You are on page 1of 2

1 of 2

D. Aliya and S. Lampman, Physical


Metallurgy Concepts in Interpretation of
Microstructures, Metallography and
Microstructures, Vol 9, ASM Handbook, ASM
International, 2004, p. 44–70

Physical Metallurgy Concepts in Interpretation of Microstructures


Debbie Aliya, Aliya Analytical, Inc.; Steve Lampman, ASM International

Solidification

The change from a solid to liquid or a liquid to solid state is a commonly recognized phase change. During solidification,
atoms arrange themselves into a crystalline structure. This means that in order to solidify, the atoms must be added to a
growing crystal one by one in the next available position. This takes time. In addition, the crystalline structure of metals in
the solid state has an important influence in the characteristics of solidification. In most common alloys, microstructure is
affected by the element that solidifies first from the melt, as described in more detail in the preceding section, “Equilibrium
Phase Diagrams,” for isomorphous and eutectic alloy systems.

In terms of initial nucleation during solidification of an alloy, tiny groups of atoms begin to solidify into a crystalline form.
These groups of atoms, or nuclei, may grow or redissolve, depending on the kinetic (or thermal) conditions of the
solidification process. For example, if the surrounding liquid is just below the liquidus of the alloy (or the melting point in a
pure metal), the so-called driving force for nucleation is low. In this condition, many nuclei may redissolve into the liquid
phase. However, some remaining nuclei also may easily grow because of the high diffusion rate associated with the high
temperature near the melting point (or near the top of the melting range of an alloy).

Thus, very slowly cooled metals will tend to have coarse microstructures, because the solid crystal phase forms at a
temperature near the top of the melting range, where nucleation of many particles is difficult, but growth of nucleated
particles is easy because diffusion is enhanced at higher temperatures. In contrast, quick cooling tends to produce relatively
fine grains, because the atoms quickly solidify with less diffusion-activated growth of the grains. Likewise, the size and
general coarseness of second-phase particles in a microstructure are determined by how far the temperature is above or
below the critical temperature of phase formation. Some nuclei also may form as a separate liquid within the matrix or main
liquid phase. Sulfur and some oxygen-containing compounds often separate out in a liquid condition from the surrounding
iron-base liquid in molten steel. When the steel cools, these separate high-sulfur areas become embedded as distinct solid
particles in the solid steel.

Because most real castings are not solidified in an equilibrium manner, deviations from the temperatures predicted by the
phase diagrams for the upper and lower ends of the freezing range may be expected. The general trend is that faster cooling
will depress the temperature range. Phase diagrams also can be used to illustrate how and why one of the common
microstructural features of coring is created in commercially produced castings.

Coring is a microlevel segregation of alloying elements that occurs during solidification of many commercial castings.
The origin of coring can be understood by referring back to the phase diagram of a simple isomorphous system ( Fig. 4).
The atoms already present in the solid state must mix with the new atoms just solidified. This mixing is very slow in the
solid state, because it depends on multiple-step diffusion of single atoms. If there is not enough time for the composition to
homogenize before new layers are added, there will be a difference in composition between the first portions of the material
to solidify and the last portions to do so. In the case of the solidification isomorphism of the hafnium-zirconium system,
very little segregation due to coring is possible, because the liquidus and solidus are close together ( Fig. 5b)

Figure 14(a) shows a heavily banded medium-carbon steel microstructure from an initially heavily microsegregated, or
cored, as-cast structure. The medium-carbon steel was produced by a continuous casting process, with higher manganese in
the pearlitic regions. This segregation of the manganese tends to persist through austenitizing in subsequent transformation
of the austenite, where martensite forms in the regions previously occupied by pearlite, while bainite or pearlite form where
the ferrite was. While carbon segregation is relatively easy to overcome, due to the small size of the carbon atoms, alloy
segregation is difficult or practically impossible to overcome. This shows that cored structures, once created, can be very
persistent, even after many thermal cycles and deformation. Figure 14(b) shows an aluminum casting with a noticeably
different composition (as revealed by the response to the etchant) in the (dark) centers of the dendrites and the lighter outer
areas.
2 of 2

Fig. 14 Examples of microstructure with alloy microsegregation from coring. (a) Heavily banded
medium-carbon steel microstructure that was produced by a continuous casting process with
alternating layers of ferrite (light) and pearlite (dark). The initial as-cast structure was heavily
cored (microsegregated), with higher manganese in the pearlitic regions indicated by
energy-dispersive spectroscopy microchemical analysis. 5% nital etch. (b) Coring during
solidification in the as-cast microstructure of an aluminum alloy revealed by differences in etching
due to compositional differences. The dark arrows show the central (core) portions of the primary
dendrites, while the light arrows show the last part of the material to solidify. Keller's reagent

Although coring may not be a concern at the eutectic concentration, because solidification occurs at one single temperature
at the eutectic, any deviation from the eutectic can create the possibility of coring. Sometimes, deviations from the eutectic
composition are not obvious by viewing a microstructure. Fast cooling near eutectic compositions may also result in
microstructures that appear to be eutectic, but only careful quantitative evaluation may reveal that the percentages of each
of the two solid phases are not equal to those predicted for the eutectic composition.

Liquation. Because the lowest freezing material in a cored microstructure is segregated to the edges of the solidifying
crystals (the grain boundaries), this material can remelt when the alloy sample is heated to temperatures below the
equilibrium solidus line. If grain-boundary melting (called liquation, or, less correctly “burning”) occurs while the sample
also is under stress, such as during hot forming, the liquefied grain boundaries will rupture, and the sample will lose its
ductility and be characterized as a hot short.
Copyright © 2005 ASM International®. All Rights Reserved.

You might also like