Sie sind auf Seite 1von 6

7 RECOVERY, RECRYSTALLIZATION, AND GRAIN GROWTH

RECOVERY, RECRYSTALLIZATION, AND GRAIN GROWTH are the main stages of


annealing when it is applied to a cold-worked metal. Such classification is approximate; some
overlapping between the stages usually occurs because of microstructural nonhomogeneity of
the specimen. To some extent, the annealing behavior of a metal may be different from metal
to metal, and for the same metal of different purity, but the basic phenomena involved in the
various annealing stages are similar.

When a metal is cold worked by plastic deformation, a small portion of the mechanical energy
expended in deforming the metal is stored in the specimen. This stored energy resides in the
crystals as point defects (vacancies and interstitials), dislocations, and stacking faults in
various forms and combinations, depending on the metal. Therefore, a cold-worked specimen,
being in a state of higher energy, is thermodynamically unstable. With thermal activation,
such as provided by annealing, the cold-worked specimen tends to transform to states of
lower energies through a sequence of processes with microstructural changes. Along with the
microstructural changes, the properties of the specimen will change correspondingly.
Deformation and annealing are important processing methods for producing desired
properties of the material by controlling its microstructures.

The Deformed State

For a better understanding of annealing, particularly recovery and recrystallization, a clear


understanding of the deformed state is useful. In addition, knowledge of the nature of the
highly deformed structure and the mechanisms that are responsible for its development
facilitates understanding the microstructural changes from a highly deformed to a fully
annealed state.

Plastic deformation is achieved principally by passage of dislocations through the lattice. In


the early stages of deformation, the dislocations are relatively long, straight, and few. With
increasing deformation, more dislocations from other slip systems are produced, causing
interactions among the various dislocations. These dislocations and clusters of short loops
tend to tangle and to align themselves roughly to broad boundaries. Small areas, or "cells,"
within which there are very few or no individual dislocations, are therefore outlined by these
broad boundaries. This sequence of cell-structure development is depicted in Fig.1 and 2.
These thin-foil specimens were prepared parallel to the rolling plane of the strip for
examination by transmission electron microscopy (TEM).
Fig. 1 Fe-3Si single crystal, cold rolled 5% in the (111)[11 ] orientation. Trails of
small dislocation loops, edge dislocation dipoles, and cusps on dislocation lines.
Thin-foil TEM specimen prepared parallel to the rolling plane. 62,000

Fig. 2 Fe-3Si single crystal, cold rolled 20% in the (111)[11 ] orientation showing
increased density of dislocations and clusters of short dislocation loops. Thin-foil TEM
specimen prepared parallel to the rolling plane. 62,000

Recovery

The earliest change in structure and properties that occurs upon annealing a cold-worked
metal is considered the beginning of recovery. As recovery proceeds, a sequence of structural
changes emerges: (1) the annealing out of point defects and their clusters, (2) the annihilation
and rearrangement of dislocations, (3) polygonization (subgrain formation and subgrain
growth), and (4) the formation of recrystallization nuclei energetically capable of further
growth. These structural changes do not involve high-angle boundary migration. Therefore,
during this stage of annealing, the texture of the deformed metal essentially does not change.
Changes in Properties. During the early stages of recovery in which the annealing out of
point defects and the annihilation and rearrangement of dislocations have occurred only to a
limited extent, the change in microstructure may not be apparent in conventional optical or
transmission electron micrographs. However, some physical or mechanical properties of the
metal, such as electrical resistivity, x-ray line broadening, or strain-hardening parameters,
may show the changes due to recovery with high sensitivities. Fig.3 and 4 show the changes
in resistivity and residual strain hardening, respectively, during isothermal recovery annealing.
These figures indicate that isothermal recovery of the various properties share the following
features: (1) there is no incubation period; (2) the rate of change is highest at the beginning,
decreasing with increasing time; and (3) at long times, the property approaches the
equilibrium value very gradually. However, hardness is less sensitive to early stages of
recovery in comparison with other properties, such as electrical resistivity, x-ray line
broadening, strain hardening, and density.

Fig. 3 Change in electrical resistivity during isothermal recovery for copper deformed by
torsion at 4.2 K

Fig. 4 Change in residual strain hardening during isothermal recovery for zone-melted iron
deformed 5% in tension at 0 C (32 F). The fraction of residual strain hardening, 1 - R =
( - 0) ( m - 0), where R is the fraction of recovery, 0 the flow stress of the
fully annealed material, m the flow stress of the strain-hardened material at a

predetermined constant strain, and a the initial flow stress after a recovery anneal.
Changes in microstructure during recovery become readily observable by transmission
electron microscopy when the density of dislocations is considerably reduced and the
appreciable rearrangement of the remaining dislocations has occurred. Fig.5(a), (b), (c), and
(d) show the sequence of dislocation substructure changes for a single crystal of Fe-3Si (wt%),
which was cold rolled 80% in the (001)[110] orientation and subsequently annealed at various
temperatures. From a structure of random arrays of dislocations (Fig.5(b)) to that of
well-defined subgrains (Fig.5(c)), the process is commonly referred to as polygonization.
Further annealing may gradually increase the average size of the subgrains (Fig.5(c) and (d)).

(a) (b)

(c) (d)

Fig. 5 Effect of annealing time and temperature on the microstructure of an Fe-3Si single crystal,
cold rolled 80% in the (001)[110] orientation. Thin-foil TEM specimens prepared parallel to the
rolling plane. All at 17,2000. (a) High density of dislocations and no well-defined cell
structure is revealed in the as-rolled condition. (b)Annealed at 400 C (750 F) for 1280
min. Reduced dislocation density and random arrays of dislocations are evident.
(c) Annealed at 600 C (1110 F) for 1280 min. Well-defined subgrains resulting from
polygonization are shown. (d) Annealed at 800 C (1470 F) for 5 min. Increased
average diameter of the subgrains is due to subgrain growth.
Recrystallization

Following recovery, recrystallization (or primary recrystallization) occurs by the nucleation


and growth of new grains, which are essentially strain-free, at the expense of the polygonized
matrix. During incubation, stable nuclei are formed by the coalescence of subgrains that leads
to the formation of high-angle boundaries. From that time on, subsequent growth of new
grains can proceed rapidly, because of the high mobility of the high-angle boundaries. The
rate of recrystallization later decreases toward completion as concurrent recovery of the
matrix occurs and more of the new grains impinge upon each other. Accordingly, isothermal
recrystallization curves are typically sigmoidal (see Fig.(6)). Because recrystallization is
accomplished by high-angle boundary migration, a large change in the texture occurs.

Sufficient deformation and a sufficiently high temperature of annealing are required to initiate
recrystallization following recovery. With a low degree of deformation and a low annealing
temperature, the specimen may recover only without the occurrence of recrystallization. In
situ recrystallization, or complete softening without the nucleation and growth of new grains
at the expense of the polygonized matrix, is a process of recovery, not recrystallization,
because it does not involve high-angle boundary migration. Consequently, there is no
essential change in texture following in situ recrystallization.

Fig. 6 Effect of penultimate grain size on the recrystallization kinetics of a low-carbon


steel, cold rolled 60% and annealed at 540 C (1005 F). Note the incubation time is
shortened as the penultimate grain size before cold rolling is decreased.

Grain Growth

After recrystallization is complete--that is, when the polygonized matrix is replaced by the
new strain-free grains--further annealing increases the average size of the grains. The process,
known as grain growth, is accomplished by the migration of grain boundaries. In contrast to
recrystallization, the boundary moves toward its center of curvature. Some of the grains grow,
but others shrink and vanish. Because the volume of the specimen is a constant, the number of
the grains decreases as a consequence of grain growth. The driving force for grain growth is
the grain-boundary free-energy, which is substantially smaller in magnitude than the driving
energy for recrystallization.

According to the growth behavior of the grains, grain growth can be further classified into
two types: normal or continuous grain growth and abnormal or discontinuous grain growth.
The latter has also been termed exaggerated grain growth, coarsening, or secondary
recrystallization.

Normal or continuous grain growth occurs in pure metals and single-phase alloys.
During isothermal growth, the increase in the average grain diameter obeys the empirical

growth law, which can be expressed as = Ktn where is the average grain diameter, t is
the annealing time, and K and n are parameters that depend on material and temperature.

Therefore, when and t are plotted on a logarithmic scale, a straight line should be obtained,
with K as the intercept and n the slope. The value of n, the time exponent in isothermal grain
growth, is usually less than, or at most equal to, 0.5. A typical example for isothermal grain
growth in zone-refined iron is shown in Fig.7. The deviation from a straight-line relationship
for very short annealing times at low temperatures is due to recrystallization, and that for long
annealing times at high temperatures is due to the limiting effect of the sheet specimen
thickness.

Fig. 7 Normal grain growth in zone-refined iron during isothermal anneals. Closed
circles represent specimens for which statistical analysis of grain-size and grain-shape
distributions was conducted.

Das könnte Ihnen auch gefallen