inclusion initiated Fatigue
Fracture
Metals Handbook V. 10
21
Notes
Fatigue of Materials
• Fatigue Crack Propagation vs Stress and
Crack Length
da
∝ f (σ , a )
dn
da
= A∆K m
dn
log
(
da
= log A∆K m
dn
)
da
= fatigue crack growth rate (mm / cycle , in / cycle )
dn
∆K = K max − K min = Stress Intensity Factor Range MPa m , ksi in
A, m = Constants f (material , enviornmen t , frequency , T , R )
(
)
Notes
Fatigue of
Materials
log
(
da
= log A∆K m
dn
)
Smith
Smith1993
1993
22
Notes
Fatigue/Fracture Toughness of Materials
Crack
CrackMonitoring
MonitoringSystem
System
Smith
Smith1993
1993
Notes
Problem Solving - Fatigue
• An alloy steel plate is subjected to repeated tensile
and compressive loading with constant amplitude.
The uniaxial fatigue cyclic stresses have magnitudes
of 120 and 50 MPa, respectively. Given the static
properties of the plate: yield strength of 1500 Mpa,
fracture toughness K IC of 45 MPa.m0.5. If the plate
contains a uniform through thickness edge crack of
1.0 mm, how many fatigue cycles are estimated to
cause fracture?
– Use da
3
−12
= 2.0 x10 ∆Κ
dN y = 1
– Assume
(Ans. Number of cycles to fracture = 2.80 x 106)
23
Notes
Creep of Materials
Creep
CreepofofLead
LeadPipes
Pipes
Ashby
Ashbyand
andJones
Jones1980
1980
Notes
Creep of Materials
• Structures that operate at high temperature (e.g.,
reactors, steam plants, chemical plants, turbines)
• Slow, continuous and permanent deformation
with load
ε = f (σ , t , T )
• Homologous temperature
Th =
T
; Th > 0.5Tm (K )
Tm
Metals : Th > 0.3 ~ 0.4
Ceramics : Th > 0.4 ~ 0.5
24
Notes
Creep of Materials
• Mechanism of Creep
– Dislocation Creep - Power Law Creep
• At Th>0.5, atom (bulk) diffusion allowing dislocation to
climb and glide away from obstacles under applied stress
• At Th between 0.3 and 0.5, core diffusion predominates
ε"ss = Aσ n e
−
Q
RT
– Diffusional Creep - Linear-Viscous Creep
• At high Th, bulk diffusion predominates
• At low Th, grain boundary diffusion and grain boundary
sliding predominate
Q
−
ε"ss =
Cσ e RT
d2
Notes
Creep of Materials
• Deformation Mechanism Maps
(Summary of competitive creep mechanisms on
normalized stress vs. homologous temperature
space)
bulk diff usion
• Creep Damage
Ashby
Ashbyand
andJones
Jones1980
1980
25
Notes
Creep of Materials
• Creep Testing
– Sample typically loaded in tension, at constant load,
and at constant temperature
– Steady-State Creep (Power Law Creep)
– Creep-Rupture Diagram
Ashby
Ashbyand
andJones
Jones1980
1980
Notes
Creep of Materials
Cast
CastNi-Alloy
Ni-Alloy
Turbine
TurbineBlade
Blade
Creep
CreepFailure
Failure
Metals Handbook V. 10
26
Notes
Creep of Materials
Cast
CastNi-Alloy
Ni-Alloy
Turbine
TurbineBlade
Blade
Stretching
Stretchingand
and
Necking
Neckingdue
due
to
toCreep
Creep
Metals Handbook V. 10
Notes
Problem Solving - Creep
• Determine the time to stress rupture at 850oC
for equiaxed MAR-M 247 alloy. The component
is loaded to 207 MPa.
(Ans. 16,627 hours)
27
Notes
Mechanical Behavior of Materials
• Development of Mechanical Properties
– Solid-Solution Strengthening
– Grain Refinement
– Precipitation Strengthening
– Microstructural Modification
Notes
Development of Mechanical Properties
• Solid Solution Strengthening
(Takeuchi 1969)
∆H El , Misfit = σ p 4πro2 (r − ro )
∆H El , Misfit = 4 µbedgeη a ro3
sin θ
r
El , Misfit
∆H screw
=0
Elastic Distortion =
∆a
⋅100
a
NNa - -size
misfit parameter
a size misfit parameter
bbedge - -edge
component of a dislocation
edge edge component of a dislocation
σσp - -hydrostatic
hydrostaticcomponent
componentofofaastress
stressfield
field
p
28
Notes
Development of Mechanical Properties
• Grain Refinement - Hall-Petch Relationship
(Petch 1959)
Hall-Petch Equation:
σ y = σ i + κ yd
− 12
Combined Solute and Grain Size
Strengthening:
σ y = σ i + ∑ κ iCi + κ y d
− 12
i
Stress
σσi - -Friction
i Friction Stress
Coefficient
κκy - -Strengthening
y Strengthening Coefficient
dd- -Grain
GrainSize
Size
CCi - -Concentration
ConcentrationofofSpecies
Speciesi i
i
Notes
Development of Mechanical Properties
• Precipitation Strengthening
Orowan
Orowan Equation
Equation
∆τ y =
Gb
L
∆τ c = Af 2 r 2
1
1
∆τ ! = B ⋅ Gbf 2 r −1
1
(Takaheshi and Nagumo 1970)
LL--Particle
ParticleInterspacing
Interspacing
f f--Volume
VolumeFraction
FractionofofPrecipitates
Precipitates
rr--Mean
Particle
Mean ParticleSize
Size
29
Part 2
Background Material
30
Background Materials
As indicated in the preface, this 4-hour module has the objective of presenting
introductory concepts in mechanical properties of materials to nonmetallurgical/materials engineers. Therefore, the content was selected and prepared based
on the author’s own experience in the field of metallurgical engineering. The references
consulted were basic, elementary texts typically used in undergraduate metallurgical
engineering education. In fact, any solid textbook on mechanical metallurgy should
contain the information presented in this module. The list below is, by no means,
exhaustive and only represents some good titles that readers can refer to and further learn
about mechanical properties of materials.
1. Engineering Materials 1 – Michael Ashby and David Jones, Pergamon Press, 1980.
2. Materials Selection in Mechanical Design – Michael Ashby, Butterworth Heinemann,
1992.
3. Materials Considerations in Design – Arnold Ruskin, Prentice Hall, 1967.
4. Foundations of Materials Science and Engineering – William Smith, McGraw Hill,
1993.
5. The Science and Design of Engineering Materials – James Schaffer, Ashok Saxena,
Stephen Antolovich, Thomas Sanders, Jr., and Steven Warner, WCB-McGraw Hill,
1999.
6. The Structure and Properties of Materials – Vol. III – H. Hayden, W. Moffatt, and J.
Wulff, John Wiley & Sons, 1965.
7. Deformation and Fracture Mechanics of Engineering Materials - Richard Hertzberg,
John Wiley & Sons, 1976.
8. The Plastic Deformation of Metals – Robert Honeycombe, Edward Arnold, 1968.
9. Mechanical Metallurgy – George Dieter, McGraw Hill, 1986.
10. Mechanical Behavior of Materials - Thomas Courtney, McGraw Hill, 1990.
11. Materials for Engineering - John Martin, The Institute of Materials, 1996.
12. The Testing of Engineering Materials – Harmer Davis, George Troxell, and George
Hauck, McGraw Hill, 1982.
13. Ensaios Mecânicos de Materiais Metálicos (Mechanical Testing of Metallic
Materials) – Sergio A. de Souza, Edgard Blücher, 1974.
14. Metals Handbook – Vol. 8 - Metallography, Structures and Phase Diagrams – ASM
Metals Handbook Series, 1973.
15. Metals Handbook – Vol. 10 – Failure Analysis and Prevention – ASM Metals
Handbook Series, 1975.
16. Metals Handbook – Desk Edition, ASM International, 1984.
17. Damage Tolerance Design Handbook – MCIC-HB-01, Battelle Columbus Labs,
1975.
18. Advanced Materials and Processes – ASM Journal, 1990.
Chapter III - Introduction to Metallurgical Principles
STRUCTURE OF METALS
Metals are crystalline solids whose atoms are arranged in regular patterns. Using highly sophisticated tools such as the
transmission electron microscope and field-ion microscope, the individual atoms could be observed to form long range geometric
patterns. This orderly atom arrangement is responsible for the crystalline structure of a metal and its many properties (physical,
chemical and mechanical).
The three most common crystalline structures found in metals are body-centered cubic, face-centered cubic and hexagonal
close-packed. Their atom arrangements are shown schematically in Fig. 3-1. The black dots represent the location of individual
atoms. The body-centered cubic (bcc) structure exhibits one atom at each corner of a cube and one in the center. Notice that the
bcc structure is not close-packed. In fact, bcc crystals are only 68% packed. Thus, the empty space in a bcc crystal can
accommodate atoms of other elements displaying a certain degree of solubility. The face-centered cubic (fcc) structure has an
atom at each corner of the cube and one in the middle of each face. By comparison, this crystal structure is more closely-packed
than bcc. Hexagonal close-packed structure exhibits a slightly different atomic arrangement than the fcc structure, with the atoms
located along horizontal planes known as the basal planes.
Figure 3-1. The three most common crystal structures in metals.
left – bcc, center – bcc, and right - hcp.
Different elements crystallize in different crystal structures. For example, iron crystallizes as bcc at room temperature, copper
and aluminum crystallizes as fcc and magnesium as hcp. Table 3-1 provides additional information regarding crystal structure of
different metals.
Table 3-1. Crystal structures of common metals.
Structure
Body-centered cubic (bcc)
Face-centered cubic (fcc)
Metals with this structure
Fe (at room temperature, α-ferrite, and near its melting point, δ
-ferrite),
Cr, Si, Nb, Mo, V, W
Fe (at intermediate high temperature, γ
-austenite), Ni, Cu, Al, Ag, Pb
Hexagonal close-packed (hcp)
Mg, Sn, Zn,
FORMATION OF CRYSTALLINE SOLIDS
Crystalline solids are usually produced by freezing or solidifying a liquid metal. In the liquid state, the molten metal loses its
long-range ordering of atoms. When a melt starts to cool to its solidification temperature, atoms aggregate to form submicroscopic particles with specific atomic ordering, characteristic of the particular metal, called nuclei. These particles form at
preferred sites such as the mold surface, second-phase particles, etc. As temperature continues to drop, more atoms will attach on
these nuclei, promoting growth of the solid phase and at the expense of the liquid phase until the completion of solidification.
Since each of the nuclei possesses a definite crystal orientation and atomic spacing, these will grow into larger solid particles
with independent crystallography that are called grains. These grains meet along grain boundaries which are typically regions
where the periodic and orderly arrangement of atoms is disrupted. Because of this atom disorder at grain boundaries, there are
often differences in metal behavior at these locations.
Most common engineering metals are alloys consisting of one major element and variable amounts of one or more alloying
elements and whose properties are usually different from those of the pure metal. Alloys provide, in general, combinations of
engineering properties that are superior, for specific applications, to those of the unalloyed metals. The crystal structure, the
purity and the prior thermal and mechanical history of an alloy all have significant influence on the engineering properties of an
alloy.
In the making of an alloy, some of the parent metal atoms are replaced by new atoms, which take the positions previously
occupied by the parent metal atoms. The replacement has a direct bearing on the properties of the alloy. The basic ways in which
the new atoms are incorporated in a metal are listed below and illustrated in Figure 3-2:
1. Direct Substitution. If the new atom is similar in size and chemical behavior to those in the original unalloyed metal, it
may directly replace one of the original lattice atoms. Thus, the new atoms are “dissolved” in the parent metal to form a
solid solution. Examples of solid solutions are gold dissolved in silver, or copper in nickel.
2. Interstitial Solid Solution. When the new atom is small in relation to the parent atom, it can dissolve in the original
structure, in the spaces between the parent metal atoms, without actually displacing any of them. In this case, an
interstitial solid solution is formed. Small amounts of carbon and nitrogen, for example, can occur interstitially in iron.
Figure 3-2. Schematic illustration of (a) substitutional and
(b) interstitial solid solution.
More often, the new atoms cannot completely dissolve in the parent metal, either interstitially or substitutionally. When the
parent and alloying metal exhibit strong chemical affinity, the mutual solubility is generally lowered with increased probability of
a chemical reaction. These circumstances usually result in the formation of intermetallic compounds of definite (or range of)
chemical composition and crystal structure. Thus, it is entirely conceivable that the solid solutions coexist with the intermetallic
phase resulting in a multiphase alloy. The individual phases may be seen and distinguished if the alloy is properly polished and
etched and then examined under a microscope at magnifications from 50 to 2000 times. This procedure is known as
metallography and results of metallographic preparations are microstructures such as those illustrated in Figs. 3-3.
The microstructure of alloys usually consists of many randomly oriented grains of the different phases present. The overall
arrangement of grains, boundaries and phases present in a metallic alloy is called its microstructure and is largely responsible for
the properties of the alloy. The microstructure is affected by the composition or alloy content, and by other factors such as
heating and cooling associated with forming and heat treating operations. The microstructure is greatly affected by the welding
operation, which, in turn, influences the properties of the alloy.
Figure 3-3. Typical microstructure of: (a) low-carbon pearlitic steel at 100X – white
phase is ferrite and dark phase is pearlite; (b) commercial nickel-base superalloy,
Udimet 700 at 1000X – multiphases present.
PHYSICAL AND CHEMICAL PROPERTIES
Metals exhibit certain properties that make them useful as engineering materials and allow them to be joined by
welding processes. Some of these properties are electrical conductivity, thermal conductivity, ionization potential,
work function, and ductility. Electrical conductivity and thermal conductivity of metals generally parallel each other
because both involve electron in the transport of energy. Copper and aluminum are good examples of metals with
high electrical and thermal conductivity. For the same reasons, they are often used as conductors and electrical
wiring. While materials with high electrical conductivity are good candidates for electrical welding, high thermal
conductivity may actually jeopardize the welding process because of the high rate of heat extraction from the weld
region. Table 3-2 illustrates some of these characteristics.
Table 3-2. Selected physical properties of metals and their weldability.
Copper
Thermal
Conductivity
W.m-1.oC-1
394
Electrical
Conductivity
(x108) Ω .m
1.72
Aluminum
222
2.83
Require high welding current
Carbon Steel
46
16.0
Readily weldable
Metal
Welding Behavior
Difficult to weld
Arc welding stability depends on the number of charge carriers (e- and positive ions, mainly) in the plasma. Low
ionization potential means easier removal of an electron from an atom. Elements such as potassium and sodium have
low ionization potential, 4.3 and 5.12 eV, respectively, and are often added to electrode coatings to stabilize the
welding arc.
Aside from physical properties, the chemical reactivity of metals is also important.
MECHANICAL AND FRACTURE PROPERTIES
At grain boundaries, the arrangement of atoms is irregular and there are many holes (vacancies) or missing atoms (see again
Fig. lid). The atom spacing may be larger than normal with the result that individual atoms can move about relatively easily in the
grain boundaries. Because of this, the diffusion of elements— i.e., the movement of individual atoms through the solid
structure— generally occurs more rapidly at grain boundaries than within the grains. Because of the disarray, it is easier for oddsized atoms to segregate at the boundaries. This segregation frequently leads to the formation of undesirable phases that
adversely affect the properties of a material by reducing its ductility or making it susceptible to cracking when welded.
Many unique things that profoundly affect the properties of an alloy occur at grain boundaries. Grain boundaries, for example,
increase the strength of materials at room temperature by inhibiting the deformation of individual grains when the material is
stressed. At elevated Temperatures, the atoms in the boundaries can move easily and slide past one another, thus reducing the
material’s strength. As a result, fine-grained materials have better properties for room temperature service while coarse-grained
materials are desirable for high temperature service. The structure of a metal could be characterized as having either few large
grains (coarse-grained), or many small grains (fine-grained) or a mixture of both large and small grains (mixed grain size).
The usefulness of a material or alloy is measured and described by its properties. Some of the more important physical and
mechanical properties and tests used to determine them are listed below.
HARDNESS
Probably the most commonly and easily measured material property is hardness. The hardness of a material is basically its
resistance to indention and has come to represent, in practice, an indication of a material’s strength. Most common hardness tests
consist of using a fixed applied load to force an indentor into the material being tested. Typical hardness indentations are shown
schematically in Fig. 16. Indentors are usually of a very hard material— in many cases diamond— and come in a variety of
shapes. Typical shapes of indenters used for the various tests are also shown in Fig. 16. The applied load that produces the indention ranges from a gram to 3000 Kg.
The factor indicative of the hardness level is either the depth of the penetration or the size of the impression. Thus the diagonal
of an impression made with a pyramid or the diameter of an impression made with a sphere can be measured. Such a
measurement is then converted into a hardness number by means of appropriate tables. The hardness number is always directly
proportional to the ability of a material to resist indention. The specific hardness test used depends on the purpose for which the
test is made as well as the size of the test material.
Heavy applied loads make large, deep impressions while small light loads make small, shallow impressions, some so small
that they can only be seen with a microscope. In order to measure the hardness of thin sheet material, a light load is used. To
measure the hardness of a larger, but still restricted region such as a weld heat-affected zone, an intermediate load of perhaps 50
to 300 grams would be used. However, if the over-all hardness of a material is to be determined, a high load, such as the 3000 Kg
Brinell test, would be used. The large impression produced by the high load and large indenter will not be influenced
significantly by very local structure sensitive hardness fluctuations since they are averaged out. In such cases, more accurate
average hardness measurement will be obtained.
As noted above, the hardness of a material is roughly proportional to its strength. However, the strength of a material must be
determined independently and then related to its hardness. Once this relationship is known, the hardness test, because of its
simplicity, becomes a valuable tool.
STRENGTH
The strength of a material is the measure of its ability to sustain an applied load without failing, or deforming significantly.
The greater a load the material can carry, the greater its strength. So that strength measurements can be universally used, they are
described as pounds of load per square inch of cross-sectional area (psi), i.e., stress. The most frequently determined strength
criterion is tensile strength even though it is often desirable to measure other kinds such as compressive or shear strengths.
In a tensile test, a specimen of some fixed geometry is used. A typical test specimen is shown in Fig. ha. The sample is loaded
(stressed) in the axial direction by means of an appropriate machine, usually hydraulic: a typical apparatus is shown in Fig. 17b.
Initially, for every increment of applied load the bar elongates a proportional amount. This is represented in Fig. 17c by the line
between points A and B. Material behavior of this type is called elastic. That is, the material is stretched somewhat like a rubber
band. As the load is applied, the material stretches, but when the load is released, the sample returns to its original size. Thus in
the elastic region, there is no permanent change in the size of the sample due to any applied load. During elastic extension, the
spacing between individual atoms increases slightly in the tensile direction but there is no relative motion between the atoms
which would cause atoms to slide past each other any great distance so that, as the load is removed these atoms move back to
their original positions. In the elastic region, the proportionality between the applied stress and the resultant elongation, or strain,
is called Young’s Modulus (E),
When the load on the test sample is increased beyond that of point B in Fig. 17c, the proportionality between stress and strain
no longer exists; that is, each increment of load produces a greater increment of strain than it did in the elastic region. Thus point
B is called the elastic limit of the material, since beyond this point the material begins to behave plastically. Plastic deformation
is permanent. When the load is removed from the sample, it does not return to its original size. On the atomic scale, when the
elastic limit has been exceeded, the atoms move within the material and do not return to their original locations when the load is
removed.
From a practical standpoint, the elastic limit is a difficult point to determine with any precision. For most engineering work, a
more practical measurement of the upper limit of elastic behavior is the yield strength. The yield strength is the stress required to
produce a small, fixed arbitrarily chosen amount of permanent strain if the load were removed at that point. The most common
amounts of permanent strain used to designate yield strengths are 0.02% and 0.2%. The 0.02% yield strength is illustrated as
point C in Fig. 1.7c.
It should also be noted that for some materials the yield strength is characterized by a particular kind of yielding rather than by
an arbitrarily defined amount of permanent strain. In soft, low-carbon steel when the elastic limit is exceeded, an abrupt and
substantial amount of strain or elongation occurs. This form of abrupt yielding is shown schematically in Fig. I 7d also. This kind
of behavior is often described as a yield point.
As the load is further increased beyond the yield point or yield strength, greater amounts of plastic strain occur. Point D on
Fig. 1 7c indicates the maximum load that the sample will support. The stress associated with this load is the ultimate tensile
strength. While the load increases up to this point the test sample not only gets longer, but also undergoes uniform reduction in
diameter. At point D an instability in the sample occurs and the gage diameter starts to decrease more rapidly in one region than
do other parts of the specimen. This behavior is called necking. Since the area of the sample decreases rapidly, the load required to
continue elongating the sample also reduces until finally the sample breaks at point F. If the fracture stress, which is the fracture
strength (or load) divided by the cross-sectional area of the sample at the fracture location, were determined, it would be greater
than the ultimate tensile stress. As an engineering value, the ultimate tensile strength is considerably more useful than the fracture
stress.
The tensile test determines the strength of a material and its ductility. The ductility of a material is basically its ability to
deform plastically without failing. Two measures of ductility are obtained from the tensile tests: the percentage of elongation and
the percentage of area reduction. The manner in which the properties are determined is shown by the equation below.
% Elongation =
Final length— Original length
Original length
—
X 100
(12)
Original area— Final area
>< 100
% Reduction in area
=——
(13)
Original area
While yield strength and tensile strength measurements are used in designing, the
ductility of a material is not. Nevertheless, ductility is an important property and
generally the more ductile a metal is, the better it is. Minimum levels of ductility are
customarily determined and specified by experience and empirical rules.
The tensile strength of a material is most frequently determined at room temperature, but it can be determined at any
temperature. In general, as the test temperature increases, the tensile and yield strength decreases while the ductility increases.
The opposite of ductility is brittleness, a familiar but often confusing term. Brittle behavior is fracture without much prior
ductility or deformation. Copper is ductile because it deforms a great deal before it breaks; glass is brittle because it breaks
almost immediately without deforming. Both are strong materials. Although related, ductility is not a direct measure of strength,
but to be useful a material must normally be both strong and ductile. A strong and ductile metal is often said to be a tough
material.
Fatigue Strength
If a stress less than a materials tensile strength is applied, it will nor break; if the same load is removed and reapplied several
times it may eventually break. Repeated bending of a paper clip to the point of failure is an example of a type of fatigue. The
strength of a material to withstand repeated load applications is called fatigue strength. The fatigue strength is usually related to
the number of cycles required to reach the point of failure. Fatigue strength is commonly determined by applying a stress first in
tension and then at the same level in compression. That is the maximum limiting stress that is cycled above and below the zero
stress level in Fig. 18. The closer the maximum stress to the tensile strength, the fewer the load cycle applications required before
fracture occurs. As the stress level is reduced, a greater number of cycles are experienced before fracture occurs. In many
materials the point at which the material will fracture is never reached, no matter how many cycles the load is applied. The stress
associated with this point is called the endurance limit. The typical fatigue behavior of a material is shown in Fig. 18. The
endurance strength of a material is frequently equal to roughly half of its tensile strength.
Creep Strength
If a load below a material’s tensile strength is applied at room temperature, the material initially elongates as the load is
applied; sustaining the same load, however, causes no further measurable elongation. Yet if the same load were to be applied at
an elevated temperature, the material would continue to gradually elongate as long as the load was maintained. This behavior is
called creep. Maintaining the load long enough would eventually cause the material to rupture.
Two factors other than stress are important in describing a material’s creep or rupture strength: time and temperature. The
higher the temperature, the shorter the time needed to produce a given amount of creep or to produce rupture failure for a given
applied load. At a given temperature, a higher load causes creep and rupture to occur sooner. Thus creep strength is designated as
the stress required to produce a given amount of creep strain in a given period of time at a given temperature, e.g., 0.19~ creep
strain in 100,000 hours at 10000F. Rupture strength is the stress to produce failure after a given period of time and at a given
temperature. The general relationships between stress, time and temperature in the case of rupture are shown in Fig. 19. Similar
curves could be drawn for creep deformation.
FRACTURE TOUGHNESS
A material with normal strength and ductility can behave quite differently if loaded under certain conditions, such as in a
notched condition, at a low temperature or very abruptly. A combination of the above factors can frequently cause a metal to fail
in a brittle manner at an effective strength much lower than its normal value. Such behavior is called notch sensitivity or brittle
failure. It is, in a sense, the opposite of notch toughness. Notch tough materials are those that are relatively insensitive to notches
and cold or impact loading. They usually fail in a reasonably ductile way in spite of the above-mentioned “embrittling” test
conditions.
A common test to measure fracture toughness is the Charpy Notched Bar Impact Test. The specimen used is shown in Fig. 20.
In this test, the energy required to break the test bar over a range of temperatures determines the notched fracture roughness. In
the test a pendulum hammer strikes the bar from a known height. As shown in Fig. 20, it is normal for the energy absorption to be
lower at lower temperatures. In testing a material over a range of temperatures, the failure mode sometimes shifts rather abruptly
from tough at high temperatures to brittle at low temperatures; the temperature at which it occurs is called the transition
temperature. Toughness or brittleness is important characteristics in welded structures. Many factors affect fracture toughness,
and not all materials are equally tough. Tougher materials have low transition temperatures and their fracture, at all temperatures,
usually requires greater energy.
TYPICAL PHASE TRANSFORMATIONS IN METAL SYSTEMS
PHASE TRANSFORMATIONS (CRITICAL POINTS)
Differences in temperature cause the atoms of many metals to vary in their crystallographic arrangements. For example the
crystalline structure of iron at temperatures up to 16700F is body centered cubic; from 16700F to 25350F it is face centered cubic;
and from 25350F to the melting temperature, 27950F, it is again body centered cubic. This change in crystalline structure is
formally called an allotropic transformation or a phase transformation. Among other metals undergoing allotropic
transformations at various temperatures are titanium, zirconium and cobalt. Many factors, including chemical composition,
cooling rate and the presence of stress, influence this type of transformation.
Another kind of transformation or change occurs when a metal melts or solidifies. When a metal melts, its orderly crystalline
arrangement of atoms becomes a completely disorganized noncrystalline array. The changes in the crystalline structure of
metals— from one solid arrangement to another, or from solid to liquid— are sometimes grouped together under the heading of
phase changes. Pure metals solidify (and melt) at a single temperature. Most alloys solidify (and melt) over a range of
temperatures. The exceptions to this will be mentioned later.
Phase Diagrams
Events such as phase changes and solidification are best shown by means of a drawing called a phase diagram (sometimes also
referred to as equilibrium diagram or constitution diagram).
A phase diagram has been called the metallurgist’s road map— from a diagram for a given alloy system, he can find for any
given alloy composition at any specified temperature what phases and what percentages of each are present. He can also
determine what phase changes tend to take place with either a change in composition, a change in temperature or both. It has at
least one significant limitation. It is only an approximation of how alloys actually behave, however; for phase diagrams describe
alloy system behavior under equilibrium conditions rarely encountered in practice. This is especially true under conditions found
in welding because of the fast heating and cooling rates. (Equilibrium implies that a metal is stable in a desired state for a given
environment— that is, extremely slow heating and cooling conditions and long hold times.) Most readily available phase
diagrams describe alloy systems containing two elements (or components), while engineering alloys generally contain many
elements. Phase diagrams for more than two element systems are quite complex and difficult to interpret. However, they are still
the best way to study most alloy characteristics.
The phase diagram shown in Fig. 14 is for the Cu-Ni alloy system. This is called an isomorphous binary system— i.e., a twoelement system in which both elements are completely soluble in each other in both the liquid and solid states. As shown in Fig.
14, the phase diagram is drawn with the alloy content plotted on the horizontal axis. The extreme left edge represents 100% of
one element (Cu in Fig. 14), and the extreme right edge is 100% of the other element (Ni in Fig. 14). Temperature is plotted on
the vertical axis. Figure 14 shows that at temperatures above the line labeled “liquidus,’ the only phase is liquid for all
compositions. At temperatures below the line labeled solidus” the only phase is solid. Furthermore, all solid alloys formed are
single phase because copper and nickel are completely soluble in each other in the solid state. Thus an alloy with 30% Cu and
70% Ni is a homogeneous solid solution, remaining solid up to 2425”F and melting completely at 24900F. The combination is the
alloy Monel. In the region between the solidus and liquidus lines both solid and liquid phases coexist simultaneously. This
illustrates the fact that most alloys solidify, or go from complete liquid to complete solid, over a range in temperature. As the diagram shows, only for pure Cu or pure Ni does complete solidification occur at a single temperature.
Figure 13 shows a more complicated phase diagram for the silver-copper alloy system. The diagram is used extensively in the
designing of brazing alloys. At temperatures above the line labeled “liquidus,” all alloys are entirely liquid. At temperatures
below the line labeled ‘solidus” and within the regions labeled a and /3, the material present is all solid, but in two phases. The a
region defines the temperature-alloy conditions where Cu is soluble in Ag and the phase is a solid solution. The same is true for
the /3 region where Ag is in solid solution in Cu. In the regions between the solidus and liquidus lines on the left side of the
diagram, the phases present are the liquid solution of Cu in Ag and solid solutions of Cu in Ag (a). For the same region on the
right-hand side the phases present are the liquid and solid solutions of Ag in Cu (/3). Finally the area labeled (a + /3) is a
mechanical mixture containing grains of both a and /3.
The Ag-Cu alloy system is one in which each element has only a limited solubility in the other. The phase diagram depicts this
and shows that as Cu is added to Ag it first dissolves, but as more is added so that the a phase boundary is exceeded, the alloy
then contains the second solid phase /3, which is basically Cu with some Ag dissolved in it.
The phase diagram illustrates another feature, i.e., an eutectic point. This is labeled on the diagram and represents an alloy
composition which, like a pure metal, solidifies at one temperature. The eutectic composition solidifies somewhat uniquely: as
the liquid cools, both the a and /3 phases are formed concurrently. Quite frequently they occur as alternating platelets and hence
have a distinctive appearance.
A final feature of the Cu-Ag phase diagram is that it shows solid solubility in relation to temperature. In principle, this is
similar to the fact that more salt can be dissolved in hot water than in cold. In Fig. 15, the boundary line between the a ± /3
region and the a solid solution region shifts to lower alloy content as the temperature decreases. This means that if a given alloy
containing copper is dissolved in silver with the copper content such that the alloy falls just within the a solid solution region at a
high temperature, the solubility limit will be exceeded at lower temperatures and some /3 phase must be formed. Control of this
behavior is important in all precipitation-hardened alloys (see Chapter 4) -
FUNDAMENTALS OF WELD SOLIDIFICATION
Figure 11 shows this in a mold (and a weld). Because of this, a fusion weld can be viewed as a tiny casting. It should be
pointed out that the thermal conditions that exist during fusion welding produce cast structures with characteristics unique to
welding; these are complex and to date not thoroughly understood. Therefore, they will not be discussed at length in this
elementary text.
SOLID-STATE TRANSFORMATIONS IN WELDMENTS
EFFECT OF ALLOYING ELEMENTS
CRACKING PHENOMENA ASSOCIATED WITH WELDING
CORROSION OF WELDMENTS
OXIDATION RESISTANCE
When most metals are exposed to the air at elevated temperatures, there is a tendency for the oxygen in the air to unite with
the metal and form an oxide. The ability of a material to resist oxide formation or to resist continuing and rapid oxidation is
called its oxidation resistance. Metals such as gold, silver and platinum are very resistant to oxidation. Iron and copper oxidize
rapidly. This is generally undesirable, particularly in the case of iron and steels, since many oxides, once formed, do not adhere to
the metal. If exposure is continuous, the material gradually deteriorates. Metals such as aluminum and chromium also form
oxides readily when exposed to air. However, in the case of these metals, the oxide is very tightly adherent to the metal and
effectively seals it and prevents further oxidation. The chromium present in stainless steels serves this same function. Oxidation
resistance decreases as the exposure temperature increases.
CORROSION RESISTANCE
Corrosion resistance of a material is broader property than oxidation resistance since it takes in the resistance of a metal to any
kind of chemical or electrochemical attack, including oxidation. The presence of water or water solutions generally increases
corrosive action. Corrosion resistance is enhanced by the formation of tight adherent oxide coatings as in the case of oxidation.
PROPERTY CHANGES RELATED TO STRUCTURE CHANGES
The particular microstructure of a metal determines its properties. The state of microstructure is determined by, among other
things, heat treatment, alloy composition and fabrication history. Of particular interest is the fact that welding is an important
determining factor in the nature of a metal’s microstructure.
SUMMARY
The crystalline nature of a metal— pure or alloyed— is reflected in its microstructure, the over-all arrangement of grains,
boundaries and phases that determines its properties. The crystallographic arrangement of a given material is influenced by,
among other things, heat treatment, alloy composition and fabrication history including, of course, welding.
Certain definable, measurable mechanical and physical properties determine the usefulness of a metal or alloy. Hardness,
oxidation resistance and corrosion resistance are self-explanatory. Hardness is roughly proportional to strength. Tensile strength,
or the resistance of a material to a load in tension, is perhaps the most frequently used measurement of strength. The strength of a
material cannot be completely assessed, however, without relating to ductility, the capacity of a material to deform plastically
without breaking. Young’s Modulus (B — _stress strain) sets forth the relationship— in the elastic region— between the applied
stress and the resultant elongation, or strain. Because the elastic limit, beyond which point plastic deformation is permanent, is
difficult to determine precisely, engineers usually rely on a more practical measurement of the upper limit of elastic behavior:
yield strength. To measure yield strength, a load is applied to the material being tested, and removed when an arbitrarily chosen
amount of permanent strain, commonly 0.02% or 0.2% is produced.
A test load not exceeding a material’s fracture strength may be reapplied in order to determine fatigue strength; reapplied at
elevated temperatures, such a test load can be used to determine the amount of creep that will produce rupture failure. Another
important property, fracture toughness, is the ability of a material to exhibit strength and ductility if loaded in a notched
condition, at low temperature or very abruptly.
Metals’ behaviors under various types of stress can be recorded on graphs. Phase diagrams are used to show phase changes
and the solidification products characteristic of individual alloy systems, given certain conditions of composition and
temperature.
QUESTIONS ON METALLURGY
1.
1-low does the atomic arrangement of a metal differ in its solid and liquid states?
Describe how a molten metal gains its crystalline structure, and explain where and why there may be differences in the
solidified metal’s behavior.
3. Define ‘alloy” and state the basic ways in which such a metal can be created.
4. What constitutes an alloys microstructure?
5. Discuss the properties of alloys exhibiting the variations of grain size, mentioning in each case the behavior of the atoms at
grain boundaries.
6. Name two types of temperature-induced transformations and tell how each affects the structure of the metal involved.
7. What is a phase diagram? What are its uses and limitations?
8. What relationship does a metal’s hardness have to its strength? How is hardness measured?
2
I-low does tensile strength differ from yield strength, fatigue strength and creep strength? In what ways are these properties
interrelated?
10. Define ductility. Define brittleness.
11. What is fracture toughness and how is it measured?
12. Under what circumstances can oxidation be an important factor in corrosion resistance?
2.
SUGGESTED READING
A If/S Welding Handbook, Section 4, Fifth Edition, ‘Metals and Their Weldability” (1966)
AWS Welding Handbook, Section 1, “Fundamentals of Welding,’Sixth Edition (1968), Fifth Edition (1962)
Welding Metallurgy, Volume 1, G. F. Linnert, AWS (1965)
Metals and How to Weld Them, Second Edition, T. B- Jefferson, G. Woods, J. F. Lincoln Arc Welding Foundation,
Cleveland, Ohio (1962)
Elements of Physical Metallurgy, A. S. Guy, Addison-Wesley Publishing Co, Inc., Reading, Mass. (1959)
Physical Metallurgy, B. Chalmers, John Wiley & Sons, Inc., New York (1959)
Welding for Engineers, H. Udin, F. Funk, J. Wulif, John Wiley & Sons, Inc., New York (1954)
Structure and Properties of Alloys, R. M. Brick, A- Phillips, McGraw-Hill Book Company, Inc., New York (1949)
YOUR PATH TO LIFELONG LEARNING
ASME offers you exciting, rewarding ways to sharpen your technical
skills, enhance personal development and prepare for advancement.
*Short Courses - More than 200 short courses offered each year keep
you up to speed in the technology fast lane--or, help you fill in any gaps
in your technical background.
*Customized Training at your organization’s site - Do you have ten or
more people at your site who could benefit from an ASME course?
Most of our courses can be offered in-house and tailored to your latest
engineering project. Bring a course to your company too.
*Self-study materials meet the needs of individuals who demand
substantive, practical information, yet require flexibility, quality and
convenience. Return to each program again and again, as a
refresher or as an invaluable addition to your reference library.
*F.E. Exam Review - A panel of seasoned educators outline a wide
range of required topics to provide a thorough review to help
practicing engineers as well as engineering students prepare for
this challenging examination. 24 hours of videotape and notes.
*P.E. Exam Review - A comprehensive review of all the major exam topics
that demonstrates the necessary math, logic and theory...and provides
shortcuts that reduce the time and effort required to prepare yourself
for this challenging exam. 15 hours of videotape and notes.
FOR INFORMATION CALL 1-800-THE-ASME AND MENTION CODE CD.
INFORMATION REQUEST FORM
Please mail to ASME at 22 Law Drive, P.O. Box 2900, Fairfield, New Jersey, 07007-2900
or fax to 201-882-1717, call 1-800-THE-ASME, or email infocentral@asme.org
Send me information on the following:
_____Short Courses
_____ In-House Training _____ Self-Study Programs
_____F.E. Exam Review _____ P.E. Exam Review
Name___________________________________________________________
Title____________________________________________________________
Organization_____________________________________________________
Business Address_________________________________________________
City____________________________State_______________Zip___________
Business Phone ( )____________________Fax ( )______________________
E-mail Address____________________________________________________
Download