You are on page 1of 66

13

CHAPTER – 2

REVIEW OF LITERATURE

2.1 GENERAL
The mechanical properties and service-life of components evoked

considerable interest, particularly for those made of composite materials. The

tailor-made properties of composites could allow flexibility in design, since the

required properties might be obtained by economic use of expensive constituents.

As a result, new aspects of materials, such as the long-term behavior under

mechanical, thermal and chemical loadings are being addressed. Composite

materials have become more popular because of increased competition in the

global market for lightweight components having greater strength and stiffness.

Composite materials have the potential to replace widely used metals or their

alloys such as steel, aluminum and their alloys. It is revealed that 60% to 80%

weight saving can be achieved by replacing steel components with composite

components, while 20% to 50% weight saving occurs if aluminum parts are

replaced by composite parts (Mazumdar, 2002). The testing and evaluation of

many available composite material systems are quite complex, costly and time

consuming. Therefore, the prediction and analysis of creep properties, for

assessing useful life of components made of composite materials, is of great

practical importance for practicing engineers. The salient applications of

composite materials along with required properties are summarized in Table 2.1.
14

Table 2.1: Typical applications of composites

Applications Required Properties


Application Areas
Automobiles (Rohatagi et Combustion chambers (SiC-SiC), Increased stiffness,

al, 1992; Yue et al, 1998; Engine cylinder liners (Al-SiC), CNG Improved wear

Peters, 1998; Fitzpatrick et storage cylinders, Diesel Engine pistons resistance, Thermal

al, 1998; Hunt, 2000; (SiCw/Al-alloy), Brake rotors, Leaf fatigue resistance,

Surappa, 2003; Bayat et al, springs (E-glass/epoxy), Drive shafts Weight reduction, High

2007; Singh, 2008; Bayat (Al-C), Flywheels, Racing car brakes thermal conductivity

et al, 2008) (Al-SiC), Motorcycle drive sprocket,

Pulleys, Torque converter reactor,

Shock absorbers (SiCp/Al-alloy),

Radiator end caps.

Sub-Marine (Peters, 1998) Propulsion shaft (Carbon and glass Weight reduction,

fibers), Cylindrical pressure hull Durability

(Graphite/Epoxy), Sonar domes

(Glass/Epoxy), Composite piping

system, Scuba diving cylinders (Al-

SiC), Floats, Boat hulls.

Commercial and Industrial Computer hard disk drive, Needle for Weight reduction,

(Fitzpatrick et al, 1998; carpet-weaving machine, Electronic Increased specific

Peters, 1998; Hunt, 2000) Packaging/Thermal Management, stiffness and strength,

Pressure vessels, Fuel tanks, Cutting durability, High elastic

tool inserts, Laptop cases, Wind turbine modulus


15

blades, Electric motors, Firefighting air

bottles, Artificial ligaments, MRI

scanner cryogenic tubes, Wheelchairs,

Hip joint implants, Eyeglass frames,

camera tripods, Musical instruments,

Drilling tubes, Drilling motor shaft,

Drill casing, Crane components, High

pressure hydraulic pipe, X-ray tables,

Heart valves, Helmets, Crucibles,

Beams.

Aerospace equipment and Rocket nozzle (TiAl-SiC fibers), Heat Light weight, Specific

structures, Space structure exchanger panels, Engine parts (Be-Al), stiffness and specific

(Yue et al, 1998; Pitcher et Wind tunnel blades, Spacecraft truss strength at elevated

al, 1998; Bache et al, 1998; structure, Reflectors, Solar panels, temperature, Creep and

Peters, 1998; Hunt, 2000) Camera housing, Hublle space telescope fatigue resistance,

metering truss assembly, Turbine rotor, Controlled CTE,

Turbine wheels (operating above 40,000 Thermal conductivity,

rpm), Nose caps and leading edge of Dimensional stability

missiles and Space shuttle.

Aircraft, Missile structures Wings, Rotary launchers, Engine casing, Stiffness, Reduced

(Pitcher et al, 1998; Rings (Al2O3/Al-alloy), Drive shaft, weight, High

Shakesheff and Purdue, Propeller blades, Landing gear doors, temperature stability

1998; Peters, 1998; Hunt, Thrust reverser (Carbon/Bismaleimide),


16

2000) Helicopter components viz. Rotor drive

shaft, Mast mount, Main rotor blades

(Carbon/Epoxy).

Nuclear Reactor (Hunt, Storage casks for spent fuel rods from Absorption of neutron

2000) nuclear reactors radiation

Sports (Fitzpatrick et al, Tennis rackets, Golf shafts, Racing Increased specific

1998; Hunt, 2000) bicycle frame (SiCw/6061), Fishing rod, stiffness and strength,

Pool cues. durability

2.2 YIELD CRITERIA

2.2.1 Mathematical Formulation

The yield criteria of materials limit the elastic domain during loading,

whereas the failure criteria give the maximum stress that can be applied.

Traditionally, we use the term yield criteria for metals or alloys and failure criteria

for geomaterials such as soil and concrete. Safe and efficient use of materials is

required for the successful design of any structural components. Therefore, for

design purposes, the onset of plastic yielding under loading conditions is of great

importance. The yield criterion gives the onset of plastic deformation. In other

words, if a state of stress satisfies yield criterion, one can say that the plastification

may start. It is assumed that the initial yielding is dependent only on the state of

stress and not on how the stress is reached. It can be assumed that there exists a

yield function f (σ ij ) such that:

Material is elastic if : f (σ ij ) < 0


17

As the yield criterion does not depend on the path of loading, it does not tell

anything about deformation. To develop a yield function, the components of the

multiaxial stress state are combined into a single quantity known as the effective

stress (σ e ) . The effective stress is then compared with the yield stress in some

appropriate form to check the inset of yielding.

The state of stress can be determined by specifying the principal stress and

orientation of principal axes. The three principal stresses and three angles

constitute the six dimensional spaces. Thus, the yield function can be written as,

f = f (σ 1 , σ 2 , σ 3 , α 1 , α 2 , α 3 )

where σ 1 , σ 2 , σ 3 are the principal stresses and α 1 , α 2 , α 3 are the orientations of

principal axes.

The yield criterion must be consistent with a number of experimental

observations, the chief of which is that pure hydrostatic pressure should not cause

yielding in a continuous solid (Dieter, 1988).

2.2.2 Geometrical representation

There exist three principal stresses and their orientation for a given state of

stress at a point. Hence one can represent the state of stress as a point in three

dimensional vector space whose bases are the principal stresses for isotropic

materials, where the orientation of the axes is not important. For isotropic materials

f (σ 1 , σ 2 , σ 3 ) = 0 represents a surface in the principal stress space. If a material is

acted upon by deviatoric stress, then the yield surface can be represented as,

f (σ 1 , σ 2 , σ 3 ) = 0 (2.1)

Subject to the condition that,


18

σ1 + σ 2 +σ 3 = 0 (2.2)

Here, σ 1 + σ 2 + σ 3 = 0 is a plane passing through the origin and equally inclined to

the principal axes. This plane is called the π - plane. Equations (2.1) and (2.2)

represent a curve C on σ 1 + σ 2 + σ 3 = 0 plane (refer Fig. 2.1). If the yield is

independent of hydrostatic stress, then the yield function is a right cylinder with

generator perpendicular to the π - plane and whose cross-section from the

hydrostatic line is the same as curve C.

2.2.3 Yield Criteria for Isotropic Material

For an isotropic material, the yield function must be independent of

orientation of principal axes i.e. α 1 , α 2 and α 3 . Thus, in this case the yield function

f is a function of σ 1 , σ 2 and σ 3 alone i.e,

f = f (σ 1 , σ 2 , σ 3 )

It was proposed by von Mises (1913) that the yielding would occur in an

isotropic material when the second invariant of stress deviator J2 exceeds some

critical value.

J2 =k 2 (2.3)

where,

J2 =
1
6
[
(σ 1 − σ 2 ) 2 + (σ 2 − σ 3 ) 2 + (σ 3 − σ 1 ) 2 ]
To estimate the value of constant k and to relate it with the yielding under tension

test, it is realized that at the onset of yielding under uniaxial

tension: σ 1 = σ e , σ 1 = σ 2 = 0 . Therefore, from Eqn. (2.3), we get,


19

σ e2 + σ e2 = 6k 2

or, σe = 3k (2.4)

Using Eqn. (2.4) into Eqn. (2.3), we obtain the following usual form of von Mises

yield criterion,

σe =
1
[(σ 1 − σ 2 ) 2 + (σ 2 − σ 3 ) 2 + (σ 3 − σ 1 ) 2 ]
1/ 2
(2.5)
2

where σ e is the equivalent stress.

Geometrically, von Mises yield criterion represents right circular cylinder

whose generator is equally inclined to the principal stress axes, as shown in Fig.

2.2. The von-Mises criterion implies that the yielding under both uniaxial tension

and compression would start at the same value of tensile and compressive stresses.

2.2.4 Yielding in Presence of Residual Stress

Arsenault and Taya (1987) pointed out that even in an isotropic metal

matrix composite yielding does not begins at the same level of tensile and

compressive stresses under uniaxial loading. Badini (1990) also noticed that the

compressive yield strength of 15 vol% SiCw/6061Al composite is higher than its

yield strength in tension. The processing of metal matrix composites often involves

cooling from the higher temperature, which results in residual tensile thermal stress

in the matrix due to restraint caused by ceramic reinforcements. To consider the

effect of different yield stresses in compression and tension, some researchers

(Schellekens and De Borst, 1990a, b; Bicanic et al, 1994; Moin, 1996) employed

the following isotropic form of Hoffman yield criterion,

(σ 12 + σ 22 + σ 32 ) − (σ 1σ 2 + σ 2σ 3 + σ 3σ 1 ) + Q(σ 1 + σ 2 + σ 3 ) − 1 = 0 (2.6)
20

where Q is a material parameter.

A number of other investigators (Bicanic et al, 1994; Moin, 1996) preferred

the following alternate form of Hoffman’s yield criterion that employs uniaxial

compression and tensile yield stresses denoted respectively by f c and f t ,

(σ 12 + σ 22 + σ 32 ) − (σ 1σ 2 + σ 2σ 3 + σ 3σ 1 ) + ( f c − f t )(σ 1 + σ 2 + σ 3 ) − f c f t = 0

(2.7)

The mechanical strength of whisker reinforced composites depends on the

orientation of whiskers in the matrix (Ledrich and Shastry, 1982; Crowe et al,

1985; McDenels, 1985). Badini (1990) observed that the compressive yield

strength of 15 vol% SiCw/6061Al composite in the longitudinal direction is higher

than its yield strength in the transverse direction. Moreover, fiber-reinforced

composites or components that are fabricated by processes such as forging, rolling

or extrusion exhibit anisotropic properties. Therefore, the use of von-Mises yield

criterion given by Eqn. (2.5) is not an appropriate choice for such cases.

The Eqn. (2.7) is used conveniently to describe the yielding of materials,

which show different yield stresses under tension and compression. Several

investigators (Arsenault and Taya, 1987; Shi et al, 1992) reported that thermal

residual stresses generally develop during cooling of composites from high

processing temperature. As a result of thermal residual stress, the yield strength of

composites is different when their direction of loading is reversed during uniaxial

loading. The magnitude of change in yield strength ( Δσ y ) between compression

and tension is dependent on the test temperature, as reported in Table 2.2 for 20

vol% SiCw/Al composite.


21

Table 2.2: Effect of processing temperature on yield strength of 20 vol%

SiCw/Al (Shi et al, 1992)

Processing Tensile Yield Stress Compressive Yield


Temperature (MPa) Stress (MPa)
(0C)
0 39.7 39.7
480 79.8 87.7

The yield stresses reported in the table above are deduced from the stress-

strain curve (tension/compression) of SiCw/Al composite with thermal history (i.e.

without cooling and cooling from 480 0C).

Badini (1990) studied the correlation between the microstructure and the

tensile and compressive properties of extruded bars made of 6061 Al alloy matrix

composite reinforced with silicon carbide whiskers. The compressive strength in

the longitudinal direction was observed to be considerably higher than the strength

in the transverse direction. The compressive and tensile strengths of cylindrical

samples, having 20 mm diameter, was found to be 218 MPa and 186 MPa

respectively. Further, the mechanical strength of material was observed to depend

on the loading direction, due to orientation of whiskers that contribute to

reinforcing the composite to different degrees depending on their alignment in the

direction of load applied (Lederich and Sastry, 1982; Crowe et al, 1985; McDanels,

1985).
22

2.3 CREEP

Apart from elastic and plastic deformations, which appear almost

instantaneously with the application of load, there is another mode of deformation,

which is a function of time and increases continuously even under constant load.

This time dependent deformation, which takes place at higher temperatures,

generally above 0.5 Tm, results in permanent deformation even when the stress is

well below the yield stress. This deformation in metals or metal matrix composites

is marked by nonlinear constitutive equations and the visco-elastic strain rate

depends on the effective stress through a power law involving material parameters,

which may also be a function of plastic strain.

Creep is the progressive time-dependent inelastic deformation under

constant load and temperature. Due to creep, a structural component undergoes

time-dependent changes in the state of stress and strain such as progressive

deformations, relaxation and redistribution of stresses, local reduction of material

strength, change of material behavior from isotropic to anisotropic etc. Kraus

(1980) pointed out that the creep behavior is a function of stress, temperature, time,

stress history, temperature history and type of material. Creep behavior also

includes the phenomenon of relaxation, which is the reduction of stress in a

structure with time while the total strain remains constant. Further, it also includes

the recovery, which is characterized by the reduction of inelastic strain with time

after the stress has been removed.

The phenomenon of creep is observed in most of the materials. The

operating temperature in various industrial and structural applications is


23

sufficiently high to cause significant creep as observed in chemical industries,

nuclear power plants, missiles, aeroengines, gas turbines etc. The testing and

estimation of creep in composite materials is tedious, costly as well as time

consuming. Therefore, the prediction and analysis of creep is extremely important

for estimating the life of components made of composite materials subjected to

severe thermo-mechanical loadings. Accordingly, it is imperative to develop a

thorough understanding of the methods for analyzing stresses and deformations in

the components subjected to creep.

The increase in stress and temperature increases the creep rate and reduces

the rupture time. Creep is generally viewed as a thermally driven process that is

influenced by the presence of stress. Several creep mechanisms like dislocation

glide, bulk diffusion, grain boundary diffusion and dislocation creep have been

conceived. More than one mechanism could operate at a given temperature. Ashby

and Frost (1975) developed a convenient summary of the various active creep

mechanisms as a function of stress and temperature in the form of deformation

maps shown in Fig. 2.3. The deformation maps depend also on the microstructure

of material. Increase in grain size reduces the amount of grain boundary diffusion

and hence reduces the creep.

2.3.1 The Creep Curve

To obtain the engineering creep curve of a metal, a standard cylindrical

specimen is heated to a constant temperature (T) of the order of 0.3 to 0.5Tm (Tm is

the melting temperature of the material) and subjected to a constant tensile force F.

The strain developed in the specimen is recorded and represented with respect to
24

time, in order to get creep curve of the material. The normal stress induced in the

specimen is usually much less than the yield limit of the material.

Curve A in Fig. 2.4 illustrates the idealized shape of a creep curve. The

curve exhibits three different regimes of creep distinguished by the nature of

variation of creep strain with time. The slope of curve A, (dε / dt or ε& ) is referred to

as creep rate. The degree to which the three stages of creep are readily

distinguishable depends strongly on the applied stress and temperature. The first

stage of creep, known as primary creep, represents a region of decreasing creep

rate. Primary creep is a period of predominantly transient creep during which the

creep resistance of material increases by virtue of its own deformation. For low

temperatures and stresses, as in the creep of lead at room temperature, primary

creep is the predominant creep process. The second stage of creep, known also as

secondary creep, is a period of nearly constant creep rate, which results from a

balance between the competing processes of strain hardening and recovery. For this

reason, the secondary creep is usually referred to as steady state creep. The average

value of the creep rate during secondary creep is called the minimum creep rate.

Third-stage or tertiary creep mainly occurs in constant load creep tests at high

stresses and high temperatures. Tertiary creep occurs when there is an effective

reduction in cross-sectional area either because of necking or internal void

formation. Third stage creep is often associated with metallurgical changes such as

coarsening of precipitate particles, recrystallization or diffussional changes in the

phases that are present. The dashed line (curve B) in Fig. 2.4 shows the shape of a

constant-stress creep curve. When the constant-stress tests are conducted, it is


25

found that the onset of stage III is greatly delayed. In engineering situations it is

usually the load not the stress that is maintained constant, so a constant-load creep

test is more important.

Andrade’s (1957) showed that the constant stress-creep curve represents the

superposition of two separate creep processes, which occur after the sudden

straining due to application of load. The first component of the creep curve is a

transient creep in which the creep rate decreases with time. Added to this is a

constant-rate viscous creep component. The superposition of these creep processes

is shown in Fig. 2.5. Andrade observed that a creep curve can be represented by the

following empirical equation:

ε = ε 0 (1 + β 1t 1 / 3 ) e kt (2.8)

where ε 0 is the instantaneous strain, ε is the strain in time t, and β 1 and k are

constants.

To analyze creep in engineering structures and components, mathematical

models of creep are needed. For this purpose, the data obtained from the tensile

tests serve as the main source of information about creep.

2.3.2 Structural Changes during Creep

If the slope of a creep curve (Fig. 2.4) is plotted versus strain, a curve of

creep rate vs total strain is obtained (Fig. 2.6). This curve illustrates the large

change in creep rate, which occurs during the creep test. Since the stress and

temperature are constant, this variation in creep rate is the result of changes in the

internal structure of the material with creep strain and time.


26

The principal deformation processes at elevated temperature are slip,

subgrain formation and grain-boundary sliding. High-temperature deformation is

characterized as extreme inhomogenity. Measurements of local creep elongation at

various locations in a creep specimen have shown that the local strain undergoes

many periodic changes with time that are not reflected while recording the changes

in strain over total gage length of the specimen. In large grained specimens, local

regions may undergo lattice rotations, which produce areas of misorientation.

2.3.3 Creep Models

Most of the models representing uniaxial creep of metals and alloys are

empirical, established after years of experimentation. These models have been

widely used and implemented as options in several finite element codes. The

models express uniaxial creep strain or creep strain rate as a function of stress, time

and temperature, and some models also use the accumulated creep strain to model

strain hardening. The mathematical form of these models is as given below,

ε = f (σ , t , T ) (2.9)

where ε is the creep strain, σ is the applied stress, t is the time period and T is the

temperature of application.

The Eqn. (2.9) can also be written in the following form,

ε = f 1 (σ ). f 2 (t ). f 3 (T ) (2.10)

Since the temperature during creep test is constant, therefore, Eqn. 2.10

becomes,

ε = f1 (σ ). f 2 (t ) (2.11)
27

Many empirical expressions, as a function of f1 (σ ) and f 2 (t ) , have been

developed in the past. The most commonly used stress function law is Norton’s law

given as,

f1 (σ ) = Bσ n

where B and n are the material constants. The approximation of a creep curve to a

straight line is possible if secondary creep region is predominant, and elastic as

well as primary creep is negligible. Under such conditions, it can be assumed that

the creep strain rate depends upon stress function only. The time dependence of

creep has been expressed in terms of Bailey’s empirical law given by,

f 2 (t ) = A1t m

where A1 and m are the material constants.

For practical applications creep does not usually occur at a fixed

temperature and stress over the entire life of component. Therefore, it is not

possible to compute the creep strain directly from a uniaxial creep equation. The

rules of time hardening and strain hardening have been developed to extend the

constant stress and temperature creep equations to time-varying stress and

temperature histories. The creep strain is determined by integrating the creep strain

rate equations for changing stress and temperature conditions. To compute the

creep strain rate at a particular instant, the mechanical and thermal loads as well as

the material history must be known. Hardening rules specify the state of material as

a function of loading history. In its simplest form, a creep-hardening rule might be

looked upon as a method of moving between constant stress and temperature creep

curves as the stress and temperature change. Two different hardening rules have
28

been used with classical creep models: time hardening and strain hardening. When

the creep rate equations are integrated for conditions of constant stress and

temperature, both the hardening rules produce identical results, namely the uniaxial

creep curve for that stress and temperature.

The time hardening theory states that for a constant temperature and

variable stress condition, creep rate ( ε&c ) is a function of stress and time i.e.,

ε&c = f (σ , t ) (2.12)

However, in case of strain hardening theory it is assumed that the creep rate

is a function of stress and accumulated strain i.e.,

ε&c = f (σ , ε ) (2.11)

The particular forms of these laws can be obtained by assuming that the

creep curve can be represented by Bailey-Norton law, which is a common

representation of creep in the primary and secondary creep ranges under isothermal

conditions and is given below,

ε c = Aσ n t m (2.12)

where A, m and n are constants whose values depend upon the type of material. The

constant A, and exponents m (>1) and n (<1), should be functions of temperature.

The form of Bailey Norton given by Eqn. (2.12) could be used to describe creep in

both tension and compression, but it would be reasonable to expect that the

parameters would be different in tension and compression. Since the value of

exponent n < 1 , therefore the creep rate at t = 0 could not be defined.

Differentiating Eqn. (2.12) with respect to time, the time hardening law can

be obtained as,
29

dε c
ε&c = = Amσ n t m −1 (2.13)
dt

From the above equation it can be seen that the creep rate decreases with

time since 0 < m < 1. Further, Eqn. (2.13) can also be written in the following form,

independent of time t, by eliminating t between Eqs. (2.12) and (2.13).

mA (1 / m )σ ( n / m )
ε&c = (2.14)
ε c(1− m ) / m

Eqn. (2.14) indicates that the creep strain rate decreases with increasing creep

strain ( ε c ) i.e. with the progression of creep strain the material hardens.

Though, both the laws are derived from the same equation, but it is

observed that for varying stress, the time and strain hardening laws give different

creep rates. This difference is procedural and not phenomenological. Quite often

the strain hardening models give more accurate predictions of experimental results

for stepwise changes of stress. Unfortunately, strain hardening models do not

always yield accurate predictions, particularly when several step changes in stress

occur during the same test (Rabotnov, 1969). Pickel et al (1971) also noticed that

the strain hardening model is unable to accurately predict the creep behavior

resulting from structural instabilities. But for structurally stable materials, the

predictions by strain hardening model are fairly reliable. However, in the case of

gradually varying stress, both the laws give approximately similar predictions.

2.3.4 Creep under Multiaxial Stress

The uniaxial creep tests allow us to establish the basic features of creep

behavior and to establish the relationship between stress, strain rate, temperature,

time, etc. However, in actual practice, most of the structural components operate
30

under multiaxial stress conditions. Metallurgical models used to describe creep

under multiaxial stress creep are limited and the available models of multiaxial

creep are mainly phenomenological in form (Kraus, 1980; Boyle and Spence, 1983;

Gooch and How, 1986). Therefore, in order to analyze the influence of stress state

on the time dependent material behavior, multiaxial creep analysis is required.

Finnie and Hellar (1959) proposed the generalized constitutive equations to

describe creep in isotropic materials under the influence of multiaxial stress as

given below,

ε&e ⎡ 1 ⎤
ε& x = ⎢σ x − 2 (σ y + σ z )⎥
σe ⎣ ⎦

ε&e ⎡ 1 ⎤
ε& y = ⎢⎣σ y − 2 (σ z + σ x )⎥⎦
σe

ε&e ⎡ 1 ⎤
ε& z = ⎢σ − (σ x + σ y )⎥
σe
z
⎣ 2 ⎦

3ε&e
ε& zx = τ zx
σe

3ε&e
ε& xy = τ xy
σe

3ε&e
ε& yz = τ yz (2.15)
σe

where σ x , σ y , σ z and τ xy , τ yz , τ zx are respectively the normal and shear stress

components in x, y and z directions respectively. Similarly ε& x , ε& y , ε& z and

ε& xy , ε& yz , ε& zx are respectively the normal and shear strain rate components in x, y and
31

z directions respectively. The effective stress ( σ e ) and the effective strain rate ( ε&e )

given above are expressed by,

σe =
1
[(σ y − σ z ) 2 + (σ z − σ x ) 2 + (σ x − σ y ) 2 + 6(τ yz2 + τ zx2 + τ xy2 ) ] 1/ 2
(2.16)
2

ε&e =
1
[(ε& y − ε& z ) 2 + (ε& z − ε& x ) 2 + (ε& x − ε& y ) 2 + 6(ε& yz2 + ε& zx2 + ε& xy2 ) ]
1/ 2
(2.17)
2

The Eqs. (2.15) - (2.17), are based on the assumption that the material is

isotropic.

2.3.5 Activation Energy for Steady State Creep

The steady state creep dominates at temperatures above about 0.5Tm. The

simplest assumption that the creep is a thermally activated process, which can be

expressed by the following Arhenius type rate equation,

ε& s = Ae − Q / RT

where the symbols are defined below:

ε&s : the steady state creep rate

Q : the activation energy for the rate controlling process

A : the pre-exponential complex constant containing the frequency of vibration of

the flow unit, the entropy change and the factor that depends on the structure

of the material.

T : the absolute temperature

R : the universal gas constant

A temperature differential creep test is often used to measure the activation

energy for creep. If the temperature interval is small such that the creep mechanism

would not be expected to change, one can write,


32

A = ε&1 e Q / RT1 = ε& 2 e Q / RT2

and,

R ln (ε&1 / ε& 2 )
Q=
(1 / T2 − 1 / T1 )

An extensive correlation between creep and diffusion data for pure metals

(refer Fig. 2.7) indicates that the activation energy for high temperature creep is

equal to the activation energy for self diffusion. The activation energy for self

diffusion is the sum of energies for the formation and movement of vacancies,

which strongly supports the view that dislocation climb is the rate controlling step

in high temperature creep. The formation of a dislocation subgrain structure is

another factor in support of this view. Therefore, it is expected that the metals in

which the vacancies move rapidly would have better creep resistance.

2.3.6 Creep Law for Aluminum based MMCs

In aluminum based composites, undergoing steady state creep, the effective

creep rate ( ε&e ) is related to the effective stress ( σ e ) through the following well

documented threshold stress ( σ o ) based creep law (Park et al, 1990; Mishra and

Pandey, 1990; Mohamed et al, 1992; Pandey et al, 1992; Gonzalez and Sherby,

1993; Pandey et al, 1994; Park and Mohamed, 1995; Cadek et al, 1995; Li and

Mohamed, 1997; Li and Langdon, 1997a, 1999a; Yoshioka et al, 1998; Tjong and

Ma, 2000; Ma and Tjong, 2001),

n
⎛σ −σo ⎞ ⎛−Q⎞
ε&e = A ⎜ e
'
⎟ exp⎜ ⎟ (2.18)
⎝ E ⎠ ⎝ RT ⎠
33

where the symbols A’, n, Q, E, R and T denote respectively the structure dependent

parameter, true stress exponent, true activation energy, temperature-dependent

Young’s modulus, gas constant and operating temperature.

The threshold stress based creep law given by Eqn. (2.18) may alternatively

be expressed as,

ε&e = [M (σ e − σ o )] n (2.19)

1/ n
1⎛ −Q⎞
where, M = ⎜ A ' exp ⎟
E⎝ RT ⎠

The creep parameters M and σ o appearing in Eqn. (2.19) are dependent on

the type of material and operating temperature (T). In a composite, the dispersoid

size (P) and its content (V) are the primary variables affecting these parameters.

Therefore, these parameters are functions of dispersoid size (P), volume content of

the dispersoid (V) and operating temperature (T). But the functional relations,

describing dependence of M and σ o on P, V and T, are scant. The values of M and

σ o can be extracted from the experimental creep results reported for aluminum or

its alloy based composites under uniaxial loading.

2.3.7 Estimation of Threshold Stress

It is evident from Fig. 2.8 that the variation of creep rate with applied stress

for the composites generally exhibits curvature at higher creep rates when creep

strain rates are measured over more than five orders of magnitude. The apparent

stress exponent, na ( = ln ε& / ln σ ), decreases with increasing stress, which is usually

considered to be an indicator for the presence of threshold stress, below which

creep does not occur. This similarity indicates that the creep behavior of these
34

composites could be explained in terms of the existence of a threshold stress for

creep (Mohamed, 1998) and the creep behavior of these composites may be

described by modified power law as given by Eqn. (2.18).

Nardone and Strife (1987) were the first to introduce the threshold stress

into power law creep equation, to explain the creep data obtained for 20 vol%

SiCw/2124Al composite under heat treated condition. Similarly, Park et al (1990)

and Pandey et al (1992) also considered the threshold stress while analyzing creep

data of SiCp/Al composites. Using threshold stress based approach, one could

easily explain the high values of apparent stress exponent and activation energy. In

an overview, Cadek et al (1995) pointed that the threshold creep behavior is

inherent for discontinuously-reinforced aluminum matrix composites. Accordingly,

the creep behavior of these composites was rationalized by using the threshold

stress approach used by numerous workers (Nardone and Strife, 1987; Park et al,

1990, 1994; Pandey et al, 1992; Mohamed, et al, 1992; Gonzalez and Sherby,

1993; Cadek et al, 1994, 1995, 1998a, b; Zhu et al, 1996; Li and Mohamed, 1997;

Li and Langdon, 1997a, 1998a, b; Tjong and Ma, 1999a; Ma et al, 1999;

Wakashima et al, 2000; Ma and Tjong, 2000; Lin et al, 2002).

The following three different methods are commonly employed by various

workers to estimate the magnitude of threshold stress. In the first method, as

proposed by Lagneborg and Bergman (1976), the threshold stress is estimated by

using linear extrapolation of the strain rate versus stress plot in which the creep

data at a single temperature is represented on ε&1 / n versus σ plot on linear scales.

Further, it is assumed that if the creep data of composites satisfy Eqn. (2.18), and
35

σ 0 is independent of the applied stress, the data points in ε&1 / n versus σ plot would

fit in a straight line. This line is extrapolated to zero strain rate to get the value of

threshold stress ( σ 0 ) at the given temperature. The most appropriate value of true

stress exponent, n, is obtained by constructing several ε&1 / n versus σ plots of creep

data by using different values of n, and selecting the value of n that describes the

best linear fit of the data points.

According to second approach, if the creep behavior of a material could be

described by modified power law given by Eqn. (2.18), the relation between the

apparent stress exponent na, true stress exponent n, applied stress σ and threshold

stress σ 0 is given by (Chaudhury and Mohamed, 1988),


na = (2.20)
(σ − σ 0 )

The values of n and σ 0 corresponding to each test temperature could be

estimated from the data in the plot of apparent stress exponent na versus applied

stress σ , resulting from experimentally determined ε& (T , σ ) relationships and by

applying Eqn. (2.20) to such a plot.

A difficulty associated with the linear extrapolation method is that it

requires a prior selection of appropriate values for n. In practice, the analysis is

generally undertaken by using either integer values of n or some specific values of

n which are related to some other well documented creep process. However, this

problem could be avoided by using the threshold stress approach suggested by Li

and Langdon (1997c). When the creep data covers at least 5 orders of magnitude of
36

strain rates and includes the strain rate as low as ~10-8 s-1, the curves through the

individual experimental points generally could be extrapolated to lie essentially

vertical at a strain rate of ~10-10 s-1. Accordingly, the predicted stress level at this

low strain rate is very close to the threshold stress. Thus, using the values of

threshold stress, estimated at a strain rate of ~10-10 s-1, the creep data could be

replotted on the logarithmic scale between the measured strain rate against the

effective stress, ( σ − σ 0 ) or ( τ − τ 0 ), where σ , σ 0 , τ and τ 0 indicate the applied

normal stress, threshold stress under normal loading, applied shear stress and

threshold stress in shear respectively. The slope of logarithmic plots of strain rates

versus effective stress yields the value of true stress exponent n.

2.3.8 Characteristics and Origin of Threshold Stress

Several investigators have proposed plausible explanations for the

characteristics and origin of threshold stress while describing creep behavior of

composites. Pandey et al (1992) examined the effects of particulate size and

content on the creep behavior of SiCp/Al composites processed through powder

metallurgy (PM) route. It is observed that the values of threshold stress

estimated from ε& 1 / 8 versus σ plot by using linear extrapolation technique are

primarily dependent upon the volume fraction of the reinforcement but are

independent of the test temperature. Based on their experimental results, it is

suggested that the load transfer to the reinforcement may be responsible for the

origin of threshold stress. However, this approach fails to account for the

experimental observation that the unreinforced matrix alloy also exhibits

essentially similar threshold stress during creep, however there is no possibility


37

of load transfer during creep of unreinforced aluminum alloy (Park and

Mohamed, 1995).

On the basis of examination of creep data of several discontinuous SiC/Al

composites, based on the value of true stress exponent (n = 8), Gonzalez and

Sherby (1993) reported that the threshold stress exhibits two main characteristics.

Firstly, it depends linearly on temperature and becomes zero in the temperature

range between 733 K and 743 K. Secondly, the threshold stress is higher for

whisker reinforced composites than the particulate composites. Gonzalez and

Sherby (1993) concluded that the threshold stress in discontinuous SiC/Al

composites originates due to the presence of SiC particles or whiskers.

Cadek et al (1998) investigated the high temperature creep behavior of 20

vol% SiCp/2124Al composite at temperatures between 623 K and 748 K. The plot

of shear creep rate versus applied shear stress ( ln γ& m vs ln τ ) indicates the origin of

threshold stress with the value of apparent stress exponent, na = ( ln γ& m / ln τ )T,

which is observed to decrease with increasing applied stress. By considering

threshold stress into analysis, it is observed that the minimum creep rate of the

composite is controlled by the matrix lattice diffusion with a true stress exponent

close to 5. The threshold stress, as estimated by the extrapolation technique,

decreases linearly with increasing temperature and disappears at a temperature near

735 K. This finding is in good agreement with that of Gonzalez and Sherby (1993),

though these workers have introduced the structure invariant model with a stress

exponent of 8.
38

Li and Langdon (1999b) suggested an alternative explanation for the

dependence of threshold stress on the reinforcement content as given by Pandey et

al (1992). Based on the scanning electron microscopic (SEM) observations on 10

vol% SiCp-2124 Al composite, it is reported that the reinforcing particles fracture

into fine SiC particles, with sizes smaller than 100 nm, as a result of thermo-

mechanical processing. These fine SiC particles, like oxide particles, partly account

for the origin of threshold stress. Moreover, it is reasonable to assume that high

volume fraction of reinforcement gives rise to a larger number of fine SiC particles.

Thus, the increase in threshold stress is expected with the increase in volume

fraction of the reinforcement.

Lin et al (2002) investigated the creep behavior of 5 vol% SiCp/2124Al

composite in double shear configuration in the temperature range between 618 K

and 678 K, and compared the results obtained with those noticed for unreinforced

2124Al, tested under similar experimental conditions. The study indicates that the

origin of anomalous stress dependence of the creep rate of the composite and the

unreinforced matrix alloy is due to the presence of threshold stress, τ 0 , which

strongly depends on temperature. An examination of the substructure of the creep

samples suggests that the most probable source of threshold stress in the composite

as well as 2124Al is related to the interaction between moving dislocations and

dispersed particles, which includes those formed during the processing of these

materials through powder metallurgy route. The creep behavior of the composite is

governed by the deformation of matrix.


39

2.3.9 True Stress Exponent

During analysis of creep data, the value of true stress exponent (n)

appearing in Eqn. (2.18) is usually taken as 3, 5 and 8, which corresponds to three

well-documented creep cases for metals and alloys: (i) n = 3 for creep controlled by

viscous glide processes of dislocation, (ii) n = 5 for creep controlled by high

temperature dislocation climb (lattice diffusion), and (iii) n = 8 for lattice diffusion-

controlled creep with a constant structure (Tjong and Ma, 2000).

Pandey and coworkers investigated high-temperature creep behavior of

SiCp/Al (Pandey et al, 1990, 1992) and TiB2p/Al (Pandey et al, 1994) composites.

The creep data obtained has been explained by using the substructure invariant

model proposed by Sherby et al (1977). Pandey et al (1992, 1994) argued that

during creep in these composites the subgrain boundaries are pinned by reinforced

particles like Al2O3 and TiB2, thereby yielding a stress exponent of 8 and an

activation energy of the order of lattice self diffusion.

The substructure invariant model assumes that the subgrain structure

remains stable during extended creep exposure and such a structure is insensitive to

stress. This is achieved by the introduction of second phase particles such as oxide

dispersoids in metals. These particles are observed to be located at the subgrain

boundaries and the subgrain size becomes equal to interparticle spacing. In such a

situation, particles stabilize the subgrain size. The substructure invariant model

predicts that the creep rate is proportional to the cube of subgrain size and a stress

exponent of 8 for particle reinforced materials that exhibit invariant microstructure


40

with stress. The substructure invariant model (Sherby et al, 1977) is described by

the following equation,

3
⎛ D ⎞ ⎛ λ ⎞ ⎛ σ − σ ⎞8
ε = S ⎜ L2 ⎟ ⎜⎜ ⎟⎟ ⎜
& 0
⎟ (2.21)
⎜ b ⎟ br ⎝ E ⎠
⎝ r ⎠⎝ ⎠

where ε& is the strain rate, σ is the applied stress, DL is the lattice self-diffusion

coefficient, λ is the subgrain size, br is the Burgers vector, E is the Young’s

modulus and S is a constant.

Gonzalez and Sherby (1993) reanalyzed the creep data of a series of SiC/Al

composites and demonstrated that the creep in SiC/Al composites may be described

by using the substructure invariant model. They also pointed out that the barrier

spacing plays a key role during creep of SiC/Al composites.

Mishra and Pandey (1990) analyzed the steady state creep data of SiC/

6061Al composites, as obtained by Nieh (1984), Nieh et al (1988) and Morimoto et

al (1988), and noticed that the substructure invariant model (n = 8) could explain

the entire set of data. It is observed that the basic steady state creep mechanism

remains unaltered for different shapes and volume fractions of SiC reinforcement,

whose influence is confined to the threshold stress.

Ma et al (1999) examined the high temperature creep behavior of in-situ 15

vol% TiB2/Al composites at temperatures between 573 K and 673 K. It is observed

that the stress exponent of 8 gives the best linear fit between ε& 1 / n and σ for this in-

situ composite. However, for n = 3 and 5, the data points of the plots exhibit a clear

curvature at low applied stress. It is further noted that the data points for all the

temperatures may be well fitted by a single straight line with a slope of 8, which
41

implies that the substructure invariant model provides satisfactory explanation for

creep in 15 vol% TiB2/Al composite. Similarly, the creep behavior of 20 vol%

SiCw/Al, 20 vol% (Al2O3+TiB2)/Al and 20 vol% (Al2O3+TiB2)/Al-Cu composites

may also be reasonably explained by means of the substructure invariant model (Ma

and Tjong, 1998; Ma et al, 1999; Tjong et al, 1999b).

Park et al (1990) in their study on creep behavior of 30 vol% SiCp/6061Al

composite demonstrated that a stress exponent of 5 rather than 8 exhibits the best

linear fit between γ&1s / n and τ , where γ& s and τ denote the shear creep rate and the

applied shear stress respectively. Mohamed et al (1992) in their study on similar

composite also revealed that the value of true exponent (n) ≈ 5 and a true

activation energy for creep (Q) ≈ 208 kJ/mol. Therefore, the prevailing creep

mechanism in these composites is climb-controlled dislocation creep corresponding

to stress exponent of 5. Li and Langdon (1998a) also made similar observations.

Cadek et al (1994a, 1995) reanalyzed the creep data of 20 vol%

SiCw/6061Al and 20 vol% SiCp/2124Al (Nieh et al, 1988), 30 vol% SiCp/6061Al

(Park et al, 1990), 20 vol% SiCp/Al (Pandey et al, 1992), 26 vol% Al2O3/Al–

5%Mg (Dragone and Nix, 1992) and 30 vol% SiCp/Al (Cadek et al, 1994b)

composites and concluded that the relationship between ε& 1 / n and σ in these

composites could be linearly described by assuming n = 5. They also observed that

the linear fitting of creep data obtained for SiCp/Al by Pandey et al (1992) could be

achieved satisfactorily for various values of stress exponent (n = 3, 5 and 8), due to

relatively narrow range of experimental creep rates, which are less than four orders

of magnitude.
42

Pandey et al (1996) investigated the creep behavior of 10 vol% SiCp/Al-4%

Mg composite and noticed that the composite exhibits a stress exponent (n) = 8 and

Q ≈ 224kJ / mol at higher stress levels. The activation energy is observed to be

higher than the activation energy for lattice self-diffusion. Li and Langdon (1997b)

during re-appraisal of the same data noticed that n ≈ 3 and Q ≈ 125 kJ / mol , which

is consistent with the creep controlled by a viscous glide process and the dragging

of magnesium atom atmosphere. It is well supported by the fact that the activation

energy for diffusion of Mg in an aluminum lattice is ~130 kJ / mol . The marked

difference noticed between these two analyses is attributed to the difficulties

associated with determining the best value of n for experimental creep data, which

span over a very limited range of strain rates (Li and Langdon, 1999b). The

reported fact is in agreement with the earlier work carried out by Cadek and Sustek

(1994), which suggests that the creep data of MMCs should extend over at least

five orders of magnitude of strain rate to enable an unambiguous determination of

the value of true stress exponent (n).

The study of creep behavior of 6061Al and 7005Al matrix composites

reinforced with 20 vol% of irregular shaped Al2O3 indicates that the true stress

exponent is close to 3 for Al2O3p/6061Al composite after taking threshold stress

into account (Li and Langdon, 1997a, 1998a). In addition, the true activation

energy obtained is close to that observed for diffusion of magnesium in aluminum

matrix (Li and Langdon, 1997a). The results indicate that the creep in 6061Al

matrix composites is controlled by viscous glide process of dislocations. However,

for Al2O3p/7005Al composite, the true stress exponent appears to be ~ 4.4 and the
43

creep is controlled by a dislocation climb process. Therefore, the creep in metal

matrix composites is controlled by creep of matrix alloys and the creep behavior of

composites could be classified into following two different classes: (i) class M

(metal-type) with n ≈ 3 and (ii) class A (alloy-type) with n ≈ 5, as observed in solid

solutions (Li and Langdon, 1998b). Similar observations are also reported by Li

and Langdon (1999a) during reanalysis of the available creep data of 20 vol%

SiCp/2024Al composite (Gonzalez and Sherby, 1993).

It is quite evident from the literature consulted that, though, some of the

researchers (Mishra and Pandey, 1990; Pandey et al, 1992; Gonzalez and Sherby,

1993; Pandey et al, 1994) have used a true stress exponent of 8 to describe the

steady state creep in Al-SiCp,w (subscript ‘p’ for particle and ‘w’ for whisker)

composites but a number of other researchers (Park et al, 1990; Mohamed et al,

1992; Park and Mohamed, 1995; Cadek et al, 1995; Li and Mohamed, 1997; Li

and Langdon, 1997a, 1999a; Yoshioka et al, 1998) have observed that a stress

exponent of either ~3 or ~5, rather than 8, provides a better description of the

steady state creep data noticed for discontinuously reinforced Al-SiC composites.

2.4 FUNCTIONALLY GRADED MATERIALS

Rapid growth in technology has lead to the development of materials for

components that exhibit a graded variation in their properties. Under severe

environments such as high temperature or thermal gradient, the conventional

materials may not survive alone. Functionally graded materials (FGMs) are a new

generation of engineered materials that are gaining interest in recent years. The

concept of FGM was first introduced in 1984 in Japan as ultra light temperature-
44

resistant material for space vehicles (Koizumi, 1993). FGMs also find applications

in structural components operating under extremely high-temperature environments

(Noda et al, 1998; Librescu and Song 2005). In FGMs the volume fraction of two

or more constituent materials is varied continuously as a function of position along

certain dimension(s) of the structure (Suresh and Mortensen, 1998; Reddy, 2000).

Due to graded variation in the content of constituent materials, the properties of

FGMs change smoothly and continously from one surface to the other, thus

eliminating the interface problems and also diminishes the concentration of thermal

stress. As an example, FGMs based on ceramic reinforcement in metal matrix are

able to withstand high-temperature environments due to better thermal resistance

offered by ceramic constituents, while the metal constituents enhance their

mechanical performance and reduce the possibility of catastrophic fracture. The

application of concept of FGMs to Metals Matrix Composites (MMCs) has led to

the development of components designed with the purpose of employing selective

reinforcement in certain regions (refer Fig. 2.9) where enhanced properties like

increased – modulus, strength and wear resistance are required (Jolly, 1990;

Koizumi, 1995, 1997; Hirai, 1996; Takezono et al, 1996; Akira and Watabane,

1997; Pattnayak et al, 2001; Zhu et al, 2001; Kieback et al, 2003). In addition to

excellent mechanical and thermal properties, the FGMs also possess the advantage

of optimizing the use of costly dispersoids, as the composition of dispersoids in

FGMs is varied from the high-temperature region to the low-temperature region.


45

2.4.1 Applications of FGMs

FGMs have great potential for applications in aircraft, space vehicles,

advanced engines and other engineering applications because of their unique

performance, acheived due to spatial tailoring of properties at microscopic level

(Noda and Tsuji, 1990; Nagata et al, 1990; Hashida and Takahashi, 1990;

Nakagaki et al, 1991; Erdogan and Wu, 1992; Kumakawa et al, 1992; Teraki et al,

1992; Arai et al, 1993; Ishizuka and Wakashima, 1994; Finot et al, 1994;

Kumakawa et al, 1994; Jin and Noda, 1994; Noda and Jin, 1994; Fukui and

Bowen, 1994; Blumm et al, 1994; Kawasaki and Watanabe, 1994; Pindera et al,

1994; Rabin and Shiota, 1995; Ho and Lavernia, 1996; Tsuda et al, 1996; Noda et

al, 1998). Aerospace industry extensively uses FGMs, as it is desired that the

materials at the surface of space crafts must withstand temperature as high as 2100

K apart from bearing a temperature difference of 1600 K, which may be easily met

by tailoring of FGMs. Some specific diversified applications of FGMs include

Functionally Graded piezoelectric actuators (Li et al, 2003), heated floor systems

(Takeuch et al, 2003), metal/ceramic armor, prosthesis joint with increased adhesive

strength and reduced pain (Quin and Datta, 2004), Thermal Barrier Coatings (TBCs)

for combustion chambers (Ivosevic et al, 2006), thermal protection systems for

spacecraft, hypersonic and supersonic planes (Leushake et al, 2004), rotating blades

in helicopters and turbomachinary (Oh et al, 2005) and smart structures (Ding et al,

2003).
46

2.4.2 Fabrication of FGMs

The various techniques have been proposed for manufacturing of FGMs

such as electrophoretic deposition (Put et al, 2003; Vanmeensel et al, 2005),

chemical vapor deposition (Kim et al, 2005), spark plasma sintering (Shen and

Nygren, 2002; Tokita, 2003) and centrifugal casting (Biesheuvel and Verweij,

2000; Velhinto, 2003). These methods are used to manufacture FGMs with their

properties varying across the thickness direction. To achieve in-plane variation of

properties in FGMs, ultraviolet irradiation method is used (Lambros et al, 1999).

The reduction of gas consumption and weight of the car are the motivating

forces for growth and innovation in the automotive industry. A significant

reduction in weight can be achieved by producing cylinder liners in Al matrix

composite manufactured by centrifugal casting process. The wear resistance of this

composite, under proper working conditions, is much superior to cast iron, which is

commonly used for the production of liners (Bonollo et al, 2004).

Discontinuously reinforced composites viz. ceramic particles reinforced in

aluminum alloy matrix produced by centrifugal casting can be considered as FGMs

as they possess varying distribution of reinforcement in the radial direction owing

to the effect of centrifugal force. Due to difference in densities of ceramic particles

and aluminum alloy matrix, centrifugal separation occurs and higher density

constituent moves towards the outer zones and vice versa. The concentration

profile of ceramic particles, in the radial direction can be controlled and optimized

by adjusting the parameters such as mould rotation speed, mould temperature,

content and size of ceramic particles, and temperature of molten aluminum. The
47

hardness profile of such FGM is directly proportional to the amount of hard

ceramic particles (Kang and Rohatgi, 1996; Liu et al, 1996; Gao and Wang, 2000).

Bonollo et al (2004) used centrifugal casting technique to manufacture

cylinder liners made of Functionally Graded (FG) Al-SiC and Al-Al2O3 composite.

An attempt was also made to analyze the role of process parameters in centrifugal

casting to optimize the distribution of reinforcement at the inner surface of the

liner. The ideal reinforcement distribution was achieved for some combination of

main process parameters, including casting temperature, mould temperature and

content of reinforcement.

Kiran et al (2009) used Al/17 wt% Si alloy to fabricate and characterize a

FG Al/Si in-situ material. The FGM was fabricated by a specially designed

centrifugal casting process and its microstructural characteristic and hardness

profile were examined.

2.4.3 Creep in FGMs

In the past few years, the elastic stresses in FGM subjected to thermo-

mechanical loading have been analyzed by many researchers (Arai et al, 1990; Fukui

and Yamanaka, 1992; Erdogan and Wu, 1993; Hirano and Teraki, 1993; Obata and

Noda, 1994; Tanigawa, 1995; You et al, 2007; Yang, 1998). However, the studies

pertaining to creep in FGMs are rather scant.

Jackson et al (1999) presented an approach to model and design the

components made of FGMs, so as to fabricate them with local composition control.

This approach is based on subdividing the solid model into sub-regions and
48

associating analytic composition blending functions with each region. These

blending functions defined the composition throughout the model as mixtures of

primary materials available to the Solid Freeform Fabrication (SSF) machine. The

role of design rules restricting the maximum and minimum concentrations has also

been discussed by them.

Zhu and Miller (1999) examined creep behavior of FGM provided with a

thermal barrier coating of Zirconium. In order to examine creep behavior, the

thermal gradient in the FGM was produced by heating the ceramic surface with

laser. The ceramic layer was observed to have primary creep. The time,

temperature and stress dependent deformation results in coating shrinkage in the

loading direction and leads to stress relaxation.

Fukui et al (1995) manufactured metal-intermetallics (Al-Al3Ni) FGM by

centrifugal casting method and conducted experiments to evaluate strength

gradients in FGM by conducting a number of 3-point bend tests. The study reveals

that the three point bending strength of Al-Al3Ni FGM could adopt the two

parameter Weibull distribution. The fracture strength of Al-Al3Ni FGM specimen

decreases with the increase in volume fraction of Al-Al3Ni. The strength of Al-

Al3Ni FGM depends on the cleavage fracture strength of Al3Ni and obeys the law

of mixtures as given below,

σ FGM = (1 − f ) σ Al + fσ Al Ni
3
(2.22)

Sadananda et al (1999) investigated creep properties of FG five layered

MoSi2-Si3N4 composite at 1200 oC under compression. The layers consisted of 0,

20, 40, 60 and 80 wt% of Si3N4 with corresponding decrease in MoSi2 content.
49

Each layer was 2 mm thick and possessed uniform distribution of Si3N4. The study

indicates that the creep rates in a single layer decrease with increasing content of

Si3N4.

Zhai et al (2005) investigated creep in FGMs subjected to high temperature

by using computational micro – mechanical method (CMM). Based on the real

microstructure of FG interlayer with different volume fractions, the emulation

experiment was implemented for the creep test numerically and the creep

parameters were estimated. The numerical results indicate that the creep

phenomenon is obvious not only for the metal-rich interlayers but also for the

ceramic-rich interlayers. The creep strain rate of the ceramic/metal interlayer is

larger than that observed for pure metal under the same load when modulus of the

ceramic phase is lower than the metal phase.

2.5 ROTATING DISK

Rotating disk provides an area of research and studies due to their vast

utilization in rotating machinery viz. steam and gas turbine rotors, turbo generators,

pumps, compressors, flywheels, automotive braking systems, ship propellers and

computer disk drives (Gupta et al, 2005; You et al, 2007; Hojjati and Hassani,

2008). In most of these applications, the disk has to operate at elevated temperature

and is simultaneously subjected to high stresses originating due to disk rotation at

high speed (Farshi and Bidabadi, 2008). As a result of severe mechanical and

thermal loadings, the material of the disk undergoes appreciable creep

deformations, thereby affecting its performance (Laskaj et al, 1999; Farshi et al,

2004; Gupta et al, 2005). For example, in turbine rotor there is always a possibility
50

that heat from the external surface is transmitted to the shaft and then to the

bearings, which may adversely affect the functioning and efficiency of the rotor

(Bayat et al, 2008).

Optimal and more reliable design of rotating disks has long been an

important issue in engineering design. By changing geometrical parameters of the

disk and physical properties of the disk material, optimal and more reliable design

of rotating disk for given operating conditions (i.e. load, speed, operating

temperature) can be achieved. The general parameters varied in the optimization

process are the geometrical parameters like the mean radius and thickness of the

disk, and material properties such as density, elastic modulus and Poisson’s ratio

(Guven and Celik, 2001).

2.5.1 Analysis of Elastic-Plastic Stresses in Rotating Disk

Analysis pertaining to estimation of stresses and strain rates in thin rotating

disk can be found in most of the standard text books and literature (Malkin, 1934;

Finnie and Heller, 1959; Lubhan and Felger, 1961; Odqvist, 1974; Findely et al,

1976; Kraus, 1980; Boyle and Spance, 1983; Skrzypek and Hetnarski, 1993;

Nabarro and Villiers, 1995; Penny and Mariott, 1995). Timoshenko and Goodier

(1970) was the first to obtain closed form solutions for rotating homogeneous disks

but without considering temperature gradient. Reddy and Srinath (1974) and Chang

(1976) investigated the influence of material density on stresses and displacements

in a rotating disk made of orthotropic material. It is demonstrated that the existence

of density gradient in the disk significantly affects the distribution of stresses and

displacements. Zhou and Ogata (2002) obtained closed form solutions for rotating
51

solid disk made of cubic anisotropic material by using direct displacement method.

The displacement, strain and stress were expressed as a simple function of polar

coordinates. Orcan and Eraslan (2002) obtained analytical solution for elastic-

plastic deformation in a variable thickness rotating disk in the form of power

function. The analysis assumed the Tresca’s yield criterion and its associated flow

rule, and linear strain hardening material behavior. The solution obtained was

verified by comparing it with the solution available for uniform thickness disk. It is

observed that with the reduction in disk thickness the plastic limit angular velocity

increases and the magnitudes of stresses and deformations reduce.

In the design and manufacturing of gas turbine rotor, optimization of weight

is an important parameter. Several numerical techniques have been proposed for

optimizing the profile of rotating disk (Fox, 1970; Zienkiewicz and Campbell,

1973; Malkov and Salganskaya, 1976; Pederson, 1981; Wang and Gallagher, 1985;

Vanderplaats, 1990; Cheu, 1990). In turbo jet engines, rotating disk is

simultaneously subjected to mechanical and thermal loads. The disk may also be

subjected to internal pressure due to shrink fitting on a shaft. In addition, the blade

effects may also be modeled by applying external tensile load at the outer radius of

the disk. When the disk rotates at significant angular velocity, while the gases cross

through fins, there is a resulting temperature gradient imposed on the disk.

Kollman (1978, 1981, 1984) solved the problem of shrink fitted disk, both

rotating as well as non-rotating, using Tresca’s yield criterion. Yeh and Han (1994)

proposed a symmetric formulation to predict elastic stresses in an inhomogeneous

rotating disk with arbitrary thickness and operating under thermal loading.
52

Gas turbine disks mostly operates under high temperature gradients and is

also subjected to high angular velocity. High speed results in large centrifugal

forces in the disk and simultaneous presence of high temperature reduces the

strength of disk material, which may ultimately result in increased deformation of

the disk. In order to attain accurate and reliable analysis of stress distribution in the

disk, the solution should consider changes in the material properties caused by

temperature change. To achieve this goal, Farshi et al (2004) considered an

inhomogeneous disk model with variable thickness. Using the variable material

properties method (Jahed and Dubey, 1997; Jahed and Sherkatti, 2000; Jahed and

Shirazi, 2001), stresses were obtained in a rotating disk operating under steady

temperature field. The stress calculation was followed by optimization process

based on inscribed hypersphere method. In the optimization process, the objective

function was the total weight of the disk and the constraints were imposed on the

stresses, which were kept less than the yield stress of the material. The disk profile

was optimized and the final solutions were obtained. The study indicates that the

solutions of optimization process for different initial profiles of the disk with

similar specifications are unique when the inscribed hypersphere radii in the last

solution stages are equal. Results obtained were compared with the published data

and were found to be in good agreement.

Farshi and Bidabadi (2008) pointed that the rotating disks are subjected to

secondary creep effects during most of their useful lives, thus it is important that

they should be optimized for minimum weight while considering only the steady-

state creep stresses. They assumed variable physical properties of the disk material,

which was assumed to operate under a high temperature gradient. The study
53

proposed a procedure for weight minimization during steady state creep. The

method estimates the disk thickness profile by keeping its weight to be minimum

while the equivalent secondary creep stresses in the rotating disk at all points

simultaneously approach, but do not exceed, the allowable stress.

Hojjati and Hassani (2008) used theoretical and numerical methods to

analyze stresses and strains in rotating disks having nonuniform thickness and

density. They assumed material of the disk to be elastic–linearly hardening. The

theoretical solution employed variable material properties (VMP) theory whereas

the numerical solution was based on the solution of governing differential equation

using Runge–Kutta’s method for elastic and plastic regimes. Finite element

modeling of the problem was also carried out by using commercially available

software. The results obtained by these three methods generally show good

agreement. The study also reveals that the VMP method is a reliable means for the

design of complex disks when exact solutions are not available.

Hasan (2007) carried out analysis of elastic–plastic stresses in a curvilinear

orthotropic rotating annular disk by considering strain – hardening material

behavior. Radial and circumferential stress components were estimated for

different angular velocity of the disk. It is observed that the magnitudes of

circumferential stress components are higher than the radial stress components. The

magnitudes of residual stress component of the circumferential stress and plastic

flow are highest at the inner surface. At all angular velocities, the radial

displacements for both elastic and plastic solutions have higher values at the inner

surface than observed at the outer surface of the disk.


54

The concept of functionally graded materials in which the properties vary

continuously from point to point, may be considered in order to make optimal use

of material and to improve the performance of structural components by suitable

control of stresses and deformations along with controlling the undesirable features

such as internal residual stress (Nagatha and Takahashi, 1995; Williamson et al,

1995). Jain et al (1999) demonstrated that a specific radial variation of elasticity

modulus leads to equal radial and circumferential stresses in the rotating disk.

Horgan and Chan (1999) also reported similar results.

Durodola and Adlington (1997) carried out predictive assessment of the

effects of various forms of gradation of material properties on the elastic stresses

and strains in rotating axisymmetric components such as disks and rotors. The

study assumed isotropic FGMs based on particulate metal-matrix composites.

Durodola and Attia (2000) carried out similar analysis for rotating hollow and solid

disks subjected to centrifugal body load. The disks were assumed to be made of FG

orthotropic materials obtained by non-uniform reinforcement of metal-matrix with

long fibers. Stresses and displacements in the disks were obtained by direct

integration of the governing differential equations and compared with those

obtained by using finite element method. A very good agreement is observed

between these results.

Kordkheili and Naghdabadi (2007) obtained semi-analytical thermo-elastic

solution for hollow and solid rotating axisymmetric disk made of FGM under plane

stress condition.
55

Bayat et al (2007) investigated FG rotating disk subjected to axisymmetric

bending and steady-state thermal loading. The material properties of the disk were

assumed to be graded in the direction of thickness according to a power law

distribution of volume fractions of the constituents. Using small deflection theory,

an exact solution for displacement field was obtained. It is observed that for

particular values of grading index (n) of material properties, the mechanical

response of the FG disk can be smaller than that of a homogeneous disk.

Bayat et al (2008) obtained elastic solutions for FG axisymmetric rotating

disks having variable thickness profile. The material properties and disk thickness

profile were assumed to be represented by separate power-law distributions. The

effects of material grading index and geometry of the disk were investigated on the

stresses and displacements in the disk. It is observed that the maximum radial stress

in the solid FG disk with parabolic thickness profile is not at the centre like

uniform thickness disk. The study reveals that FG disk with parabolic concave or

hyperbolic convergent thickness profile can be more efficient than the uniform

thickness disk.

Sharma and Sahni (2009) used Seth’s transition theory to obtain elastic-

plastic stresses in thin rotating disks made of transversely isotropic and purely

isotropic materials. It is observed that the rotating disk made of transversely

isotropic material yields at a higher angular speed as compared to disk made of

isotropic material. Rotating disk made of transversely isotropic material is observed

to be much safer than the disk made of isotropic material.


56

Asghari and Ghafoori (2010) obtained semi-analytical three-dimensional

elasticity solution for FG hollow and solid rotating disks. The study aimed to

generalize an available two-dimensional plane-stress solution to a three-

dimensional one. It is revealed that for thin disks the two-dimensional solution

provides appropriate results but for thick disks a three-dimensional elasticity

solution should be considered to avoid poor results.

Sharma and Sahni (2011) used transition theory to obtain elastic – plastic

and transitional stresses in a variable thickness rotating disk having inclusion. It is

observed that the rotating disk made of incompressible material with inclusion

requires higher angular speed to yield at the internal surface as compared to disk

made of compressible material. For disk with exponentially varying thickness, high

angular speed is required for initial yielding at the internal surface as compared to

flat disk. It is also concluded that the disk made of isotropic compressible material

is on the safer side of design as compared to disk made of isotropic incompressible

material as it requires higher percentage increase in angular speed to become fully

plastic from its initial yielding.

2.5.2 Analysis of Creep in Rotating Disk

Disks of gas turbines, jet engines, and automotive and aerospace braking

systems usually operate at relatively higher angular speed and high temperature or

thermal gradient. Therefore, the prediction of creep deformations is extremely

important for these applications. Most of the published work on creep in rotating

disk is dedicated to the steady state creep behavior since a major part of the

component’s creep life is spent in the secondary stage.


57

Wahl et al (1954) analyzed steady state creep deformations and stress

distributions in a rotating forged disk made of 12% chromium steel at 1000 0F. The

results obtained were validated experimentally. Theoretical analysis of creep was

carried out using von Mises and Tresca yield criteria, while the creep behavior was

described by power law. The stress distributions obtained using Tresca and Mises

criteria do not differ significantly. The normalized strain curves calculated for

various values of stress exponent (n) corresponding to both the yield criteria are

observed to practically coincide for n = 6 to 9. The study also reveals that the

theoretical and experimental stresses are in better agreement if the Tresca criterion

is used. The creep deformations estimated using Mises theory are found to be quite

low as compared to those obtained experimentally, which may be attributed to

anisotropy of the disk material.

Wahl (1956) derived some formulas for calculating stress distributions in

rotating disks having constant and variable thickness, undergoing steady-state

creep at elevated temperature. The formulas were based on the Tresca’s yield

criterion and the associated flow rule and give reasonable results when compared

with the available experimental data (Wahl et al, 1954). The method proposed was

also applied to calculate the transient change in stress when the stress distribution

changes from an initial to a steady state condition during the starting period.

Wahl (1957) utilized the formulas derived in their previous work (Wahl et

al, 1954) to construct the design charts of stress distribution in constant thickness

disks, undergoing steady state creep, for different values of stress exponent (n) and

diameter ratios. In all the cases, the disks were subjected to a radial peripheral load
58

to simulate the effect of blade loading. The steady state creep rate was expressed as

a product of power function of stress multiplied by a function of time.

Wahl (1958) extended his previous work (Wahl, 1957) on rotating disks

having central holes and undergoing steady state creep to cases where the radial

and tangential stresses are equal over all or a portion of the disk. Both constant and

variable thickness disks were considered and the charts were presented for

determining the ratios of peak stress to average stress for various diameter ratios,

disk contours and stress exponent. The study indicates that the variable thickness

disk has somewhat lower ratios of peak to average stress than those observed for

constant thickness disk.

Ma (1959) presented a mathematical approach to calculate creep

deformations and stress distributions in rotating solid disks of variable thickness

operating at uniform temperature. The approach was based on maximum shear

stress theory (Tresca yield criterion) associated with the Mises flow rule and the

assumption that the tangential stresses in the disk are invariably greater than the

radial stresses, except at the center of the disk. The results obtained using Mises

criterion are found to be in excellent agreement with the available experimental

creep data (Wahl et al, 1954).

Ma (1960) extended his work for variable thickness solid disks, used in gas

turbine and jet engine, operating at uniform temperature. The study used Tresca’s

criterion and its associated flow rule while the steady state condition was described

by exponential creep law. It is revealed that the stress distributions over central
59

portion of variable thickness disk are quite different from those observed in a

constant thickness disk.

Ma (1961; 1964) further extended his analysis for variable thickness disk

operating under variable temperature. The steady state creep was described by

either exponential creep law (Ma, 1961) or power law (Ma, 1964). The study

reveals that the proposed analyses can be used to obtain closed form solutions for

complex disk design problem with great simplicity instead of using tedious

numerical solutions.

Wahl (1963) investigated the effects of initial transient period on the long-

time creep tests of rotating disks by using both time-hardening and strain-hardening

relations. The results obtained were applied to the long-time spin-tests conducted

on steel disks at 1000 0F (Wahl et al, 1954). The study reveals that by considering

the effects of transient period, there is no appreciable impact on the over-all creep

deformation noticed during the spin tests. However, when the creep deformations

are of order of elastic strains it is necessary to include such transient effects.

Arya and Bhatnagar (1979) analyzed the creep stresses and deformations in

rotating disk made of orthotropic materials by assuming the creep rate to be a

function of time. A numerical example was worked out to investigate the effect of

transition (non-steady) creep and orthotropicity of material on stress and strain

distributions in the disk. It is observed that the tangential stress at any radius and

the tangential strain at the inner radius of the disk decrease at all times for an

anisotropic material. The time taken to reach steady state distribution decreases

with increasing anisotropy of the disk material.


60

Bhatnagar et al (1986) carried out analysis of steady state creep in rotating

disks having constant and variable (linear and hyperbolic) thickness. The creep

stresses and strains were obtained for different cases of anisotropy by using

Norton’s power law creep model. The study reveals that the selection of a certain

type of material anisotropy and an optimum disk profile would lead to a better disk

design.

Gupta et al (2000) used Seth’s transition theory to estimate creep stresses

and creep rates in a thin rotating disk having variable thickness and variable

density. The study reveals that a rotating disk with radially varying density and

thickness ratio is on the safer side of design as compared to a flat disk having

variable density.

Singh (2000) investigated the effect of reinforcement-size, content and

operating temperature on the steady state creep behavior of a rotating disk made of

isotropic Al-SiCp composite, creeping according to Norton’s law. The creep

parameters appearing in the creep law were extracted from the available

experimental creep results reported by Pandey et al (1992) for Al-SiCp. The study

reveals that particle size, particle content and operating temperature do not have

significant effect on the distribution of radial stress but their effect on the

distribution of tangential stress is sizable. Both the radial and tangential strain rates

in the disk decrease by several orders of magnitude with decreasing particle size,

increasing particle content and decreasing operating temperature. Gupta et al

(2004a) also made similar observations, however they used Sherby’s law to

describe the steady state creep behavior of the disk material (Al-SiCp). The study
61

also reveals that for a given operating condition, the strain rates in rotating disks

can be controlled by selecting optimum particle content and/or particle size of the

reinforcement.

Singh and Ray (2001; 2003a) analysed steady state creep in a rotating disk

made of isotropic FGM (Al-SiCp) and operating at a constant temperature. The disk

was assumed to have linearly decreasing content of SiCp from the inner to outer

radius and creeping according to Norton’s power law. It is observed that the steady

state creep response of the FGM disk in terms of strain rates is significantly

superior as compared to a similar disk having uniform distribution of SiCp.

Singh and Ray (2002) studied the influence of anisotropy, induced due to

processing, on the steady state creep in a rotating disk made of 6061Al-20 wt%

SiCw and operating at 561 K. The creep behavior of the disk material was described

by Norton’s power law. It is observed that the presence of anisotropy in the disk

leads to significant reduction in both the tangential and radial strain rates over the

entire disk radius.

Jahed and Bidabadi (2003) presented a method to analyze primary and

secondary creep in axisymmetric rotating disks and pressure vessels subjected to

different types of loading, such as internal and external pressure, centrifugal

loading and temperature gradient. The method uses the basic solution for a rotating

disk made of uniform isotropic material and generates the solution for disk made of

non-uniform inhomogeneous material. Primary and secondary creep was predicted

and the results obtained were compared with those estimated by FEM technique. In

all the cases, a good agreement is observed.


62

Singh and Ray (2003b) analyzed steady state creep in a rotating disk made

of anisotropic 6061Al-20 wt% SiCw composite in the presence of thermal residual

stress. The study used Norton’s power law and the newly proposed yield criterion,

which at appropriate limits reduces to Hill anisotropic and Hoffman isotropic yield

criterion. It is observed that the presence of residual stress in an anisotropic disk

results in significant changes in the distribution of tangential stress but similar

distribution of radial stress when compared with those obtained in a similar

anisotropic disk but without residual stress. The presence of tensile residual stress

in the disk leads to significant increase in creep rate as compared to that observed

in a similar anisotropic disk without residual stress. In the presence of residual

stress, the radial strain rate becomes tensile in the middle of the disk, however, it

remains compressive towards the inner and outer radii. Similar results were also

noticed in the subsequent study by Singh and Ray (2004), based on the effect of

thermal residual stress in a rotating disk made of isotropic 6061Al-20 vol% SiCw

and undergoing steady state creep. The study used Hoffman yield criterion for

isotropic material.

Gupta et al (2004b, 2005) investigated steady state creep in a rotating disk

made of isotropic FGM containing SiCp in a matrix of pure aluminum and

operating in the presence of radial thermal gradient. The creep behavior of the

composite was described by Sherby’s law. The study indicates that the steady-state

strain rates in the FGM disk are significantly lower than that observed in an

isotropic disk having uniform distribution of SiCp, when both the disks operate

under thermal gradient. The study also reveals that the strain rates in composite
63

disk operating under thermal gradient are lower as compared to a similar disk

operating at a constant temperature, estimated by averaging the imposed thermal

gradient.

Gupta et al (2007) used Artificial Neural Network (ANN) for predicting the

creep response of a rotating Al-SiCp composite disk operating at elevated

temperature. The analysis of steady state creep, described by Sherby’s law, was

carried out for various combinations of SiCp size and content, and operating

temperature. The creep parameters were extracted from the limited experimental

uniaxial creep data available for Al-SiCp. The results obtained were used to train

the ANN model based on back propagation learning algorithm with SiCp size,

SiCp content and temperature as the input and stresses and strain rates as the output

parameters. The predictions obtained from the ANN model were compared with

the corresponding values obtained by analytical procedure. A nice agreement is

observed between the creep stresses and strain rates predicted by ANN model and

estimated analytically.

Singh (2008) used Norton’s creep law to analyze steady state creep in a

rotating disk made of 6061Al-20 vol% SiCw with varying extent of anisotropy,

characterized by a parameter α. It is observed that the radial strain rate, which

remains compressive for isotropic composite disk (α = 1.0) and anisotropic disk

having α > 1, becomes tensile in the middle of the disk when the extent of

anisotropy parameter α < 1. By changing the extent of anisotropy from α > 1 to α <

1, the variation of tangential strain rate in the disk remains similar, however, its

magnitude reduces by about five orders. Thus, the presence of anisotropy


64

introduces significant change in the strain rates, though its effect on the resulting

stress distribution is relatively small.

Gupta et al (2009a) developed a mathematical model to predict the steady

state creep response of a rotating disk made of SiC (particle/whisker) reinforced

6061Al matrix composite. The model was used to investigate the effect of SiC

morphology on the creep behavior of composite disk. The steady state creep

behavior was described by Sherby’s creep law. The study reveals that the creep

stresses and creep rates in the disk are significantly affected by the morphology of

SiC. The steady state creep rates in a whisker reinforced disk are observed to be

significantly lower than those observed in a particle reinforced disk.

Singh and Rattan (2010) analyzed steady state creep in a rotating disk made

of Al-SiCp composite by using isotropic Hoffman yield criterion. The results

obtained were compared with those using von Mises yield criterion. It is observed

that the distribution of stress in the disk is not too much affected in the presence of

phase specific thermal residual stress. The presence of residual stress leads to

increase the tangential strain rate, particularly in the region near the outer radius of

the disk, as compared to that observed in a similar disk but without residual stress.

The radial strain rate, which is compressive, changes significantly in the presence

of residual stress and even becomes tensile in the middle of the disk.

Gupta et al (2009b) formulated a mathematical model to describe steady

state creep behavior in rotating disks made of isotropic aluminum matrix composite

containing linear and quadratic distributions of SiCp in the radial direction. The

disks were assumed to operate under a radial thermal gradient, originating due to
65

braking action as estimated by FEM analysis. The steady state creep behavior of

the disks was described by Sherby's law. Based on the developed model, the

distributions of stresses and strain rates were obtained and compared for various

FGM disks containing the same average amount (20 vol%) of SiCp. The study

reveals that the creep stresses and steady state creep rates in a rotating FGM disk

can be significantly reduced by employing more SiCp reinforcement in the middle

of the disk as compared to the inner and the outer radii.

Loghman et al (2011) investigated time-dependent redistribution of creep

stress in a rotating disk made of Al–SiC composite by using Mendelson’s method

of successive elastic solution. The material creep behavior was described by

Sherby’s constitutive model by using available experimental results on Al–SiC

composite. It is concluded that the uniform distribution of SiC reinforcement does

not have considerable influence on stresses. However, the minimum and the most

uniform distribution of circumferential and effective thermoelastic stresses belongs

to FGM disk made of aluminum with 0% SiC at the inner surface and 40% SiC at

the outer surface. It has also been found that the stresses, displacement and creep

strains change with time at a decreasing rate in such a manner that after almost

50 years the solution approaches the steady state condition.

2.6 FORMULATION OF THE PROBLEM

It is revealed from the literature consulted that the problem of estimating

elastic stresses and deformations in rotating disk made of monolithic material has

been investigated by several workers, however, the studies pertaining to analysis of

creep deformation in rotating disk, especially of composite materials, are rather


66

scant. Gupta et al (2004a; b; 2005; 2007) analyzed creep response of rotating

composite disks made of MMCs based on aluminum or aluminum alloy matrix,

while describing its creep behavior by substructure-invariant model with a stress

exponent of 8. A detailed analysis of the published creep data for aluminum-based

composites suggests that the substructure-invariant model is untenable because it

leads to consistently higher value of activation energies for creep than anticipated

for lattice self-diffusion in aluminum. It is suggested that the creep behavior of

aluminum / aluminum alloy based composites could be analyzed in a better way by

assuming the flow to be controlled by creep of matrix material so that the true

stress exponent (n) is ~3 or ~5 (Cadek et al, 1995; Li and Langdon, 1999). In this

context, it is important that a study may be conducted to investigate the impact of

stress exponent (n) on the creep response of a rotating composite disk.

The stresses produced in the disk are due to rotary motion, which can be

reduced by increasing the thickness of the disk, as is evident from a number of

studies reported for variable thickness rotating disk made of monolithic

material. Based on these observations, a need is felt to carry out a detailed

analysis of the steady state creep in rotating disks made of composite material

and having different kinds of thickness profiles.

In view of the potential of FGMs to withstand severe thermo-mechanical

loadings, studies have been conducted to analyze creep in rotating disk made of

isotropic FG composites (Singh and Ray; 2001; 2003a Gupta et al, 2004b;

2005). In all these analyses, the disk was assumed to be of constant thickness

and having a radially varying distribution of SiCp in a matrix of pure aluminum.


67

Keeping this in view, it may be interesting to extend the above analysis for

rotating disk made of isotropic FG composite having variable thickness profile.

The processing of composites may lead to thermal residual stresses,

which are responsible for different yield strengths of the composite under

tension and compression. Singh and Ray (2003b) investigated steady state

creep, using Norton’s power law, in a constant thickness rotating composite

disk in the presence of thermal residual stresses. In this light, it is decided to

investigate the effect of thermal residual stress on the creep behavior of rotating

composite disk having variable thickness and creeping according to a threshold

stress based law with a stress exponent of 5.

The study conducted by Wahl et al (1954; 1963) reveals that while

evaluating the long-time creep behavior of steel disks, it is necessary to

consider the effects of transient period. The literature consulted so far,

pertaining to creep behavior of rotating composite disk, reveals that the effect of

transient period has not been considered while evaluating creep response of the

composite disk. Neglecting the transient phase can lead to errors in estimating

the creep deformation of the disk, whose magnitude may be of the order of

elastic strains in the disk. Therefore, it is imperative to undertake a study to

investigate the effects of initial transient period on the long-time creep behavior

of the rotating composite disk.


68

2.7 OBJECTIVES

On the basis of literature consulted and the facts reported in the previous

section, the following objectives are set for the present study.

i. To carry out an analysis of the steady state creep in a rotating

composite disk for different values of stress exponent (n) and to

investigate its impact on the creep behavior of the composite disk.

ii. To analyze the steady state creep in a rotating composite disk having

different kinds of thickness profiles.

iii. To investigate the effect of thermal residual stress on the steady

state creep behavior of the rotating composite disk of variable

thickness profile.

iv. To analyze the steady state creep behavior of an isotropic

functionally graded composite disk having variable thickness and

tailored distribution of reinforcement.

v. To analyze the effects of transient period in evaluating the long-time

creep behavior of the rotating composite disk.

2.8 METHODOLOGY

In order to meet the objectives outlined in the proposed research work, the

following methodology shall be adopted.

(i) The constitutive equations describing creep under multi-axial stress

will be developed on the basis of different yield criteria as applicable

to disk material.
69

(ii) The equilibrium equation of rotating disk will be solved along with

the developed constitutive equations to formulate a mathematical

model to estimate stresses and strains/strain rates in the composite

disk.

(iii) The material constants appearing in the creep law and yield criteria

will be extracted from the available experimental results or regression

analysis.

(iv) The developed mathematical model will be solved using suitable

numerical scheme and by developing a suitable software.

(v) The numerical computations will be carried out to calculate creep

response of the composite disk.

(vi) The results obtained will be analyzed to draw salient conclusions.


70

Fig. 2.1: Intersection of yield surface with π -plane


71

Fig. 2.2: Representation of von Mises yield criterion


72

Fig. 2.3: Simplified deformation mechanism map (Ashby and Frost, 1975)
73

Fig. 2.4: Typical creep curve showing three stages of creep


74

Fig. 2.5: Analysis of the competing processes determining the creep curve (Andrade, 1957)
75

Fig. 2.6: Creep strain rate as function of total strain (Deiter, 1988)
76

Fig. 2.7: Correlation between activation energies for high-temperature


creep and self-diffusion (Deiter, 1988)
77

Fig. 2.8: Variation of steady state creep rate with applied stress for PM 15 vol
% SiCp/8009Al composite (Ma and Tjong, 2000)
78

THot
Ceramic Phase

Ceramic matrix with


metallic inclusions

Transition region
`

Metallic matrix with


ceramic inclusions

Metallic Phase

TCold

(a) Continuously graded microstructure.

Metallic matrix
Ceramic inclusion

(b) Discretely graded microstructure

Ceramic phase
Inclusion phase 1

Inclusion phase 2

(c) Multiphase graded microstructure

Fig. 2.9: Characteristics of functionally graded materials

You might also like