You are on page 1of 2

The Heat Treat Doctor

Daniel H. Herring | 630-834-3017 | heattreatdoctor@industrialheating.com

Segregation and Banding in Carbon


and Alloy Steel

o one likes surprises in manufacturing, and this


is especially true during heat treatment. To avoid
them, our attention is often focused on the type of
material being supplied to us and the process/equipment variables we must control. What we
dont often consider is the condition of the
incoming raw material. In this regard, the
result of alloy segregation and banding after heat treatment can cause considerable
angst between the heat treater and their
customer. Lets learn more.
Many people believe steels are classified
only by their chemistry. In point of fact,
the steelmaking process used; the casting
process employed (ingot or continuous casting); the size/shape of
semi-finished (blooms, billets or slabs) or finished products (plate,
sheet or bars); and the properties they are required to have for the
end-use application play a major role.

Segregation
Steels are not chemically homogeneous; they
form chemistry throughout the entire crosssectional area of their manufactured shapes.
When steel is cast, the first material to solidify is the outer area adjacent to the mold
walls (this surface zone is referred to as the
chill zone). This results in a thin layer of
equiaxed crystals with the same composition
as the liquid metal. These crystals continue
to grow inward as columnar grains in a dendritic shape parallel to the thermal gradient. Finally, solidification ends as the liquid
temperature drops and neighboring grains
impinge upon each other in the central zone
of the as-cast shape (Fig. 1).
The solidification process causes macroscopic as well as microscopic partitioning of
chemical elements as the liquid metal cools.
Macroscopically, segregation occurs at the

centerline of continuously cast products and at the centerline,


top and bottom of ingots. Microscopically, segregation occurs
between dendrites throughout the solidified section. Subsequent
mechanical hot working (e.g., rolling) creates longitudinal bands
of varying elemental chemistry.
Microsegregation is the difference in composition between the
center of the dendritic stem (i.e. the dendritic core) and the region
between the dendrite arms. The first liquid to solidify (in the center of the dendrite) will be alloy-rich if the alloy addition raises the
melting temperature, whereas the alloy (i.e. solute-rich) area will
be concentrated at the interdendritic regions if the alloy addition
lowers the melting temperature.
For the heat treater, alloy segregation can produce differences
in the hardenability of the steel. The difference in hardenability
between alloy-rich and alloy-lean regions can manifest itself by
creating harder and softer areas of martensite and mixed transformation products such as bainite.

Banding
Banding is caused by segregation of the alloying elements during
solidification. Subsequent hot-working operations result in segredo not have unigation aligned in the direction of working,
which results in the banded appearance
delineated in the microstructure. The distribution of microsegregation (in wrought
steels) depends on how much working has
been done to shape the part. The diffusion
rates of the alloying elements in steel control the homogenization of the casting. For
example, chromium and molybdenum homogenize readily, while nickel homogenizes
very slowly. The alternating bands of varying alloy chemistry result in different microstructures, orientated parallel to the rolling
direction of the material (Fig. 2).
Fig. 1. Schematic diagram of zones of crystal
morphologies in an as-solidified section of
Additional mechanical and/or thermal
steel. Illustrated are the outer chill zone, the
treatments to remove or reduce microsegrecolumnar zone and the interior equiaxed
gation add to the cost of manufacture and
zone. Macro and microsegregation are not
oftentimes are not considered economically
illustrated in this view.[1]

In this month's IH Monthly Prescription we talk about some reasons for having a lab. Every month, Dan Herring sits down with IHs
editor, Reed Miller, to talk technical. If you have a topic you would like them to discuss, drop us an
e-mail at reed@industrialheating.com. Find the podcast on our website. IH Monthly Prescription
is sponsored by Praxair.

16 October 2013 - IndustrialHeating.com

Fig. 2. Banded 4340 microstructure (50x, Vilellas Reagent)

feasible. For example, a high degree of homogenization can result from soaking the
segregated steel at an elevated temperature (often above 1200C or 2200F) for a
very long time (up to 100 hours or more in
some cases).
The primary cause of banding is due to
the segregation of substitutional alloying
elements (e.g., manganese, chromium,
molybdenum) during (dendritic) solidification. Cooling rate, austenitic grain
size and austenitizing temperature also
influence the severity of microstructural
banding.
Banding occurs in all steels. While
reheating of as-cast products and hot
rolling tend to reduce chemical
segregation, other factors (related to solidstate phase transformations and residual
solidification) result in a greater or lower
degree of banding in the microstructure
of all finished steel products.
Effect of Banding on
Heat Treatment
In general, hardness and microstructure
will be heavily influenced by segregation
and banding. Alloy-rich areas tend to
transform to martensite or bainite, while
alloy-lean areas show increasing amounts
of pearlite and ferrite (due to slower cool18 October 2013 - IndustrialHeating.com

Fig. 3. 4340 clutch plate. A region of reformed (untempered) martensite


(white) within the banded area of the heat-affected zone. This area exhibited
hardness values significantly higher than those of the surrounding area.
(1250x, Vilellas Reagent)

ing rates). For case-hardened parts, the


hardness of the case will be impacted,
particularly if higher concentrations of
retained austenite or bainite are formed
in the primarily martensite structure. In
certain applications, retained austenite
converts to untempered martensite in
service (Fig. 3), resulting in hardness and
property variation that can cause dimensional change or even component failure.
Tensile, yield and fatigue strength tend
to be relatively unaffected by the presence
or absence of microsegregation, although
ductility and toughness properties are negatively impacted.
As an example, spotty hardness and
poor mechanical properties can result
from segregated or banded microstructures having significant amounts of
coarse pearlite, coarse ferrite or ferrite
clusters. In induction hardening, higher
temperatures and longer heating times are
required to fully austenitize these structures and may result in grain growth, formation of coarse martensite on quenching, surface oxidation/decarburization
and increased chance of distortion. Thus,
having a prior microstructure consisting
of fine pearlite or a quench-and-tempered
structure is the most highly desired for
induction hardening.

Summary
Segregation and banding may or may not
be detrimental to the final product, but the
effects of these conditions should either be
evaluated before the raw material is heat
treated or negated (to the extent possible)
by appropriate thermal treatments (e.g.,
annealing, normalizing). IH
References
1. Krauss, George, Solidification, Segregation
and Banding in Carbon and Alloy Steels,
Metallurgical and Material Transactions B,
Volume 34B, December 2003, pp. 781 792.
2. Majka, Ted F., David K. Matlock and George
Krauss, Development of Microstructural
Banding in Low-Alloy Steel with Simulated
Mn Segregation, Metallurgical and Materials
Transactions A, Volume 33A, June 2002, pp.
1627 - 1637.
3. Shewman, Paul G., Transformations in Metals, McGraw-Hill, Book Company, 1969.
4. Parrish, Geoffrey, Carburizing: Microstructure and Properties, ASM International,
1999.
5. VanAken, David, Engineering Concepts:
Segregation and Banding in Steel, Industrial Heating, April 2001.
6. Rudenev, Valery I., Can the Fe-Fe3C Phase
Transformation Diagram be Directly Applied in Induction Hardening of Steel?,
Heat Treating Progress, June/July 2003.
7. Rudnev, V., D. Loveless, R. Cook and M.
Black, Handbook of Induction Heating, Marcel Dekker, Inc. 2002.

You might also like