Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

1

The Need for Ductility in Steel Reinforcement


Fahmida Gulshan
Assistant Professor
Materials and Metallurgical Engineering Department
Bangladesh University of Engineering and Technology

Introduction
Ductility is a measure of the ability of a material to sustain plastic deformations before collapse.
Ductile structures suffer relatively large deformations before failure and this provides warning of
impending failure. Ductility also provides robustness and helps in dissipating the internal energy
generated by impact loading, such as earthquake actions, dynamic impact and blast loadings. A
material that experiences very little or no plastic deformation upon fracture is termed brittle.

Concrete is a brittle material. The ductility within a concrete structure is provided by the steel
reinforcement. Therefore the steel used for reinforcing concrete must be sufficiently ductile so
that every reinforced concrete section has the capacity to deform by an adequate amount.
Initially it was assumed that the concrete was the limiting factor for plastic rotation and it was,
therefore, considered that the capacity for plastic rotation was independent of the type of steel
used. The steel was assumed to have sufficient ductility so that it did not limit rotation. This
presumption was made because at that time steel had comparatively low strength and high
ductility owing to its chemical composition and the manufacturing processes employed. Recent
developments in design codes have recognized the importance of ductility of the reinforcing steel
and provide designers with the means of using the additional ductility provided by high ductility
steels. This paper aims to highlight the importance of ductility in reinforced concrete structures
and to elaborate the factors that affect ductility in steel reinforcements.

Structure of steel
Steel is an alloy of iron and carbon (containing up to 1.7% carbon). As in all other metals, the
atoms in iron occupy definite positions, and this gives rise to crystalline structure of iron. The
atoms of metals can be considered to resemble spheres and thus any type of arrangement of
atoms gives rise to the formation of interstices (voids) within the structure. In steel the carbon
atoms occupy the interstitial positions in the crystal structure formed by the iron atoms. The
atoms of iron arrange themselves either in the form of a body centred cubic (BCC) or a face
centred cubic (FCC) crystal (Figure 1).

Dr. Fahmida Gulshan, e-mail: fahmidagulshan@mme.buet.ac.bd


2

Figure 1: B.C.C. and F.C.C. structures

In iron, the arrangement of atoms is also dependent on temperature. A metal in which the
arrangement of atoms depends on temperature is known as an allotropic metal. At room
temperature the crystal structure is body centered cubic and this form is known as alpha iron.
When heated to 910C the structure becomes face centred cubic, known as gamma iron. At 1400C
the face centered cubic structure reverts again to body centered cubic (known as delta iron) form.
The essential difference between alpha iron and delta iron is only in the temperature range over
which each exists (Figure 2).

Figure 2: Cooling curve for pure iron

The density of packing [defined as that fraction of a crystal actually occupied by atoms] depends
on the nature of packing of the atoms. The sizes of the interstitial spaces between the atoms of
iron depend on the arrangement of the atoms. The solubility of carbon in body centred iron is
3

only 0.025% at 723oC while the maximum solubility of carbon in face centred cubic gamma iron
is 1.7% at 1130oC. This allotropic change in iron and the resultant change of solubility of carbon
in iron make it possible to change the properties of iron by controlling the rate of cooling from
temperatures in which the structure is face centred cubic (austenite). We shall come back to this
point again.

How a material deforms:


The deformation in a metal on the application of an external force occurs by a process known as
slip. At the atomic level slip occurs as plane of atoms, called slip planes, slide relative to one
another. In metal crystals slip occurs preferentially on planes of high atomic density. In general
the planes of greatest atomic density are the most widely spaced planes and therefore are
sheared most easily. Not only does slip tend to take place on preferred crystallographic planes,
but also the slip direction of a crystal has been found to be almost exclusively a close packed
direction.

Figure 3: Slip planes and directions

The experimentally determined slip direction coincides with close packed direction. This can be
clearly explained with the aid of Figure 3(a). This figure shows a hard ball model of simple cubic
structure. Line mn is a close packed direction and qr is an arbitrarily chosen non-close packed
direction. In figure 3(b),the upper half of lattice has been sheared to the right by an amount “a”,
the distance between atoms in the mn direction. The symmetry in the crystal has been retained.
A shear of “c” in figure 3(c) (the distance between atom centres in this direction) also preserves
the lattice. However a simple calculation shows that c is larger than a (c = 1.414a).
4

Depending on the magnitude of the external forces, materials may experience two kinds of
deformation: elastic and plastic. Plastic deformation is permanent, and strength and hardness are
measures of a material’s resistance to this deformation. It may be interesting to note that actual
deformation of a metal requires much less energy than are determined by theoretical calculations.
An attempt to explain this ease with which metals deform plastically has led to the concept of
dislocations.

The dislocations may be regarded as defects in the crystal structures and are introduced during
solidification, during plastic deformation and as a consequence of thermal stresses that result
from rapid cooling. Virtually all crystalline materials contain dislocations. The number of
dislocations, or dislocation density in a material, is expressed as the total dislocation length per
unit volume. Dislocation densities as low as 103 mm-2 are typically found in carefully solidified
metal crystals. For heavily deformed metals the density may run as high as 10 9 to 1010 mm-2.
Heat treating a deformed metal specimen can diminish the density to on the order of 105 to 106
mm-2. Thus not only dislocations are produced during solidification and during plastic deformation,
but they may be partially eliminated during an annealing heat treatment. Dislocations cannot
easily cross over into crystals of different orientation (at grain boundaries) or crystals of different
modules (second phase particles). Both these therefore act as obstacles for the dislocation
movements on a slip plane.

Dislocation motion is analogous to the mode of locomotion employed by a caterpillar. The


caterpillar forms a hump near its posterior end by pulling in its last pair of legs a unit leg distance.
The hump is propelled forward by repeated lifting and shifting of leg pairs. When the hump
reaches the anterior end, the entire caterpillar has moved forward by the leg separation distance.
The caterpillar hump and its motion correspond to the extra half plane of atoms in the dislocation
model of plastic deformation.

On a microscopic scale, plastic deformation corresponds to the net movement of large numbers
of atoms in response to an applied stress. In crystalline solids, plastic deformation most often
involves the motion of large number of dislocations. The process by which plastic deformation is
produced by dislocation motion is termed slip; the crystallographic plane along which the
dislocation line traverses is the slip plane as indicated in Figure 4.

Dislocations do not move with the same degree of ease on all crystallographic planes of atoms
and in all crystallographic directions. Ordinarily there is a preferred plane and in that plane there
are specific directions along which dislocation motion occurs. This plane is called the slip plane; it
follows that the direction of movement is called the slip direction. This combination of the slip
plane and the slip direction is termed the slip system. The slip system depends on the crystal
5

structure of the metal and is such that the atomic distortion that accompanies the motion of a
dislocation is a minimum.

Figure 4: Atomic rearrangements that accompany the motion of an edge dislocation as it moves in response to an applied
shear stress. (a) The extra half plane of atoms is labeled A. (b) The dislocation moves one atomic distance to the right as
A links up to the lower portion of plane Bl in the process the upper portion of B becomes the extra half plane (c) A step
forms on the surface of the crystal as the extra half plane exists.

For a particular crystal structure, the slip plane is that plane having the most dense atomic
packing, that is has the greatest planar density. The slip direction corresponds to the direction, in
this plane, that is most closely packed with atoms, that is, has the highest linear density. Metals
with FCC or BCC crystal structures have a relatively large number of slip systems (at least 12).
These metals are quite ductile because extensive plastic deformation is normally possible along
the various systems.

Strain hardening

In a crystal the atoms have equilibrium positions. If removed from the equilibrium positions the
atoms tend to restore the equilibrium positions. If the deformations are carried out at high
temperatures, the atoms have sufficient energy to restore the equilibrium positions. If the
deformation is carried out at low temperatures, the atoms cannot go back to the equilibrium
positions. This raises the energy of the system and the material becomes harder. This
phenomenon is known as strain hardening or work hardening and it limits the extent of the
shape changes possible at low temperatures. The classic and comic picture of a carnival strong
man bending a bar of steel and then asking someone in the crowd to straighten it is an example
of the effect of work hardening. If the bar were originally soft enough to allow a person to bend
it, work hardening would occur during the bending operation. It would then be impossible for
anyone of comparable strength to restore the bar to its original shape because of the increased
6

yield strength caused by strain hardening. Thus working or straining a material causes an
increase in its strength with increasing deformation. Because of this increase in strength, it
becomes impractical if not impossible to deform materials beyond a certain point (Figure 5).

Figure 5: The influence of cold work (CW) or strain hardening on the stress strain behaviour for a low carbon steel

The mechanism of formation of dislocations is not clearly understood and it is assumed that
defects capable of generating dislocations are more or less uniformly distributed within each
crystal and according to their geometry and chance location various defects produce both
positive and negative dislocations in response to same shearing load. As slip occurs in any given
crystal, both types of dislocations are present and are moving simultaneously in opposite
directions across the crystal. In such a case, a positive dislocation moving along one plane may
encounter the stress field of a negative dislocation moving along some nearby plane. Additional
force will then be required to overcome the attraction of each dislocation for the other. This
increment of load necessary to free the struck dislocations represents strain hardening. As the
dislocation density increases, this resistance to dislocation motion by other dislocations becomes
more pronounced. Thus the imposed stress necessary to deform a metal increases with
increasing cold work.

Since slip is itself localized on specific planes, this strain hardening is at first a local effect
confined to the immediate vicinities of the planes which happened to slip at relatively low stress.
But as slip recurs on other planes throughout the crystal, the dislocation lattice is created within a
continuously more extensive volume of metal and general strain hardening of the crystal
structure.
7

Strain hardening is often utilized commercially to enhance the mechanical properties of metals
during fabrication procedures. The effects of strain hardening may be removed by heat treatment
because an increase in temperature gives the atoms enough energy to restore equilibrium sites
and thus reduce the overall energy.

Heat Treatment of steel

We have seen that iron is an allotropic metal. The high temperature modification (γ-iron) can
contain in solution a maximum of 1.7 % carbon, while the room temperature modification (α-
iron) can contain in solution only up to 0.025% carbon). Thus carbon atoms have to move out to
allow transformation to the low temperature modification.

Figure 6: The lattice structure of Martensite (Atom X is displaced because of a carbon atom vertically above it in the
lattice).

If a γ solid solution (high temperature face centred cubic crystal of iron) is quenched in water to
prevent diffusion of carbon atoms, the carbon atoms get trapped in the structure (Figure 6)
setting up intense local lattice strains that block movement of dislocations. As a result, the
structure becomes hard and extremely strong, but also very brittle. Under an optical microscope
it looks like an array of needles (Figure 7). The needle-like structure formed when carbon is
trapped by the iron crystal lattice is called martensite. It represents the maximum hardness
obtainable with any given carbon content.
8

The degree of hardness of a quenched austenitic structure is proportional to the lattice strain.
The lower the carbon content, the lower the strain. The maximum hardness appears to be
reached in the region of 0.6 to 0.8% carbon. Industrially, steels in which the carbon is too low to
give a useful hardening effect on quenching (below 0.25%) are known as mild steels, and those
that do show a useful hardening effect are termed medium carbon or structural steels. Those
containing more than 0.8% carbon offer greater strength and hardness on heat-treatment and
these so-called high-carbon steels are often used for cutting and forming tools.

Figure 7: Microstructure of martensite formed by quenching steel from the austenite region: very hard and brittle

Unfortunately, the high hardness and strength of a wholly martensitic structure is difficult to
exploit in practice because it is so brittle. Steel in this condition requires considerable support
from a tougher material to be of use in service. An added disadvantage of quenched steels is
dimensional change during transformation, due to the different lattice spacings of iron atoms in
α-, and y-iron. Martensite expands by a few percent on formation, the extent of expansion
increasing with the carbon content. This expansion causes distortion and internal stresses that
can lead to cracking. Water-, or oil-quenched steel with more than about 0.4% carbon will almost
certainly develop quench cracks, and the main reason why the welding of such steels is difficult is
that cracking can occur even on cooling in air.

One way of overcoming the brittleness of quenched martensitic steel, to give desired
combinations of hardness and toughness throughout the section, is to produce martensite and
then temper it. Tempering is a controlled heat-treatment allowing some of the trapped carbon to
escape from the interstitial spaces between the iron atoms and thus reduce the stress.

The temperature at which tempering is carried out is critical. Between 200°C and 300°C diffusion
rates are slow and only a small amount of carbon is released. As a result the structure retains
9

much of its hardness but loses some of its brittleness. Between 500°C and 600°C diffusion is
much faster, allows most of the carbon atoms to diffuse out from between the iron atoms to
form cementite and greatly enhances ductility. The dramatic change in mechanical properties
brought about by tempering a martensitic 0.4% carbon steel is shown in Figure 8.

Figure 8: Property changes with tempering of a martensitic (quenched) carbon steel. Notice how the most drastic
changes take place between 300 and 400°C

Measure of Ductility

There are two basic parts to a typical stress-strain relationship (Figure 9): elastic region and
plastic region. In the elastic region the strain (deformation) is linearly proportional to the applied
load until the yield strength is reached. The elastic strain is recovered if the load is removed. The
yield stress is a measure of resistance to plastic deformation of material. It is the maximum
stress that a material can withstand before starting permanent deformation. In most ferrous
materials, the elastic-plastic transition is well defined and occurs abruptly. Some steels show two
distinct yield points (Figure 10). At the upper yield point, plastic deformation is initiated, with an
actual decrease in strength. Continued deformation fluctuates slightly at some constant stress
value, termed the lower yield point. The yield strength is then taken as the average stress at the
lower yield point.

Once the yield strength is exceeded, deformation is no longer proportional to load. Small
increments of load produce larger plastic deformation, which cannot be reversed on unloading.
Deformation continues with increasing load until the maximum (ultimate) load is reached (the
maximum of the stress-strain curve). Beyond this point, the load decreases as the cross section
reduces in a local area before fracture.
10

Figure 9: Stress-strain diagram

Where, σp-proportional limit (maximum stress in linear region)


σy –yield strength (stress that result in a specific amount of permanent strain)
σuts-ultimate tensile strength
εf-elongation or strain to failure (total strain at failure)
E-modulus of elasticity (slope of curve in linear region)

Figure 10: (a) Typical stress-strain behavior for a metal showing elastic and plastic deformations, the proportional limit P,
and the yield strength σy (b) Stress strain behavior found for some steels demonstrating the yield point phenomenon.
11

After yielding, plastic deformation becomes more and more difficult. The stress necessary to
continue deformation rises with increasing strain, and the stress-strain curve tends to flatten out.
This is called strain hardening (as explained earlier). With deformation, the number of
dislocations inside the material is increased which hinder further movement of dislocation, and
the material becomes stronger. The stress necessary to continue plastic deformation in metals
increases to a maximum, and then decreases to the eventual fracture. The tensile strength (TS)
corresponds to the maximum stress that can be sustained by a structure in tension; if this stress
is applied and maintained, fracture will result. All deformation up to this point is uniform
throughout the narrow region of the tensile specimen. At this maximum stress a small
constriction or neck begins to form at some point, and all subsequent deformation is confined at
this neck. This phenomenon is termed necking and fracture ultimately occurs at the neck.

Figure 11: Schematic tensile stress-strain diagram showing the phenomenon of elastic strain recovery and strain
hardening. The initial yield strength is designated as σy1; σy2 is the yield strength after releasing the load at point D and
then upon reloading.

If a material is deformed plastically and the stress is then released, the material ends up with a
net permanent strain (Figure 11). If the stress is reapplied the material again responds elastically
at the beginning up to a new yield point that is higher than the original yield point. This is due to
strain or work hardening of material. The amount of elastic strain that it will take before reaching
the yield point is called elastic strain recovery.
12

Calculation of Ductility:

Ductility may be expressed quantitatively as either percent elongation or percent area reduction.
The percent elongation %EL is the percentage of plastic strain at fracture

lf  lo
% EL = x 100
lo

Where lf is the fracture length and lo is the original gauge length.


Percent area reduction %RA is defined as

 Ao  Af 
%RA=   x 100
 Ao 

Where Ao is the original cross sectional area and Af is the cross sectional area at the point of
fracture.
The tensile stress-strain behaviours for both ductile and brittle materials are schematically shown
in figure 12.

Figure 12: Schematic representations of tensile stress-strain behaviour for brittle and dultile materials loaded to fracture

Ductile materials typically exhibit substantial plastic deformation with high energy absorption
before fracture. The fracture process normally occurs in several stages (Figure 13). First after
necking begins, small cavities or microvoids form in the interior of the cross section. Next, as
deformation continues, these microvoids enlarge, come together, and coalesce to form an
elliptical crack, which has its long axis perpendicular to the stress direction. The crack continues
13

to grow in a direction parallel to its major axis by this microvoid coalescence process. Finally,
fracture ensues by the rapid propagation of a crack around the outer perimeter of the neck, by
shear deformation at an angle of about 45 degree with the tensile axis. Fracture having this
characteristic surface contour is termed a cup and cone fracture because one of the mating
surfaces is in the form of a cup, the other like a cone. In this type of fractured specimen the
central interior region of the surface has an irregular and fibrous appearance, which is indicative
of plastic deformation. Brittle fracture takes place without any appreciable deformation and by
rapid crack propagation. The direction of the applied tensile stress and yields a relatively flat
fracture surface as indicated in figure 14.

Figure 13: Stages in the cup and cone fracture (a) initial necking (b)Small cavity formation (c)Coalescence of cavities to
form a crack (d) Crack propagation (e) Final shear fracture at 45 degree angle relative to the tensile direction

Figure 14: (a) Cup and cone fracture (b) Brittle fracture
14

Steel reinforcement: The quest for strength and ductility


Strength is a characteristics of steel required in calculations and recognized in standards and
design codes, because steel plays an integral role in the mechanical behaviour of reinforced
concrete. Standards define strength requirements by means of two parameters: the elastic limit
or yield strength (which is used in the steel grade designation) and the ultimate tensile strength.
Strength is a necessary requirement but it is not sufficient in itself to define the behaviour of
reinforcing steel in concrete. There is also a requirement for ductility.

Until the 1960s mostly plain mild steel rebars with yield strength of about 250 N/mm 2 were used.
Around 1960 ribbed mild steel bars were introduced to allow for a better bond with concrete.
Both the plain and the ribbed bars had very high ductility as indicated by the elongation values.

In the late 1960s, the cold twisted deformed (CTD) rebars with yield strength of around 400
N/mm2 and elongation values of 14-15% were first produced. The cold twisted deformed (CTD)
bars are produced by cold working process, which is basically a mechanical process. It involves
stretching and twisting of mild steel, beyond the yield plateau, and subsequently releasing the
load.

The cold twisted bars inherent problems of inferior ductility, weldability and increased rate of
corrosion due to presence of residual stresses and higher carbon content. Additionally, cold
twisting being labour intensive, enhances cost of production with limitation on production rates.
Increasingly varied and innovative applications have resulted in growing demand for larger
diameter bars with similar strength, elongation, weldability and bendability as the small size bars.
Along with this, there is also a need for these steel bars to be welded and fabricated on the site
easily. For this strength and ductility have to be achieved at the lowest possible carbon content.
The most challenging requirement is to achieve all these superior properties at relatively lower
cost.

Since high strength was achieved at the cost of ductility, higher strength CTD bars did not gain
global acceptance as elongation values dropped to 12 % or less. Europe, where the CTD process
was developed, gave up its use in the 1970s, a few years after its development The demand of
civil engineers for rebars of yield strength 500 N/mm 2 with good ductility & weldability remained
unfilled.

In recent years, there has been another major effort for the production of high strength
deformed bars by using small amounts of alloying elements. Microalloying with Nb, V, Ti & B, in
15

combination or individually has made it possible to produce high strength reinforcement bars
having yield strength of 500 - 550 MPa. Production of reinforcement bars by the addition of
micro-alloys gives the desired results of high strengths but at a cost, which is prohibitive.
Moreover, the achieved ductility is also low.

The objective of guaranteed minimum 500 N/mm 2 yield strength with adequate ductility for
seismic zones was finally met through the development of the “Quenching & Tempering”
technology in early 1980s. Two such processes were developed in Europe, Thermex and
Tempcore, and both received world patents – and global acceptance amongst the civil engineers
because it met all their requirements. The steel mills all over have increasingly resorted to these
unique technologies and demand for such rebars continues to increase.

In Q and T, the steel bars are passed through a specially designed water-cooling system where
they are kept till the outer surface of the bars becomes colder while the core remains hot. This
creates a temperature gradient in the bars. After the intensive cooling, the bar is exposed to air
and the core re-heats the quenched surface layer by conduction, therefore tempering the
external martensite. When the bars are taken out of the cooling system, the heat flows from the
core to the outer surface, further tempering of the bars, which helps them attain a higher yield
strength. The resulting heat-treated structure imparts superior strength and toughness to the
bars.

The pre-determined cooling of the bar periphery transforms the peripheral structure to
martensite and then annealed through the heat available at the core. The peripheral and core
temperature difference finally equalises at around 600 0C and the resultant bar structure is of
tempered martensite at the periphery and of fine-grained ferrite-pearlite at the core(figure 15).

Generally speaking, the resultant soft core forms about 65-75 per cent of the area (depending
upon the desired minimum yield strength) and the rest is the hardened periphery. The equalizing
temperature together with the final rolling temperature is the most important parameter to
achieve the required mechanical properties.

Finally, when the bar is discharged on to the cooling beds, the remaining austenite transforms
into a very fine-grained pearlite structure.
16

(a) (b) (c)

Figure 15: microstructure of quenched and tempered bar (a)Outer case: Tempered Martensite (b)Inner case: Bainite like
intermediate layer and (c)Core: Ferrite/Pearlite

Q and T bars having uniform and concentrated hardened periphery and the softer core will have
the desired tensile strengths coupled with high elongation as required in seismic zones.
Depending on the size and grade, rebars with hardened periphery of about 15 to 30 per cent of
the cross sectional area of the bar are ideal for civil constructions (constructions of houses,
offices, etc.)

Concluding Remarks

The unsatiable demand for still higher strength rebars with adequate ductility and weldability has,
in the past, led to interesting developments in the technology of rebar production. The increasing
awareness about the quality of the bars is sure to lead to further developments. In Bangladesh it
is time for us to rise and face this reality and prepare ourselves to make our contributions to
future developments in the field of production of rebars with still higher strength and ductility.

You might also like