Chapter 1 - Creep Introduction

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Chapter 1

Introduction

1.1 DESCRIPTION OF CREEP

Creep of materials is classically associated with time-dependent plasticity under a


fixed stress at an elevated temperature, often greater than roughly 0.5 Tm, where Tm
is the absolute melting temperature. The plasticity under these conditions is
described in Figure 1 for constant stress (a) and constant strain-rate (b) conditions.
Several aspects of the curve in Figure 1 require explanation. First, three regions are
delineated: Stage I, or primary creep, which denotes that portion where [in (a)] the
creep-rate (plastic strain-rate), e_ ¼ de=dt is changing with increasing plastic strain or
time. In Figure 1(a) the primary-creep-rate decreases with increasing strain, but with
some types of creep, such as solute drag with ‘‘3-power creep,’’ an ‘‘inverted’’
primary occurs where the strain-rate increases with strain. Analogously, in (b), under
constant strain-rate conditions, the metal hardens, resulting in increasing flow
stresses. Often, in pure metals, the strain-rate decreases or the stress increases to a
value that is constant over a range of strain. The phenomenon is termed Stage II,
secondary, or steady-state creep. Eventually, cavitation and/or cracking increase the
apparent strain-rate or decrease the flow stress. This regime is termed Stage III, or
tertiary creep, and leads to fracture. Sometimes, Stage I leads directly to Stage III
and an ‘‘inflection’’ is observed. Thus, care must sometimes be exercised in
concluding a mechanical steady-state (ss).
The term ‘‘creep’’ as applied to plasticity of materials likely arose from the
observation that at modest and constant stress, at or even below the macroscopic
yield stress of the metal (at a ‘‘conventional’’ strain-rate), plastic deformation occurs
over time as described in Figure 1(a). This is in contrast with the general observation,
such as at ambient temperature, where, a material deformed at, for example,
0.1–0.3 Tm, shows very little plasticity under constant stress at or below the
yield stress, again, at ‘‘conventional’’ or typical tensile testing strain-rates (e.g.,
104103 s1). {The latter observation is not always true as it has been observed
that some primary creep is observed (e.g., a few percent strain, or so) over relatively
short periods of time at stresses less than the yield stress in some ‘‘rate-sensitive’’ and
relatively low strain-hardening alloys such as titanium [1] and steels [2].}
We observe in Figure 2 that at the ‘‘typical’’ testing strain rate of about 104 s1,
the yield stress is sy1 . However, if we decrease the testing strain-rate to, for example,
107 s1, the yield stress decreases significantly, as will be shown is common for
metals and alloys at high temperatures. To a ‘‘first approximation,’’ we might

3
4 Fundamentals of Creep in Metals and Alloys

Figure 1. Constant true stress and constant strain-rate creep behavior in pure and Class M
(or Class I) metals.

consider the microstructure (created by dislocation microstructure evolution with


plasticity) at just 0.002 plastic strain to be independent of e_ . In this case, we might
describe the change in yield stress to be the sole result of the e_ change and predicted
by the ‘‘constant structure’’ stress-sensitivity exponent, N, defined by

N ¼ ½q ln e_ =q ln sT,s ð1Þ

where T and s refer to temperature and the substructural features, respectively.


Sometimes, the sensitivity of the creep rate to changes in stress is described by
a constant structure strain-rate sensitivity exponent, m ¼ 1/N. Generally, N is
relatively high at lower temperatures [3] which implies that significant changes
in the strain-rate do not dramatically affect the flow stress. In pure fcc metals,
Introduction 5

Figure 2. Creep behavior at two different constant strain-rates.

N is roughly between 50 and 250 [3]. At higher temperatures, the values may
approach 10, or so [3–10]. N is graphically described in Figure 3. The trends of N
versus temperature for nickel are illustrated in Figure 4.
Another feature of the hypothetical behaviors in Figure 2 is that (at the identical
temperature) not only is the yield stress at a strain rate of 107 s1 lower than at
104 s1, but also the peak stress or, perhaps, steady-state stress, which is maintained
over a substantial strain range, is less than the yield stress at a strain rate of 104 s1.
(Whether steady-state occurs at, for example, ambient temperature has not been
fully settled, as large strains are not easily achievable. Stage IV and/or
recrystallization may preclude this steady-state [11–13]). Thus, if a constant stress
sss2 is applied to the material then a substantial strain may be easily achieved at a
low strain-rate despite the stress being substantially below the ‘‘conventional’’ yield
stress at the higher rate of 104 s1. Thus, creep is, basically, a result of significant
strain-rate sensitivity together with low strain hardening. We observe in Figure 4 that
N decreases to relatively small values above about 0.5 Tm, while N is relatively high
below about this temperature. This implies that we would expect that ‘‘creep’’ would
6 Fundamentals of Creep in Metals and Alloys

Figure 3. A graphical description of the constant-structure strain-rate sensitivity experiment, N (¼ 1/m)


and the steady-state stress exponent, n.

Figure 4. The values of n and N as a function of temperature for nickel. Data from Ref. [7].
Introduction 7

be more pronounced at higher temperatures, and less obvious at lower temperatures,


since, as will be shown subsequently, work-hardening generally diminishes with
increasing temperature and N also decreases (more strain-rate sensitive). The
above description/explanation for creep is consistent with earlier descriptions [14].
Again, it should be emphasized that the maximum stress, sss2 , in a constant
strain rate (_e) test, is often referred to as a steady-state stress (when it is the result of
a balance of hardening and dynamic recovery processes, which will be discussed
later). The creep rate of 107 s1 that leads to the steady-state stress ðsss2 Þ is the same
creep-rate that would be achieved in a constant stress test at sss2 . Hence, at sss2 ,
107 s1 is the steady-state creep rate. The variation in the steady-state creep rate
with the applied stress is often described by the steady-state stress exponent, n,
defined by

n ¼ ½d ln e_ ss =d ln sss T ð2Þ

This exponent is described in Figure 3. Of course, with hardening, n is expected to


be less than N. This is illustrated in Figures 3 and 4. As just mentioned, generally,
the lower the strain-rate, or higher the temperature, the less pronounced the
strain hardening. This is illustrated in Figure 5, reproduced from [15], where the
stress versus strain behavior of high-purity aluminum is illustrated over a wide range
of temperatures and strain-rates. All these tests utilize a constant strain-rate. The
figure shows that with increasing temperature, the yield stress decreases, as expected.
Also, for a given temperature, increases in strain rate are associated with increases
in the yield stress of the annealed aluminum. That is, increases in temperature
and strain-rate tend to oppose each other with respect to flow stress. This can be
rationalized by considering plasticity to be a thermally activated process. Figure 5
also illustrates that hardening is more dramatic at lower temperatures (and higher
strain-rates). The general trend that the strain increases with increasing stress to
achieve steady-state (decreasing temperature and/or increasing strain-rate) is also
illustrated.

1.2 OBJECTIVES

There have been other, often short, reviews of creep, notably, Sherby and Burke [16],
Takeuchi and Argon [17], Argon [18], Orlova and Cadek [19], Cadek [20],
Mukherjee, [21], Blum [22], Nabarro and de Villiers [23], Weertman [24,25],
Nix and Ilschner [26], Nix and Gibeling [27], Evans and Wilshire [28], Kassner and
Pérez-Prado [29] and others [30–32]. These, however, often do not include some
important recent work, and have sometimes been relatively brief (and, as a result, are
8 Fundamentals of Creep in Metals and Alloys

Figure 5. The stress versus strain behavior of high-purity aluminum. Data from Ref. [15].

not always very comprehensive). Thus, it was believed important to provide a new
description of creep that is both extensive, current and balanced. Creep is discussed
in the context of traditional Five-Power-Law Creep, Nabarro-Herring, Coble,
diffusional creep, Harper-Dorn, low-temperature creep (power-law-breakdown or
PLB) as well as with 3-power Viscous Glide Creep. Each will be discussed separately.
Figure 6 shows a deformation map of silver [33]. Here, several deformation regimes
are illustrated as a function of temperature and grain size. Five-Power-Law Creep
is indicated by the ‘‘dislocation creep’’ regime bounded by diffusional creep (Coble
and Nabarro-Herring) and ‘‘dislocation glide’’ at low temperatures and high
stress. Deformation maps have been formulated for a variety of metals [33].
Additionally, Superplasticity particle-strengthening in creep and Creep Fracture will
be discussed.
Introduction 9

Figure 6. Ashby deformation map of silver from [33]. grain sizes 32 and 100 mm, e_ ¼ 108 s1,
A – dislocation glide, B – Five-Power-Law Creep, C – Coble creep, D – Nabarro-Herring creep,
E – elastic deformation.

You might also like