Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 26

NAME; OSAKWE OZICHUKWU

MATRIC; 090404049

COURSE; MME 309

DEPARTMENT; MECHANICAL

ENGINEERING

MANUAL NO; 28

TITLE;

ROCKWELL HARDNESS TEST

COLD WORKING

METALLOGRAPHY

PHASE DIAGRAMS
 ROCKWELL HARDNESS TEST

OBJECTIVES

 To understand what hardness is, and how it can be used to indicate some properties of

materials.

 To conduct typical engineering hardness tests and be able to recognise commonly

used hardness scales and numbers.

 To be able to understand the correlation between hardness numbers and the

properties of materials.

 To learn the advantages and limitations of the common hardness test methods.

EQUIPMENT; Rockwell Standard Hardness Testing machine

INTRODUCTION

Many materials, when in service, are subjected to forces or loads; examples include the

aluminium alloy from which an airplane wing is constructed and the steel in an automobile

axle. In such situations it is necessary to know the characteristics of the material and to

design the member from which it is made such that any resulting deformation will not be

excessive and fracture will not occur. The mechanical behaviour of a material reflects the

relationship between its response or deformation to an applied load or force. Important

mechanical properties are strength, hardness, ductility, and stiffness.

It is a common practice to test most materials before they are accepted for processing, and

before they are put into service to determine whether or not they meet the specifications

required. One of these tests is for hardness. The Rockwell machine is one commonly used for

this purpose.
THEORY;

HARDNESS

Another mechanical property that may be important to consider is hardness, which is a

measure of a material’s resistance to localized plastic deformation (e.g., a small dent or a

scratch). Early hardness tests were based on natural minerals with a scale constructed solely

on the ability of one material to scratch another that was softer. A qualitative and somewhat

arbitrary hardness indexing scheme was devised,termed the Mohs scale, which ranged from 1

on the soft end for talc to 10 for diamond. Quantitative hardness techniques have been

developed over the years in which a small indenter is forced into the surface of a material to

be tested, under controlled conditions of load and rate of application. The depth or size of the

resulting indentation is measured, which in turn is related to a hardness number; the softer the

material, the larger and deeper the indentation, and the lower the hardness index number.

Measured hardnesses are only relative (rather than absolute), and care should be exercised

when comparing values determined by different techniques.

Hardness tests are performed more frequently than any other mechanical test for several

reasons:

1. They are simple and inexpensive; ordinarily no special specimen need be prepared, and the

testing apparatus is relatively inexpensive.

2. The test is non-destructive; the specimen is neither fractured nor excessively deformed; a

small indentation is the only deformation.

3. Other mechanical properties often may be estimated from hardness data, such as tensile

strength.
ROCKWELL HARDNESS TESTS

The Rockwell tests constitute the most common method used to measure hardness because

they are so simple to perform and require no special skills. Several different scales may be

utilized from possible combinations of various indenters and different loads, which permit the

testing of virtually all metal alloys (as well as some polymers). Indenters include spherical

and hardened steel balls having diameters of 1/16 , 1/8 , ¼ and 1/3 in. (1.588, 3.175, 6.350,

and 12.70 mm), and a conical diamond (Brale) indenter, which is used for the hardest

materials. With this system, a hardness number is determined by the difference in depth of

penetration resulting from the application of an initial minor load followed by a larger major

load; utilization of a minor load enhances test accuracy. On the basis of the magnitude of

both major and minor loads, there are two types of tests: Rockwell and superficial Rockwell.

For Rockwell, the minor load is 10 kg, whereas major loads are 60, 100, and 150kg. Each

scale is represented by a letter of the alphabet; several are listed with the corresponding

indenter and load in Tables 7.4 and 7.5a. For superficial tests, 3 kg is the minor load; 15, 30,

and 45 kg are the possible major load values. These scales are identified by a 15, 30, or 45

(according to load), followed by N, T, W, X, or Y, depending on indenter. Superficial tests

are frequently performed on thin specimens. When specifying Rockwell and superficial

hardnesses, both hardness number and scale symbol must be indicated. The scale is

designated by the symbol HR followed by the appropriate scale identification.

For example, 80 HRB represents a Rockwell hardness of 80 on the B scale, and 60 HR30W

indicates a superficial hardness of 60 on the 30W scale. For each scale, hardnesses may range

up to 130; however, as hardness values rise above 100 or drop below 20 on any scale, they

become inaccurate; and because the scales have some overlap, in such a situation it is best to

utilize the next harder or softer scale. Inaccuracies also result if the test specimen is too thin,
if an indentation is made too near a specimen edge, or if two indentations are made too close

to one another. Specimen thickness should be at least ten times the indentation depth,

whereas allowance should be made for at least three indentation diameters between the center

of one indentation and the specimen edge, or to the centre of a second indentation.

Furthermore, testing of specimens stacked one on top of another is not recommended. Also,

accuracy is dependent on the indentation being made into a smooth flat surface.

The modern apparatus for making Rockwell hardness measurements is automated and very

simple to use; hardness is read directly, and each measurement requires only a few seconds.

The modern testing apparatus also permits a variation in the time of load application. This

variable must also be considered in interpreting hardness data.

1
hardness α
toughness

HRC – hardness Rockwell scale C (diamond indenter)

MINOR LOAD – load that holds the indenter in place.

MAJOR LOAD – load that pulls down the indenter in order for indentation of the material

(about 150kg).

35·85 HRC – hardness of the standard calibration block.


PROCEDURE

1. Understand thoroughly the operation of each machine, and check its operation before

proceeding.

2. Check the calibration of the Rockwell machines with standard calibration test blocks

for the scale selected.

3. Using the appropriate scale;

 Check the hardness of each test specimen on a Rockwell test machine


 Tabulate the results
 Convert all readings to either Rb or Rc values

A modern Rockwell hardness testing machine.


RESULTS

HRC (for standard calibration block)

1; 35·98 HRC

2; 37·34 HRC

3; 36·37 HRC

4; 37·09 HRC

5; 36·77 HRC

6; 36·43 HRC

AVERAGE VALUE = (35·98+37·34+36·37+37·09+36·77+36·43) ÷ 6 = 36·66 HRC

HRC (for mild steel)

1; 15·36 HRC

2; 22·86 HRC

AVERAGE VALUE = 19·11 HRC

35·85 HRC – hardness of the standard calibration block.

ERROR TERM = 36·85 - 36·66 = 0·19 HRC

AVERAGE VALUE OF MILD STEEL HRC TEST = 19·11 + 0·19 = 19·30 HRC

On starting the test we add 0·19 HRC to it because it falls short of the standard HRC value.
 COLD WORKING

OBJECTIVE

 To study the effect of cold working on the hardness and microstructure of brass or any

other material.

EQUIPMENT;

 Rockwell Hardness Tester

 Rolling Mill

 Micrometer

 Hacksaw

 Belt Sander

 Polishing and Etching Equipment

 Metallograph

INTRODUCTION

Cold working promotes rapid change in the mechanical properties and in the appearance of

copper alloys. The most rapid method of accomplishing it is by cold rolling as in rolling sheet

and bar.

MATERIAL

One piece annealed cartridge brass ¾” (13·5mm) ˟ 4” (50mm) ˟ 1/8 ” (3·125mm) thick.
THEORY;

STRAIN HARDENING

Strain hardening is the phenomenon whereby a ductile metal becomes harder and stronger as

it is plastically deformed. Sometimes it is also called work hardening, or, because the

temperature at which deformation takes place is ‘‘cold’’ relative to the absolute melting

temperature of the metal, cold working. Most metals strain harden at room temperature. It is

sometimes convenient to express the degree of plastic deformation as percent cold work

rather than as strain. Percent cold work (%CW) is defined as

%CW = { Ao−AoAd }˟ 100


where A0 is the original area of the cross section that experiences deformation, and

Ad is the area after deformation. Figures 1a and 1b demonstrate how steel, brass, and copper

increase in yield and tensile strength with increasing cold work. The price for this

enhancement of hardness and strength is in the ductility of the metal. This is shown in Figure

1c, in which the ductility, in percent elongation, experiences a reduction with increasing

percent cold work for the same three alloys. The influence of cold work on the stress–strain

behaviour of a steel is vividly portrayed in Figure 2. The strain hardening phenomenon is

explained on the basis of dislocation. The dislocation density in a metal increases with

deformation or cold work, due to dislocation multiplication or the formation of new

dislocations, as noted previously. Consequently, the average distance of separation between

dislocations decreases—the dislocations are positioned closer together. On the average,

dislocation–dislocation strain interactions are repulsive. The net result is that the motion of a

dislocation is hindered by the presence of other dislocations. 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.
Figure 1; for 1040 steel, brass, and copper, (a) the increase in yield strength, (b) the increase
in tensile strength, (c) the decrease in ductility (%EL) with percent cold work.
Figure 2. the influence of cold work on the stress-strain behaviour for a low-carbon steel

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 an

annealing heat treatment.

strain hardening exponent, is a measure of the ability of a metal to strain harden; the larger

its magnitude, the greater the strain hardening for a given amount of plastic strain.

These properties and structures may revert back to the precold-worked states by appropriate

heat treatment (sometimes termed an annealing treatment). Such restoration results from two

different processes that occur at elevated temperatures: recovery and recrystallization,

which may be followed by grain growth.

RECOVERY

During recovery, some of the stored internal strain energy is relieved by virtue of dislocation

motion (in the absence of an externally applied stress), as a result of enhanced atomic

diffusion at the elevated temperature. There is some reduction in the number of dislocations,
and dislocation configurations are produced having low strain energies. In addition, physical

properties such as electrical and thermal conductivities and the like are recovered to their pre-

cold-worked states.

RECRYSTALLIZATION

Even after recovery is complete, the grains are still in a relatively high strain energy state. Re-

crystallization is the formation of a new set of strain-free and equiaxed grains (i.e., having

approximately equal dimensions in all directions) that have low dislocation densities and are

characteristic of the pre cold-worked condition. The driving force to produce this new grain

structure is the difference in internal energy between the strained and unstrained material.

The new grains form as very small nuclei and grow until they completely replace the parent

material, processes that involve short-range diffusion. Also, during re-crystallization, the

mechanical properties that were changed as a result of cold working are restored to their pre

cold-worked values; that is, the metal becomes softer, weaker, yet more ductile. Some heat

treatments are designed to allow re-crystallization to occur with these modifications in the

mechanical characteristics Re-crystallization is a process the extent of which depends on both

time and temperature. The degree (or fraction) of re-crystallization increases with time.

Table 1. Re-crystallization and melting temperatures for various metals and alloys.
GRAIN GROWTH

After re-crystallization is complete, the strain-free grains will continue to grow if the metal

specimen is left at the elevated temperature; this phenomenon is called grain growth. Grain

growth does not need to be preceded by recovery and re-crystallization; it may occur in all

polycrystalline materials, metals and ceramics alike.

An energy is associated with grain boundaries. As grains increase in size, the total boundary

area decreases, yielding an attendant reduction in the total energy; this is the driving force for

grain growth. Grain growth occurs by the migration of grain boundaries. Obviously, not all

grains can enlarge, but large ones grow at the expense of small ones that shrink. Thus, the

average grain size increases with time, and at any particular instant there will exist a range of

grain sizes. Boundary motion is just the short-range diffusion of atoms from one side of the

boundary to the other. The directions of boundary movement and atomic motion are opposite

to each other as shown in Figure 3.

Figure 3. schematic representation of grain growth via atomic diffusion.


ADVANTAGES

1. Increases the hardness of the material at the expense of ductility.


2. Dimension accuracy.
3. Good surface finishing.

PROCEDURE

1. Using the hacksaw, cut the annealed cartridge brass into 7 pieces of about ½”
(12·5mm) length and mark them as 0, 10, 20, 30, 40, 50, 60.
2. Measure the Rb or Rc (Rockwell B or Rockwell C scale) of the piece marked “0”. This
means that sample “0” has no cold work. Take three readings and average them.
3. Using the rolling mill reduce the thickness of sample “10” by 10%. This means that
the sample “10” represents 10% cold work. Measure the hardness and record.
4. Continue rolling and measuring the hardness numbers of samples 20, 30, 40, 50 and
60 and record.
5. Examine the microstructure of the cold worked specimens, parallel to the rolling
direction.

RESULTS

Diameter of sample (mild steel): 9mm (0·9cm)

% OF ORIGINAL LENGTH % REDUCTION (mm) 9mm - % REDUCTION

10% of 9mm 0·9 8·1

20% of 9mm 1·8 7·2

30% of 9mm 2·7 6·3

40% of 9mm 3·6 5·4

50% of 9mm 4·5 4·5

60% of 9mm 5·4 3·6


 METALLOGRAPHY

OBJECTIVE

 To see how cold-work treatment affect material grain structure

MATERIALS

Control sample and Cold work treated sample

INTRODUCTION

This involves a study of the structural characteristics of mild steel (0·2% carbon) in relation

to its mechanical and physical properties.

Microstructure comprises;

 ways grains are arranged

 grain boundaries

 grain structures

MICROSCOPIC TECHNIQUES

In cause of rolling, grains get compact and work hardened; the material become elongated as

a result of ductility.

STEPS INVOLVED

1. SECTIONING

2. GRINDING (rough / smooth)- flattening the material by use of sandpaper to remove

scratches. Sandpaper grades include 80, 150, 220, 300, 600 micron-electrons. (600

micron-electrons is the smoothest and 80 micron-electrons is the roughest). As you

grind, water is applied to make temperature remain ambient (prevent heating).

3. POLISHING (cloth / powder)- to enhance a better surface finish.

4. ETCHING- dipping the material into a suitable chemical reagent to attack and

dissolve the grain boundary.

5. MICROSCOPY- Taking to the high power microscope to see the microstructures.


OPTICAL MICROSCOPY

With optical microscopy, the light microscope is used to study the microstructure; optical and

illumination systems are its basic elements. For materials that are opaque to visible light (all

metals and many ceramics and polymers), only the surface is subject to observation, and the

light microscope must be used in a reflecting mode. Contrasts in the image produced result

from differences in reflectivity of the various regions of the microstructure. Investigations of

this type are often termed metallographic, since metals were first examined using this

technique. Normally, careful and meticulous surface preparations are necessary to reveal

the important details of the microstructure. The specimen surface must first be ground and

polished to a smooth and mirror-like finish. This is accomplished by using successively finer

abrasive papers and powders. The microstructure is revealed by a surface treatment using an

appropriate chemical reagent in a procedure termed etching. The chemical reactivity of the

grains of some single-phase materials depends on crystallographic orientation. Consequently,

in a polycrystalline specimen, etching characteristics vary from grain to grain. Also, small

grooves form along grain boundaries as a consequence of etching. Since atoms along grain

boundary regions are more chemically active, they dissolve at a greater rate than those within

the grains. These grooves become discernible when viewed under a microscope because they

reflect light at an angle different from that of the grains themselves. When the microstructure

of a two-phase alloy is to be examined, an etchant is often chosen that produces a different

texture for each phase so that the different phases may be distinguished from each other.

ELECTRON MICROSCOPY

The upper limit to the magnification possible with an optical microscope is approximately

2000 diameters. Consequently, some structural elements are too fine or small to permit

observation using optical microscopy. Under such circumstances the electron microscope,

which is capable of much higher magnifications, may be employed.


(a)polished and etched grains as they might appear when viewed with an optical microscope.

(b)section taken through these grains showing how the etching characteristics and resulting

surface texture vary from grain to grain because of differences in crystallographic orientation.

(c)photomicrograph of a polycrystalline brass specimen.


Mild steel – Nitric acid (30%) + Ethanol (Nital) for Etching

If forged, the grain of the material will just only be compacted.

PROCEDURE

1. Flatten the material to remove scratches with the use of sandpaper.

2. Dip the material into a suitable reagent (acid).

3. Bring out the material after 10-15 seconds; we noticed the surface of the material no

longer glittered after being attacked by the acid (reagent).

4. Rinse the material in water.

5. Leave to dry for some time.

6. Render the material visible under the electron microscope.

OBSERVATION

20% - the cold work treated material is observed to be compacted and flowing in the
direction of roll.

30% & 40% - the higher the deformation, the grains become much more compacted and rate
of flow is increased

50%- complete deformation of material grain and the ferrites are almost indistinguishable.
 PHASE DIAGRAMS

OBJECTIVE
 Construction of equilibrium diagrams.

DEVELOPMENT OF MICROSTRUCTURE IN EUTECTIC ALLOYS

Depending on composition, several different types of microstructures are possible for the

slow cooling of alloys belonging to binary eutectic systems. These possibilities will be

considered in terms of the lead–tin phase diagram. The first case is for compositions ranging

between a pure component and the maximum solid solubility for that component at room

temperature [20ºC (70ºF)]. For the lead–tin system, this includes lead-rich alloys containing

between 0 and about 2 wt% Sn (for the α phase solid solution), and also between

approximately 99 wt% Sn and pure tin (for the ᵦ phase). For example, consider an alloy of

composition C1 (Figure 1) as it is slowly cooled from a temperature within the liquid-phase

region, say, 350ºC; this corresponds to moving down the dashed vertical line ww_ in the

figure. The alloy remains totally liquid and of composition C1 until we cross the liquidus line

at approximately 330ºC, at which time the solid α phase begins to form. While passing

through this narrow α + L phase region, solidification proceeds ; that is, with continued

cooling more of the solid α forms. Furthermore, liquid and solid-phase compositions are

different, which follow along the liquidus and solidus phase boundaries, respectively.

Solidification reaches completion at the point where ww1 crosses the solidus line. The

resulting alloy is polycrystalline with a uniform composition of C1 , and no subsequent

changes will occur upon cooling to room temperature. This microstructure is represented

schematically by the inset at point c in Figure 1. The second case considered is for

compositions that range between the room temperature solubility limit and the maximum
solid solubility at the eutectic temperature. Let us examine an alloy of composition C2 as it is

cooled along the vertical line xx1 in Figure 2. Down to the intersection of xx1 and the solvus

line, changes that occur are similar to the previous case, as we pass through the

corresponding phase regions (as demonstrated by the insets at points d, e, and f ). Just above

the solvus intersection, point f, the microstructure consists of α grains of composition C2 .

Upon crossing the solvus line, the α solid solubility is exceeded, which results in the

formation of small ᵦ phase particles; these are indicated in the microstructure inset at point g.

With continued cooling, these particles will grow in size because the mass fraction of the ᵦ

phase increases slightly with decreasing temperature. The third case involves solidification of

the eutectic composition, 61.9 wt% Sn (C3 in Figure 3). Consider an alloy having this

composition that is cooled from a temperature within the liquid-phase region (e.g., 250ºC)

down the vertical line yy1 in Figure 10.11. As the temperature is lowered, no changes occur

until we reach the eutectic temperature, 183ºC. Upon crossing the eutectic isotherm, the

liquid transforms to the two α and ᵦ phases. This transformation may be represented by the

reaction

L ( 61∙ 9 wt % Sn ) ↔ α ( 18 ∙3 wt %Sn ) + β ( 97∙ 8 Wt %Sn )

in which the α and ᵦ phase compositions are dictated by the eutectic isotherm

end points.

During this transformation, there must necessarily be a redistribution of the lead and tin

components, inasmuch as the α and ᵦ phases have different compositions neither of which is

the same as that of the liquid (as indicated in Equation 1). This redistribution is accomplished

by atomic diffusion. The microstructure of the solid that results from this transformation

consists of alternating layers (sometimes called lamellae) of the α and ᵦ phases that form

simultaneously during the transformation. This microstructure, represented schematically in

Figure 3, point i, is called a eutectic structure, and is characteristic of this reaction.


The micro-structural change that accompanies this eutectic transformation is represented

schematically in Figure 4; here is shown the α - ᵦ layered eutectic growing into and replacing

the liquid phase. The process of the redistribution of lead and tin occurs by diffusion in the

liquid just ahead of the eutectic–liquid interface. The arrows indicate the directions of

diffusion of lead and tin atoms; lead atoms diffuse toward the α-phase layers since this α

phase is lead-rich (18.3 wt% Sn–81.7 wt% Pb); conversely, the direction of diffusion of tin is

in the direction of the ᵦ , tin-rich (97.8 wt% Sn–2.2 wt% Pb) layers. The eutectic structure

forms in these alternating layers because, for this lamellar configuration, atomic diffusion

of lead and tin need only occur over relatively short distances.

Fig 4. schematic representations of the formation of the eutectic structure for the lead-tin

system. Directions of diffusion of tin and lead atoms are indicated by the arrows.
Figure 2, schematic representations of the equilibrium microstructure for a lead-tin alloy of
composition c2 as it is cooled from the liquid-phase region.
Figure 1. schematic representations of the equilibrium microstructure for a lead-tin alloy of
composition c1 as it is cooled from the liquid phase region.
Figure 3.schematic representations of the equilibrium microstructure for a lead-tin alloy of
eutectic composition c3 above and below the eutectic temperature.
The fourth and final micro-structural case for this system includes all compositions other than

the eutectic that, when cooled, cross the eutectic isotherm. Consider, for example, the

composition C4 , Figure 5, which lies to the left of the eutectic; as the temperature is lowered,

we move down the line zz1, beginning at point j. The micro-structural development between

points j and l is similar to that for the second case, such that just prior to crossing the eutectic

isotherm (point l), the α and liquid phases are present having compositions of approximately

18.3 and 61.9 wt% Sn, respectively, as determined from the appropriate tie line. As the

temperature is lowered to just below the eutectic, the liquid phase, which is of the eutectic

composition, will transform to the eutectic structure (i.e., alternating α and ᵦ lamellae);

insignificant changes will occur with the α phase which formed during cooling

through the α + L region. This microstructure is represented schematically by the inset at

point m in Figure 5. Thus, the α phase will be present both in the eutectic structure and also as

that phase that formed while cooling through the α + L phase field. To distinguish one α from

the other, that which resides in the eutectic structure is called eutectic α, while the other that

formed prior to crossing the eutectic isotherm is termed primary α; both are labelled in

Figure 5. In dealing with microstructures, it is sometimes convenient to use the term Micro-

constituent, that is, an element of the microstructure having an identifiable and characteristic

structure. For example, in the point m inset, Figure 5, there are two micro-constituents,

namely, primary α and the eutectic structure. Thus, the eutectic structure is a micro-

constituent even though it is a mixture of two phases, because it has a distinct lamellar

structure, with a fixed ratio of the two phases. It is possible to compute the relative amounts

of both eutectic and primary α. Micro-constituents. Since the eutectic micro-constituent

always forms from the liquid having the eutectic composition, this micro-constituent may be

assumed to have a composition of 61.9 wt% Sn. Hence, the lever rule is applied using a tie

line between the α–(α + ᵦ) phase boundary (18.3 wt% Sn) and the eutectic composition.
Figure 5. schematic representations of the equilibrium microstructure for a lead-tin alloy of

eutectic composition c4 as it is cooled from the liquid phase region.

You might also like