Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 82

NARRATIVE REPORT

Item No. 1: Introduction to Steel

The use of iron dates back to about 1500 B.C., when primitive furnaces were used to heat the ore in

a charcoal fire. Ferrous metals were produced on a relatively small scale until the blast furnace was

developed in the 18th century. Iron products were widely used in the latter half of the 18th century and the

early part of the 19th century. Steel production started in mid-1800s, when the Bessemer converter was

invented. In the second half of the 19th century, steel technology advanced rapidly due to the development

of the basic oxygen furnace and continuous casting methods. More recently, computer-controlled

manufacturing has increased the efficiency and reduced the cost of steel production.

The structural steel industry consists of four components (AISC 2008):

1. producers of structural steel including hot-rolled structural shapes and hollow sections

2. service centers that function as warehouses and provide limited preprocessing of structural

material

3. structural steel fabricators that prepare the steel for the building process

4. erectors that construct steel frames on the project site

Civil and construction engineers will primarily interact with the service centers to determine the

availability of products required for a project, the fabricators for detail design drawings and preparation of the

product, and the erectors on the job site.

Currently, steel and steel alloys are used widely in civil engineering applications. In addition, wrought

iron is still used on a smaller scale for pipes, as well as for general blacksmith work. Cast iron is used for

pipes, hardware, and machine parts not subjected to tensile or dynamic loading.

1
Steel products used in construction can be classified as follows:

1. structural steel (Figure 3.1 vertical columns) produced by continuous casting and hot rolling for

large structural shapes, plates, and sheet steel

2. cold-formed steel (Figure 3.1 trusses and decking) produced by cold-forming of sheet steel into

desired shapes

3. fastening products used for structural connections, including bolts, nuts and washers

4. reinforcing steel (rebars) for use in concrete reinforcement (Figure 3.2)

5. miscellaneous products for use in such applications as forms and pans

2
Civil and construction engineers rarely have the opportunity to formulate steel with specific properties.

Rather, they must select existing products from suppliers. Even the shapes for structural elements are

generally restricted to those readily available from manufacturers. While specific shapes can be made to

order, the cost to fabricate low-volume members is generally prohibitive. Therefore the majority of civil

engineering projects, with the exception of bridges, are designed using standard steel types and structural

shapes. Bridges are a special case, in that the majority of bridge structures are fabricated from plate steel

rather than hot-rolled sections or hollow structural sections (AISC 2008).

Even though civil and construction engineers are not responsible for formulating steel products, it is

beneficial to understand how steel is manufactured and treated and how it responds to loads and

environmental conditions. This chapter reviews steel production, the iron–carbon phase diagram, heat

treatment, steel alloys, structural steel, steel fasteners, and reinforcing steel. The chapter also presents

common tests used to characterize the mechanical properties of steel. The topics of welding and corrosion

of steel are also

introduced.

3
Item No. 2: Steel Production

The overall process of steel production is shown in Figure 3.3. This process consists of the following

three phases:

1. reducing iron ore to pig iron

2. refining pig iron (and scrap steel from recycling) to steel

3. forming the steel into products

The materials used to produce pig iron are coal, limestone, and iron ore. The coal, after transformation

to coke, supplies carbon used to reduce iron oxides in the ore. Limestone is used to help remove impurities.

Prior to reduction, the concentration of iron in the ore is increased by crushing and soaking the ore. The iron

is magnetically extracted from the waste, and the extracted material is formed into pellets and fired. The

processed ore contains about 65% iron.

Reduction of the ore to pig iron is accomplished in a blast furnace. The ore is heated in the presence of

carbon. Oxygen in the ore reacts with carbon to form gases. A flux is used to help remove impurities. The

molten iron, with an excess of carbon in solution, collects at the bottom of the furnace. The impurities, slag,

float on top of the molten pig iron.

4
5
The excess carbon, along with other impurities, must be removed to produce high-quality steel. Using

the same refining process, scrap steel can be recycled. Two types of furnaces are used for refining pig iron

to steel:

1. basic oxygen

2. electric arc

The basic oxygen furnaces remove excess carbon by reacting the carbon with oxygen to form gases.

Lances circulate oxygen through the molten material. The process is continued until all impurities are

removed and the desired carbon content is achieved.

Electric furnaces use an electric arc between carbon electrodes to melt and refine the steel. These

plants require a tremendous amount of energy, and are used primarily to recycle scrap steel. Electric furnaces

are frequently used in minimills, which produce a limited range of products. In this process, molten steel is

transferred to the ladle. Alloying elements and additional agents can be added either in the furnace or the

ladle.

During the steel production process, oxygen may become dissolved in the liquid metal. As the steel

solidifies, the oxygen can combine with carbon to form carbon monoxide bubbles that are trapped in the steel

and can act as initiation points for failure. Deoxidizing agents, such as aluminum, ferrosilicon and manganese,

can eliminate the formation of the carbon monoxide bubbles. Completely deoxidized steels are known as

killed steels. Steels that are generally killed include:

■ Those with a carbon content greater than 0.25%

■ All forging grades of steels

■ Structural steels with carbon content between 0.15 and 0.25 percent

6
■ Some special steel in the lower carbon ranges

Regardless of the refining process, the molten steel, with the desired chemical composition, is then

either cast into ingots (large blocks of steel) or cast continuously into a desired shape. Continuous casting

with hot rolling is becoming the standard production method, since it is more energy efficient than casting

ingots, as the ingots must be reheated prior to shaping the steel into the final product. Cold-formed steel is

produced from sheets or coils of hot rolled steel which is formed into shape either through press-braking

blanks sheared from sheets or coils, or more commonly, by roll forming the steel through a series of dies. No

heat is required to form the shapes (unlike hot-rolled steel), and thus the name cold-formed steel. Cold-

formed steel members and other products are thinner, lighter, and easier to produce, and typically cost less

than their hot-rolled counterparts (Elhajj, 2001).

Methods for manufacturing steel have evolved significantly since industrial production began in the

late 19th century. Modern methods, however, are still based on the same premise as the original Bessemer

Process, which uses oxygen to lower the carbon content in iron.

Today, steel production makes use of recycled materials as well as traditional raw materials, such

as iron ore, coal, and limestone. Two processes, basic oxygen steelmaking (BOS) and electric arc furnaces

(EAF), account for virtually all steel production.

Modern steelmaking can be broken down into six steps:

Ironmaking, the first step, involves the raw inputs of iron ore, coke, and lime being melted in a blast

furnace. The resulting molten iron also referred to as hot metal still contains 4-4.5 percent carbon and other

impurities that make it brittle.

7
Primary steelmaking has two primary methods: BOS (Basic Oxygen Furnace) and the more modern

EAF (Electric Arc Furnace) methods. BOS methods add recycled scrap steel to the molten iron in a converter.

At high temperatures, oxygen is blown through the metal, which reduces the carbon content to

between 0-1.5 percent. EAF methods, however, feed recycled steel scrap through use high-power electric

arcs (temperatures up to 1650 C) to melt the metal and convert it into high-quality steel.

Secondary steelmaking involves treating the molten steel produced from both BOS and EAF routes to

adjust the steel composition. This is done by adding or removing certain elements and/or manipulating the

temperature and production environment. Depending on the types of steel required, the following secondary

steelmaking processes can be used:

 Stirring

 Ladle furnace

 Ladle injection

 Degassing

 CAS-OB (composition adjustment by sealed argon bubbling with oxygen blowing)

Continuous casting sees the molten steel cast into a cooled mold causing a thin steel shell to solidify.

The shell strand is withdrawn using guided rolls and fully cooled and solidified. The strand is cut into desired

lengths depending on application; slabs for flat products (plate and strip), blooms for sections (beams), billets

for long products (wires) or thin strips.

In primary forming, the steel that is cast is then formed into various shapes, often by hot rolling, a process

that eliminates cast defects and achieves the required shape and surface quality. Hot rolled products are

divided into flat products, long products, seamless tubes, and specialty products.

8
Finally, it's time for manufacturing, fabrication, and finishing. Secondary forming techniques give the steel

its final shape and properties. These techniques include:

 Shaping (cold rolling), which is done below the metal's recrystallization point, meaning mechanical

stress—not heat—affects change.

 Machining (drilling)

 Joining (welding)

 Coating (galvanizing)

 Heat treatment (tempering)

 Surface treatment (carburizing)

Item No. 3: Steelmaking

Steelmaking is the process of producing steel from iron ore and/or scrap. In

steelmaking, impurities such as nitrogen, silicon, phosphorus, sulfur and excess carbon (most important

impurity) are removed from the sourced iron, and alloying elements such as manganese, nickel, chromium,

carbon and vanadium are added to produce different grades of steel. Limiting dissolved gases such

as nitrogen and oxygen and entrained impurities (termed "inclusions") in the steel is also important to ensure

the quality of the products cast from the liquid steel.

Steelmaking has existed for millennia, but it was not commercialized on a massive scale until the

late 19th century. The ancient craft process of steelmaking was the crucible process. In the 1850s and 1860s,

the Bessemer process and the Siemens-Martin process turned steelmaking into a heavy industry. Today

there are two major commercial processes for making steel, namely basic oxygen steelmaking, which has

9
liquid pig-iron from the blast furnace and scrap steel as the main feed materials, and electric arc

furnace (EAF) steelmaking, which uses scrap steel or direct reduced iron (DRI) as the main feed materials.

Oxygen steelmaking is fuelled predominantly by the exothermic nature of the reactions inside the

vessel; in contrast, in EAF steelmaking, electrical energy is used to melt the solid scrap and/or DRI materials.

In recent times, EAF steelmaking technology has evolved closer to oxygen steelmaking as more chemical

energy is introduced into the process.

Item No. 4: Cementation Process

The process begins with wrought iron and charcoal. It uses one or more long stone pots inside a

furnace. Typically, in Sheffield, each was 14 feet by 4 feet and 3.5 feet deep. Iron bars and charcoal are

packed in alternating layers, with a top layer of charcoal and then refractory matter to make the pot or "coffin"

airtight. Some manufacturers used a mixture of powdered charcoal, soot and mineral salts, called cement

10
powder. In larger works, up to 16 tons of iron was treated in each cycle,though it can be done on a small

scale, such as in a small furnace or blacksmith's forge.

Depending on the thickness of the iron bars, the pots were then heated from below for a week or

more. Bars were regularly examined and when the correct condition was reached the heat was withdrawn

and the pots were left until cool—usually around

fourteen days. The iron had gained a little over

1% in mass from the carbon in the charcoal, and

had become heterogeneous bars of blister steel.

The bars were then shortened, bound,

heated and forge welded together to

become shear steel.It would be cut and re

welded multiple times, with each new weld

producing a more homogeneous, higher quality

steel. This would be done at most 3-4 times, as

more is unnecessary and could potentially

cause carbon loss from the steel.

The cementation process of making

steel is also called the converting process. This

process consists in impregnating bars of

wrought iron or soft steel with carbon, at a

temperature below its melting point, and was used (chiefly in England) for the production of high carbon bars

to be employed in the manufacture of crucible steel or shear steel. The bars were usually of pure Swedish

iron made by the Walloon process. They are packed in layers, separated by charcoal (sometimes called

11
cement) in fire-brick chambers (converting pots) heated externally by flues, and forming part of the cementing

furnace. The top of the pot is closed with an arch of wheel swarf, which later frits and forms an air-tight cover.

The furnace attains its full temperature in about 3 to 4 days, at which it is maintained about 7 to 8 days for

mild heats, about 9 days for medium heats, and about 11 days for high carbon heats; the cooling down

requires about 4 to 6 days. To test the progress of the operation, trial bars (test bars or tap bars) are drawn

at intervals through a special small aperture, provided for the purpose, and examined. If wrought iron has

been employed, the finished bars will be found covered with blisters formed by the reaction between the

contained slag and the carbon, from which comes the name blister bar or blister steel; at one time this was

sometimes termed German steel.

12
Item No. 5: Crucible Process

Crucible steel is steel made by melting pig iron (cast iron), iron, and sometimes steel, often along

with sand, glass, ashes, and other fluxes, in a crucible. In ancient times steel and iron were impossible to

melt using charcoal or coal fires, which

could not produce temperatures high

enough. However, pig iron, having a

higher carbon content thus a lower

melting point, could be melted, and by

soaking wrought iron or steel in the

liquid pig-iron for a long time, the

carbon content of the pig iron could be

reduced as it slowly diffused into the

iron. Crucible steel of this type was

produced in South and Central Asia

during the medieval era. This generally

produced a very hard steel, but also a

composite steel that was

inhomogeneous, consisting of a very

high-carbon steel (formerly the pig-iron)

and a lower-carbon steel (formerly the wrought iron). This often resulted in an intricate pattern when the steel

was forged, filed or polished, with possibly the most well-known examples coming from the wootz steel used

in Damascus swords. Due to the use of fluxes the steel was often much higher in quality (lacking impurities)

and in carbon content compared to other methods of steel production of the time.

13
Crucible process, technique for producing fine or

tool steel. The earliest known use of the technique

occurred in India and central Asia in the early 1st

millennium CE. The steel was produced by heating

wrought iron with materials rich in carbon, such as

charcoal in closed vessels. It was known as wootz and

later as Damascus steel. About

800 CE the crucible process appeared in northern

Europe—likely as a result of trade contact with the Middle East—where it was used to make the high-

quality Ulfbehrt swords used by the Vikings. The process was devised again in Britain about 1740

by Benjamin Huntsman, who heated small pieces of carbon steel in a closed fireclay crucible placed in

a coke fire. The temperature he was able to achieve (1,600 °C [2,900 °F]) was high enough to permit melting

steel for the first time, producing a homogeneous metal of uniform composition that he used to

manufacture watchand clock springs. After 1870 the Siemens regenerative gas furnace replaced the coke-

fire furnace; it produced even higher temperatures. The Siemens furnace had a number of combustion holes,

each holding several crucibles, and heated as many as 100 crucibles at a time. All high-quality tool

steel and high-speed steel was long made by the crucible process, but in the 20th century the electric

furnace replaced it in contries where electric power was cheap.

14
Item No. 6: Bessemer Process

Bessemer Process: In this process pig iron is melted in a cupola and poured into Bessemer converter

which is pear shaped and has a steel shell lined with refractory material. It’s pivoted on trunnions so as to

facilitate tilting, pouring or charging.

Once the above converter is charged with molten pig iron, a strong thrust of air is blasted across the

molten mass for about 20 minutes through nozzles provided at the bottom of the vessel. The process oxidizes

all traces of the carbon and silicon present, leaving the converter with pure iron.

After this the blasting of air is stopped and the specified amount of ferro-manganese is added to it for the

sake of including the recommended content of carbon and manganese to the steel.

The air blasted procedure is again initiated for some time, ensuring perfect mixing of the alloy.

The converter is then tilted so that the molten material can be

discharged into the ladles. In the final step the molten alloy is shifted

into rectangular moulds where it’s obtained in the form of solid ingots.

15
Item No. 7: Open-Hearth Process

Open Hearth Process: The specialty of open-hearth furnaces is the extreme heat that can be

obtained from them due to their regenerative process. The charge of pig iron, steel scrap, iron ore, and flux

are together kept in a shallow container with a flame burning above it. The process is initiated inside

reverbaratory gas-fired regenerative furnaces for greater efficiency.

Regenerators are placed below the furnace and positioned in two pairs. The pairs are heated

alternately through the passage of hot gases given out from the furnace in their route to the chimney. This

heat is retained by the regenerators and is reversed and given back to the furnace. This heat exchange

procedure helps the furnace to maintain high temperatures even with less fuels.

Once the furnace is charged with pig iron, pure oxidizing ores like haematite are added to it from time

to time, which helps oxidization and the removal of impurities like silicon, carbon, and manganese in the pig

iron. Spiegel is also introduced when the carbon content becomes less than 0.1%, and ferro-manganese

after the metal is tapped out into the ladle. Ferro-manganese becomes important for restoring malleability and

also for carburizing the iron.

16
The open-hearth furnace (OHF) uses the heat of combustion of gaseous or liquid fuels to convert a

charge of scrap and liquid blast-furnace iron to liquid steel. The high flame temperature required for melting is

obtained by preheating the combustion air and, sometimes, the fuel gas. Preheating is done in large, stovelike

regenerators or checker chambers, located beneath the furnace (see figure). These contain checker bricks

stacked in such a way that they absorb heat from furnace off-gases as they are directed through the chamber.

After one chamber has been heated for about 20 minutes, a sliding valve is activated, directing the off-gases

to the other chamber and simultaneously bringing air into the heated chamber. This combustion air, after

picking up the heat from the checker brick, then enters the furnace through an end-wall above the checker

chamber and burns the fuel, which also enters the furnace at the same wall. The combustion flames heat the

charge, and the off-gases, after moving across the hearth to the other end wall, are directed downward to

17
heat the other chamber. This cycle, with entry ports becoming exit ports, is reversed every 15 to 20 minutes.

After heating the regenerator, the off-gases flow through a heat-recovery boiler and a gas-cleaning system

before they are discharged into the atmosphere through a stack.

The OHF itself consists of a shallow, rectangular hearth that holds the charge, liquid steel,

and slag (see figure). Depending on the furnace size, the long front wall on the charging side usually has

three to seven rectangular openings fitted with water-cooled doors. These are used for charging scrap and

iron, adding flux and alloying agents, running off slag, conducting tests, and repairing the hearth refractory.

On the opposite side of the furnace, at the back wall, is the taphole and a spout for tapping into one or two

ladles. The two end walls are used as inlets or outlets for gas and air, and they also hold the injection burners

for heavy oil, tar, or natural gas, when used.

Above the hearth, an arched roof contains the flames and reflects the heat onto the melt. Since thermal

exposure is intense here, the roof is made of high-grade chrome-magnesite refractory bricks suspended from

a steel structure. Many furnaces have one to four retractable oxygen lances installed in the roof to increase

the flame temperature and melting rate.

OHFs vary considerably in size, having been built for heats of 10 to 600 tons. The hearth of a 150-

ton-capacity OHF is approximately 15 metres long and 5 metres wide. There are often up to a dozen furnaces

in one shop, lined up end wall to end wall only a few metres apart with all front doors on one line and at the

same level. This permits the charging of all furnaces by the same charging machine, crane, and rail system.

Bulk materials, such as scrap, cold blast-furnace iron, ore, limestone, coke, and alloying agents, are charged

through the furnace doors in small boxes of one- to two-cubic-metre capacity. The boxes are brought to the

OHF on small railroad buggies, and a charging machine then moves one box after another through a door,

turns them over, and dumps their contents onto the hearth.

The process. When starting a heat, the hearth is first covered by limestone flux, and scrap is charged

on top of that. Charging a large furnace may require two to three hours and as many as 150 full charging

18
boxes. The burners and oxygen lances are on during charging, so that most of the scrap has been melted

by the end of the scrap charge. Afterward a special pouring spout is placed into one of the doors, and blast-

furnace iron is slowly poured from an iron ladle into the melt. Composition of the metallic charge varies from

20 percent scrap and 80 percent blast-furnace iron to 100 percent scrap; a common proportion is 60 percent

iron and 40 percent scrap.

Carbon in the poured iron reacts with the oxidized molten scrap and generates the carbon

monoxide boil. This stirs the shallow (about 300 millimetres deep) bath and accomplishes a high heat

transfer and a good mixing of the slag and metal. The carbon monoxide boil may last two to three hours,

during which time carbon is oxidized and lowered, slag is flushed off through the doors, and the temperature

is raised. Increasing heat causes the limestone charged beneath the scrap to calcine and release carbon

dioxide, according to the following reaction:

This begins the lime boil, which has a beneficial stirring effect similar to that of the carbon monoxide

boil. After about one hour, the calcined lime rises through the melt and is dissolved in the slag.

19
During the subsequent refining period, flux and alloys are added, and oxygen or carbon is injected to lower

or raise the carbon content. When temperature and chemical composition are in the specified range, the

furnace is tapped by blowing the taphole open with a small explosive charge. Tap-to-tap time is six to nine

hours, often including one hour for inspection, cleaning, and hearth repair. After 200 to 300 heats, there is

usually a three-day process of checker cleaning and more extensive repair work. The roof is usually replaced

after about 1,000 heats, which shuts the furnace down for one week. The hearth, being made up anew after

every heat, lasts many years.

20
Item No. 8: Electric Process

Electric Process: In this process electric arc or electric high frequency furnaces are used. In electric

arc furnaces which are more common among the two processes, high voltage electric arc struck between

carbon electrodes and the charge becomes the source of a very high temperature. The charge is collected

directly from an open hearth furnace, the intense arc heat keeps the charge in its molten state, and the

impurities are removed in the form of slag.

The high frequency furnace is based on the principle that when high frequency alternating current is

applied to steel, eddy currents starts flowing in them. If this induction is made very strong, it can heat up the

steel and melt it.

Electric furnaces are more advantageous compared to the other steel manufacturing processes due

to the absence of evolving gases, fumes, etc., which normally become a major problem with fuel operated

furnaces.

21
Process. After tapping a heat, the roof is moved away, and the hearth is inspected and, when

necessary, repaired. An overhead crane then charges the furnace with scrap from a cylindrical bucket that is

open on the top for loading and fitted with a drop bottom for quick charging. Scrap buckets are loaded in such

a manner as to assure a cushioning of heavy scrap when the load drops onto the hearth in order to obtain

good electrical conductivity in the charge, low risk of electrode breakage, and good furnace wall protection

during meltdown. Carbon and slag formers are sometimes added to the charge to prevent overoxidation of

the steel and to quicken slag formation. After charging one bucket, the roof is moved back to the furnace,

and the electrodes are lowered. Meltdown begins with a low power setting until the electrodes have burned

themselves into the light scrap on top of the charge, protecting the sidewalls from overheating during higher-

power meltdown. Leaving some scrap unmelted at the furnace wall for its protection, a second bucket is

charged and the same meltdown procedure is followed. Melting very light scrap sometimes requires the

charging of a third or even fourth bucket.

22
After meltdown, the carbon level in the steel is about 0.25 percent above the final tap level, which

prevents overoxidation of the melt. By this time a basic slag has formed, typically consisting of 55 percent

lime, 15 percent silica, and 15 to 20 percent iron oxide. Slag foaming is often generated by injecting carbon

or a lime-carbon mixture, which reacts with the iron oxide in the slag to produce carbon monoxide gas. This

foam shields the sidewall and permits a higher power setting. As required, the carbon content of the steel is

either decreased by oxygen blowing or increased by carbon injection. Samples are taken, the temperature is

checked, additions are made, and, when all conditions are right, the furnace is tapped by rotating it forward

so that the steel flows over the spout or through the vertical taphole into a ladle. When slag appears, a quick

back tilt is applied and the slag is poured through the rear door of the furnace into a slag pot. Some shops

leave 15 percent of the liquid steel in the furnace. This “hot heel” practice permits complete slag separation.

Very clean steel—i.e., with low oxygen and sulfur content—can be produced in the EAF by a two-

slag practice. After removal of slag from the first oxidizing meltdown, new slag formers are added that contain

carbon or aluminum or both as reducing agents. The new reducing slag may consist of 65 percent lime, 20

percent silica, calcium carbide or alumina (or all three), and practically no iron oxide. Alloys, which oxidize

easily, are added at this time to minimize losses and to improve metallurgical control. Refining continues

under the reducing slag until the heat is ready for tapping. Total heat time is one to four hours, depending on

the type of steel made—that is, on the amount of refining applied and auxiliary heating used. Many shops do

not apply a two-slag practice but treat the steel, after scrap meltdown and tapping, in ladle treatment stations.

These secondary metallurgical plants, discussed below, allow the EAF to run only as a highly efficient scrap

melter.

From time to time, as the arc erodes their tips and the high-temperature furnace atmosphere oxidizes their

bodies, new electrodes are added to the top of the electrode strings at the furnace. Electrodes are consumed

at the rate of three to six kilograms per ton of steel, depending on the type of operation.

23
The furnace. The electric-arc furnace (EAF) is a squat, cylindrical vessel made of heavy steel plates.

It has a dish-shaped refractory hearth and three vertical electrodes that reach down through a dome-shaped,

removable roof (see figure). The shell diameter of a 10-, 100-, and 300-ton EAF is approximately 2.5, 6, and

9 metres. The shell sits on a hydraulically operated rocker that tilts the furnace forward for tapping and

backward for slag removal. The bottom—i.e., the hearth—is lined with tar-bonded magnesite bricks and has

on one side a slightly inclined taphole and a spout or, as shown in the figure, an oval hearth and a vertical

taphole. With this latter arrangement, a furnace needs be tilted only 10° for tapping, producing a tight and

short tap stream that decreases heat loss and reoxidation of the liquid steel. Before charging, the vertical

taphole is closed from the outside by a movable bottom plate and is filled with refractory sand.

Most furnace walls are made of replaceable, water-cooled panels; these are covered inside by

sprayed-on refractories and slag for protection and to keep heat loss down. The roof is also made of water-

cooled panels and has three circular openings, equally spaced, for insertion of the cylindrical electrodes.

Another large roof opening, the so-called fourth hole, is used for off-gas removal. Additional openings in the

24
furnace wall, with water-cooled doors, are used for lance injection, sampling, testing, inspection, and repair.

The roof and electrodes can be lifted and moved away for charging scrap and for hearth maintenance.

The graphite electrodes, produced to high standards by a specialized industry, are actually strings

of individual electrodes bolted end to end by short graphite nipples. This is done because shorter electrodes

are easier to manufacture, transport, and handle. Electrode diameters depend on furnace size; a 100-ton

EAF typically uses 600-millimetre electrodes. Three electrode strings are each clamped to arms that extend

over the furnace roof and that are bolted to a vertically movable mast located beside the furnace. The mast

controls the distance between each electrode tip and the scrap or melt, thereby regulating the arc length and

current flow. Power-supply equipment—normally a step-down transformer, vacuum circuit breakers, a tap

changer for electrode voltage control, and a furnace transformer—is installed in a concrete vault a short

distance from the furnace. Heavy water-cooled cables and the power-carrying arms connect the furnace

transformer with the electrodes.

EAF plants are smaller and less expensive to build than integrated steelmaking plants, which, in

addition to basic oxygen furnaces, contain blast furnaces, sinter plants, and coke batteries for the making of

iron. EAFs are also cost-efficient at low production rates—e.g., 150,000 tons per year—while basic oxygen

furnaces and their associated blast furnaces can pay for themselves only if they produce more than 2,000,000

tons of liquid steel per year. Moreover, EAFs can be operated intermittently, while a blast furnace is best

operated at very constant rates. The electric power used in EAF operation, however, is high, at 360 to 600

kilowatt-hours per ton of steel, and the installed power system is substantial. A 100-ton EAF often has a 70-

megavolt-ampere transformer.

25
The Electric Arc Furnace (EAF) uses only cold scrap metal. The process was originally used solely

for making high quality steel, such as those used for machine tools and spring steel, as it gave more precise

control over the composition. Today, however, it is also employed in making more widely used steels,

including alloy and stainless grades as well as some special carbon and low-alloy steels. Modern electric arc

furnaces can make up to 150 tonnes of steel in a single melt.

The electric arc furnace consists of a circular bath with a movable roof, through which three graphite

electrodes can be raised or lowered. At the start of the process, the electrodes are withdrawn and the roof

swung clear. The steel scrap is then charged into the furnace from a large steel basket lowered from an

overhead travelling crane. When charging is complete, the roof is swung back into position and the electrodes

lowered into the furnace. A powerful electric current is passed through the charge, an arc is created, and the

26
heat generated melts the scrap. Lime and fluorspar are added as fluxes and oxygen is blown into the melt.

As a result, impurities in the metal combine to form a liquid slag.

The Electric Arc Furnace (EAF). Samples of the steel are taken and analysed to their check composition

and, when the correct composition and temperature have been achieved, the furnace is tapped rapidly into a ladle.

Final adjustments to precise customer specification can be made by adding alloys during tapping or, subsequently, in

a secondary steel making unit.

(Because EAF steel is made with 100% scrap, there is a temptation to specify this in preference

to BOS steel in a well meaning effort to reduce environmental impact. This is discouraged by the metals

industries and the reasons are explained in the Resources section).

Item No. 8: Secondary Steelmaking

The ladele. An open-topped cylindrical container made of heavy steel plates and lined with refractory,

the ladle is used for holding and transporting liquid steel. Here all secondary metallurgical work takes place,

including deslagging and reslagging, electrical heating, chemical heating or cooling with scrap, powder

injection or wire feeding, and stirring with gas or with electromagnetic fields. The ladle receives liquid steel

27
during tapping while sitting on a stand beneath the primary steelmaking furnace. It is moved by cranes, ladle

cars, turntables, or turrets. A ladle turret has two liftable forks, usually 180° apart, that revolve around a tower,

each fork capable of holding a ladle. Ladles have two heavy trunnions on each side for crane pickup. Support

plates under each trunnion are used for setting the ladles onto stands or ladle cars.

The shell. The side wall of a ladle is slightly cone-shaped, with the larger diameter on top for easy

removal of a skull—i.e., solidified steel and slag. A ladle capable of holding 200 tons of steel has an outside

diameter of approximately four metres and is about five metres high. Inside the ladle there is usually a 60-

millimetre-thick refractory safety lining next to the shell. The working lining, that part contacting the steel and

slag, is 180 to 300 millimetres thick, depending on ladle size and location in the ladle. The lining thickness

and type of brick in one ladle are often different to counteract increased wear at certain locations—for

example, at the impact area of the tapping stream or at the slag line. This results in more equal wear on the

ladle lining and an extended ladle service life.

Sometimes, fired clay bricks are used because they bloat—that is, they expand during heating and seal the

joints between them. Their thermal shock resistance is high, but their resistance to slag corrosion is low, so

that the working lining has to be replaced every 6 to 12 heats. Because ladle rebricking takes about eight

hours, up to 12 ladles are sometimes in use in large steelmaking shops in order to assure availability. For

ladle operations requiring longer holding times, higher-grade refractory linings are made of high alumina or

magnesia bricks. These give greater slag resistance, but they do not bloat and are less resistant to thermal

shock. For these reasons, they are kept hot at special preheating stations. Ladles that use these bricks have

service lives of up to 80 heats, so that fewer ladles are required. Preheating also decreases the heat loss of

liquid steel during tapping and holding.

Tapping. Except for very small ladles, which pour over the lip and a spout or through a teapot

arrangement when tilted, most ladles have a funnel-shaped nozzle with a closing device installed in the

28
bottom. Depending on ladle size, these nozzles have an orifice diameter of 15 to 100 millimetres and are

made of high-grade refractory material. Often they are opened and closed by a vertical steel stopper rod,

which is enclosed in refractory sleeves and partly immersed in the liquid steel. The head of the stopper rod

closes the nozzle and is lifted a specific distance for controlling the flow rate; on top it is connected to a

vertical slide that is either manually operated by a lever or remotely controlled from the crane pulpit.

Many shops use a slide-gate nozzle, which consists, in principle, of a fixed upper and a movable

lower refractory plate. Both plates have holes that are adjusted relative to each other for closed, throttled,

and full-open position. The lower plate is hydraulically shifted and is usually replaced after every heat. In a

similar system, an old plate is pushed out by a new plate while pouring, and flow control is accomplished by

using bottom plates with different orifice diameters. Having the entire flow-control system on the outside of

the ladle and the inside of the ladle completely unrestricted is necessary for operating with long holding times

and for certain steel treatments conducted in the ladle.

Stirring and Storing. Ladles are often built with one or more permeable refractory bottom blocks and

argon hookups for gas stirring. Ladles can also be placed against an electromagnetic stirring coil installed on

a ladle car; in this case, their shells are made of a nonmagnetic alloy.

A number of shops use ladle lids to limit the liquid-steel heat loss. Lid-handling systems are normally

mechanized, and removing, storing, and placing lids onto the ladles is done automatically.

Ladle metallurgy. The carrying out of metallurgical reactions in the ladle is a common practice in

practically all steelmaking shops, because it is cost-efficient to operate the primary furnace as a high-speed

melter and to adjust the final chemical composition and temperature of the steel after tapping. Also, certain

metallurgical reactions, for reasons of equipment design and operation, are more efficiently performed in the

29
ladle. The simplest form of steel treatment in the ladle takes place when the mixing effect of the tapping

stream is used to add deoxidizers, slag formers, and small amounts of alloying agents. These materials are

either placed into the ladle before tapping or are injected into the tapping stream.

Controlling temperature. Deoxidation reactions carried out in the ladle are exothermic and thus raise

the temperature of the liquid steel, but the steel also loses heat by radiation from the top surface, by heating

of the ladle lining, and by heat flux through the lining and shell. Temperature drops that take place when just

holding the steel can range from 0.3° to 2° C per minute. (Small ladles, owing to their high surface-to-volume

ratio, have a greater temperature loss than large ladles.) The rate of temperature drop then slows as the

refractories become heated and a steady flow of heat prevails through the lining and slag layer.

Tapping at the right temperature is necessary in order to meet critical temperature windows for

teeming or casting operations. Heat losses during and after tap can usually be predicted by computer, using

a process model that considers the temperature and configuration of the tap stream, the thermal condition of

the ladle before tap, the thicknesses of the ladle lining and slag layer, the expected holding times and stirring

conditions, and the thermal effects of alloying additions. Actual control over steel temperature can be

achieved in a ladle furnace (LF). This is a small electric-arc furnace with an 8- to 25-megavolt-ampere

transformer, three electrodes for arc heating, and the ladle acting as the furnace shell—as shown in A in

the figure. Argon or electromagnetic stirring is applied for better heat transfer. Most LFs can raise the

temperature of the steel by 4° C per minute, and several shops accomplish an increase of 4° to 6° C by

inducing a strong exothermic chemical reaction (for instance, by feeding aluminum and injecting oxygen) at

the stirring station. Subsequent argon stirring removes most of the alumina inclusions formed by this process.

Both heating technologies permit long holding times of full ladles and improve the continuous caster

operation.

30
Slag removal. Keeping furnace slags on the molten steel too long can result in a reversion of

elements such as phosphorus back into the steel. To avoid this, slag can be removed at slag-skimming

stations, where the ladle is tilted forward and a rake scrapes the slag into a slag pot parked beneath the ladle.

Some shops use a vacuum system, which sucks the slag off the liquid steel and granulates it instantaneously.

In either case, after slag removal the steel is covered with slag formers or an insulating layer to minimize heat

loss and reoxidation. Special equipment is used to quickly place a blanket of material on the steel surface.

Stirring and injecting

In most continuous casting operations, it is necessary to maintain minimal fluctuation in steel

temperature, and this requires the use of a ladle stirring station to establish a uniform temperature and

chemical composition throughout the ladle. The steel can be stirred by argon injected through a refractory-

31
lined lance or through a permeable refractory block in the bottom of the ladle, or it can be stirred by an

electromagnetic coil.

Additions are usually made at the stirring station by a wire feeder, which runs a heavy wire at

controlled speed through a refractory-covered lance and into the steel. Aluminum wire is often used for

trimming; other materials, such as calcium-silicon, zirconium, and rare-earth metals, are often enclosed in

thin steel tubes and are fed by the same machines. The wires and filled tubes are normally shipped to steel

plants in large coils, but there are also machines that fill the tubes with the appropriate materials on-site.

Another widely used treatment is powder injection. Powdered metal is fluidized by argon in a

pressure vessel and injected by a refractory-lined lance deep into the liquid steel. Because powder has a

large contact surface area, it reacts quickly with the steel. Deep injection is beneficial when adding materials

such as calcium or magnesium, which evaporate at steelmaking temperature, because ferrostatic pressure

suppresses the evaporation of these metals for some time. Powders are shipped to the shop in sealed

containers or in special tank cars topped with inert gas.

Desulfurizing. Many powder-injection stations are used for desulfurization. One effective desulfurizer

is a calcium-silicon alloy containing 30 percent calcium. Metallic calcium desulfurizes by forming the very

stable compound calcium sulfide (CaS), and it is alloyed with silicon because pure calcium reacts

instantaneously with water and is therefore difficult to handle. Injecting four kilograms of calcium-silicon per

ton of steel can remove approximately three-quarters of the sulfur, so that the sulfur content will drop, for

example, from 0.016 to 0.004 percent. For steel grades that do not permit silicon additions, a magnesium-

lime mixture is used. Magnesium is a good desulfurizer, and it also acts as a deoxidizer by combining locally

with dissolved oxygen. This makes it possible for the lime to desulfurize the steel according to the following

reaction:

32
Like magnesium, lime has a double function, because it helps to prevent the very low-melting

magnesium powder from melting inside the lance.

Adding calcium accomplishes another important function. Sulfur is normally present in solidified steel

in the form of manganese sulfide inclusions, which are soft at hot-rolling temperatures and are rolled into long

strings or platelets. This results in poor physical properties of the steel in directions perpendicular to that of

the rolling. The addition of calcium improves these properties by forming strong inclusions, containing mainly

calcium sulfide, that are not plastic at hot-rolling temperatures. This phenomenon, called inclusion shape

control, can also be achieved by small additions of zirconium or rare earth.

Vacuum treatment. Exposing steel to vacuum conditions has a profound effect on all metallurgical

reactions involving gases. First, it lowers the level of gases dissolved in liquid steel. Hydrogen, for example,

is readily removed in a vacuum to less than two parts per million. Nitrogen is not as mobile in liquid steel as

hydrogen, so that only 15 to 30 percent is typically removed during a 20-minute vacuum treatment.

Another important process is vacuum decarburization and deoxidation. In

theory, oxygen and carbon, when dissolved in steel, react to form carbon monoxide until they

reach equilibrium at the following relationship:

This means that, under vacuum conditions (when there are only small amounts of carbon monoxide

in the surrounding gas and therefore little carbon monoxide pressure), carbon and oxygen will react

vigorously until they reach equilibrium at very low levels. For instance, liquid steel at 1 atmosphere pressure

may contain 0.043 percent carbon and 0.058 percent oxygen, but, if the pressure is lowered to 0.1

atmosphere, the two elements will react until they reach equilibrium at 0.014 percent carbon and 0.018

percent oxygen. Under a pressure as low as 0.01 atmosphere, equilibrium will be reached at 0.004 percent

carbon and 0.006 percent oxygen. In practical operation, the obtainable levels of carbon and oxygen are far

33
above equilibrium conditions, because the movement of carbon and oxygen atoms in liquid steel is time-

consuming and treatment time is limited. In addition, the steel is continuously reoxidized by multiple sources

of oxygen. Nevertheless, it is common practice to produce ultralow-carbon steel, containing less than 0.003

percent carbon, in 20 minutes at a vacuum treatment station under pressure of one torr. (In vacuum

technology, pressures are often expressed in torr, which is equivalent to the pressure of a column of one

millimetre of mercury. One atmosphere equals 760 torr.)

There are several types of vacuum treatment, their use depending on steel grade and required

production rates. In the tank degasser (shown in B in the figure), the ladle is placed in an open-top vacuum

tank, which is connected to vacuum pumps. The vacuum pumping system often consists of two or three

mechanical pumps, which lower the pressure to about 0.1 atmosphere, and four or five stage steam ejectors,

which bring the pressure to under 1 torr, or 0.0013 atmosphere. Practical treatment time is 20 to 30 minutes.

The ladles used in tank degassing stations are large and, when filled with steel, retain about one metre of

freeboard in order to contain the melt during a vigorous boil.

A modification of the tank degassers is the vacuum oxygen decarburizer (VOD), which has an oxygen

lance in the centre of the tank lid to enhancecarbon removal under vacuum. The VOD is often used to lower

the carbon content of high-alloy steels without also overoxidizing such oxidizable alloying elements

as chromium. This is possible because, in the pressure-dependent carbon-oxygen reaction outlined above,

oxygen reacts with carbon before it combines with chromium. The VOD is often used in the production of

stainless steels.

There are also tank degassers that have electrodes installed like a ladle furnace, thus permitting arc

heating under vacuum. This process is called vacuum arc degassing, or VAD.

For higher production rates (e.g., 25 ladles treated per day) and large ladles (e.g., 200 tons), a recirculation

degasser is used, as shown in C in the figure. This has two refractory-lined snorkels that are part of a high,

34
cylindrical, refractory-lined vacuum vessel and are immersed in the steel. As the system is

evacuated, atmospheric pressure pushes the liquid steel through the snorkels and up into the vessel. One

atmosphere lifts liquid steel about 1.3 metres. Injecting argon into one of the snorkels then circulates the steel

through the vessel, continuously exposing a portion of the steel to the vacuum. Recirculation facilities are

often very elaborate, using fast vessel-exchange systems or even two operating vessels at one station to

achieve high production rates. Some units also inject oxygen during vacuum treatment, through either the

side or the top of the vessel. This is done to speed up decarburization or, by simultaneously adding aluminum,

to increase the steel temperature. Some shops apply a similar system but use a vacuum vessel with only

one snorkel. Here, a portion of the steel in the ladle flows in and out of the vacuum vessel and is exposed to

the vacuum by a continuous raising and lowering of either the vessel or the ladle.

Item No. 9: Coke

Coke is produced from carefully selected grades of coal. Different grades of coal are stocked

separately and blended before transfer to coke ovens. The coal is heated, or ‘carbonised’ in the ovens until

it becomes coke. It is then removed from the oven, cooled and graded before use in the blast furnace. The

coal gas produced during carbonisation is collected and used as a fuel in the manufacturing process while

by-products such as tar, benzole and sulfur are extracted for further refining.

35
Coke, ore and sinter are

fed, or ‘charged’, into the top of

the blast furnace, together with

limestone. A hot air blast, from

which the furnace gets its name,

is injected through nozzles,

called ‘tuyeres’, in the base of the

furnace. The blast air may be

oxygen-enriched and coal or oil is

sometimes also injected to

provide additional heat and reduce coke requirements. The blast fans the heat in the furnace to white-hot

intensity, and the iron in the ore and sinter is melted out to form a pool of molten metal in the bottom, or

hearth, of the furnace.

The limestone combines with impurities and molten rock from the iron ore and sinter, forming a liquid

‘slag' which, being lighter than the metal, floats on top of it.

Coking coal is converted to coke by driving off impurities to leave almost pure carbon. The physical properties

of coking coal cause the coal to soften, liquefy and then resolidify into hard but porous lumps when heated

in the absence of air. Coking coal must also have low sulphur and phosphorous contents. Almost all

metallurgical coal is used in coke ovens.

The coking process consists of heating coking coal to around 1000-1100ºC in the absence of oxygen

to drive off the volatile compounds (pyrolysis). This process results in a hard porous material - coke. Coke is

produced in a coke battery, which is composed of many coke ovens stacked in rows into which coal is

36
loaded. The coking process takes place over long periods of time between 12-36 hours in the coke ovens.

Once pushed out of the vessel the hot coke is then quenched with either water or air to cool it before storage

or is transferred directly to the blast furnace for use in iron making.

Item No. 10: Smelting

Smelting, process by which a metal is obtained, either as the element or as a simple compound, from

its ore by heating beyond the melting point, ordinarily in the presence of oxidizing agents, such as air, or

reducing agents, such as coke. The first metal to be smelted in the ancient Middle East was

probably copper (by 5000 BCE), followed by tin, lead, and silver. To achieve the high temperatures required

for smelting, furnaces with forced-air draft were developed; for iron, temperatures even higher were required.

Smelting thus represented a major technological achievement. Charcoal was the universal fuel

until coke was introduced in 18th-century England. Meanwhile, the blast furnace had achieved a high state

of development.

37
In modern ore treatment, various preliminary steps are usually carried out before smelting in order

to concentrate the metal ore as much as possible. In the smelting process a metal that is combined

with oxygen—for example, iron oxide—is heated to a high temperature, and the oxide is caused to combine

with the carbon in the fuel, escaping as carbon monoxide or carbon dioxide. Other impurities, collectively

called gangue, are removed by adding a flux with which they combine to form a slag.

In modern copper smelting, a reverberatory furnace is used. Concentrated ore and a flux,

commonly limestone, are charged into the top, and molten matte—a compound of copper, iron, and sulfur—

and slag are drawn out at the bottom. A second heat treatment, in another (converter) furnace, is necessary

to remove the iron from the matte.

Smelting involves more than just melting the metal out of its ore. Most ores are the chemical

compound of the metal and other elements, such as oxygen (as an oxide), sulfur (as a sulfide), or carbon

and oxygen together (as a carbonate). To extract the metal, workers must make these compounds undergo

a chemical reaction. Smelting therefore consists of using suitable reducing substances that combine with

those oxidizing elements to free the metal.

Mined ores are processed to concentrate the minerals of interest. In the case of metal ores, these

mineral concentrates usually need to be further processed to separate the metal from other elements in the

ore minerals. Smelting is the process of separating the metal from impurities by heating the concentrate to a

high temperature to cause the metal to melt. Smelting the concentrate produces a metal or a high-grade

metallic mixture along with a solid waste product called slag.

The principal sources of pollution caused by smelting are contaminant-laden air emissions and

process wastes such as wastewater and slag.

38
One type of pollution attributed to air emissions is acid rain. The smelting of sulfide ores results in

the emission of sulfur dioxide gas, which reacts chemically in the atmosphere to form a sulfuric acid mist. As

this acid rain falls to the earth, it increases the acidity of soils, streams, and lakes, harming the health of

vegetation and fish and wildlife populations.

In older smelters, air emissions contained elevated levels of various metals. Copper and selenium,

for example, which can be released from copper smelters, are essential to organisms as trace elements, but

they are toxic if they are overabundant. These metals can contaminate the soil in the vicinity of smelters,

destroying much of the vegetation. In addition, particulate matter emitted from smelters may include oxides

of such toxic metals as arsenic (cumulative poison), cadmium (heart disease), and mercury (nerve damage).

39
Item No. 11: Blast furnace

Blast furnace, a vertical shaft furnace that produces liquid metals by the reaction of a flow

of air introduced under pressure into the bottom of the furnace with a mixture of metallic ore, coke,

and flux fed into the top. Blast furnaces are used to produce pig iron from iron ore for subsequent processing

into steel, and they are also employed in processing lead, copper, and other metals. Rapid combustion is

maintained by the current of air under pressure.

Blast furnaces produce pig iron from iron ore by the reducing action of carbon (supplied as coke) at

a high temperature in the presence of a fluxing agent such as limestone. Ironmaking blast furnaces consist

of several zones: a crucible-shaped hearth at the bottom of the furnace; an intermediate zone called a bosh

between the hearth and the stack; a vertical shaft (the stack) that extends from the bosh to the top of the

furnace; and the furnace top, which contains a mechanism for charging the furnace. The furnace charge, or

burden, of iron-bearing materials (e.g., iron-ore pellets and sinter), coke, and flux (e.g., limestone) descends

through the shaft, where it is preheated and reacts with ascending reducing gases to produce liquid iron and

slag that accumulate in the hearth. Air that has been preheated to temperatures from 900° to 1,250° C

(1,650° and 2,300° F), together with injected fuel such as oil or natural gas, is blown into the furnace through

multiple tuyeres(nozzles) located around the circumference of the furnace near the top of the hearth; these

nozzles may number from 12 to as many as 40 on large furnaces. The preheated air is, in turn, supplied from

40
a bustle pipe, a large-diameter pipe encircling the furnace. The preheated air reacts vigorously with the

preheated coke, resulting in both the formation of the reducing gas (carbon monoxide) that rises through the

furnace, and a very high temperature of about 1,650° C (3,000° F) that produces the liquid iron and slag.

The bosh is the hottest part of the furnace because of its close proximity to the reaction between air

and coke. Molten iron accumulates in the hearth, which has a taphole to draw off the molten iron and, higher

up, a slag hole to remove the mixture of impurities and flux. The hearth and bosh are thick-walled structures

lined with carbon-

type refractory blocks, while

the stack is lined with high-

quality fireclay brick to

protect the furnace shell. To

keep these refractory

materials from burning out,

plates, staves, or sprays for

circulating cool water are

built into them.

The stack is kept full with alternating layers of coke, ore, and limestone admitted at the top during

continuous operation. Coke is ignited at the bottom and burned rapidly with the forced air from the tuyeres.

The iron oxides in the ore are chemically reduced to molten iron by carbon and carbon monoxide from the

coke. The slag formed consists of the limestone flux, ash from the coke, and substances formed by the

reaction of impurities in the ore with the flux; it floats in a molten state on the top of the molten iron. Hot gases

rise from the combustion zone, heating fresh material in the stack and then passing out through ducts near

the top of the furnace.

41
Blast furnaces may have the following ancillary facilities: a stock house where the furnace burden is

prepared prior to being elevated to the furnace top by skip cars or a belt conveyor system; a top-charging

system consisting of a vertical set of double bells (cones) or rotating chutes to prevent the release of furnace

gas during charging; stoves that utilize the furnace off-gases to preheat the air delivered to the tuyeres; and

a cast house, consisting of troughs that distribute liquid iron and slag to appropriate ladles for transfer to

steelmaking furnaces and slag-reclamation areas.

Modern blast

furnaces range

in size from 20 to

35 m (70 to 120

feet), have

hearth diameters

of 6 to 14 m (20

to 45 feet), and

can produce

from 1,000 to

almost 10,000

tons of pig iron

daily.

42
Item No. 12: Basic Oxygen Surface

Basic Oygen Furncace (BOF) is a pear shaped vessel where the pig iron from blast furnace, and

ferrous scrap, is refined into steel by injecting a jet high-purity oxygen through the hot metal. More

specifically, in a BOF:

the carbon content of pig iron, which is typically 4-5%, is reduced to varying levels below 1%

depending on the product specifications; unwanted impurities are removed; concentration of desired is

brought to product specifications.

Approximately 50 Nm3 of oxygen, which is produced in another plant, is required for every ton of

liquid steel. As the reactions taking place in the BOF are highly exothermic, the temperatures in the furnace

usually reach 1600-1650 oC. Scrap, or scprap substitutes, that meet purity requirements are often added to

control excessive temperature rises. However, the pig iron input stays at the levels of 65 to 90% for every

ton of steel produced. Impurities are dissolved by the added limestone and fomed into a slag. During the

BOF processes a gas with high CO content is formed. If no gas recovery is exercised, CO is converted to

CO2 by combustion at the mouth of the furnaces with open hood, and through flaring after gas cleaning in

furnaces with a closed hood. BOF shops are often followed by secondary metallurgy processes - in laddle

or in vacuum - to give the product its final characteristics. BOF operations can be a net producer of energy if

the sensible and chemical energy leaving the furnace is recovered and utilized.

43
Basic oxygen process (BOP), a steelmaking method in

which pure oxygen is blown into a bath of molten blast-furnace iron

and scrap. The oxygen initiates a series of intensively exothermic

(heat-releasing) reactions, including the oxidation of such impurities

as carbon, silicon, phosphorus, and manganese.

The advantages of using pure oxygen instead of air in refining

pig iron into steel were recognized as early as 1855 by Henry

Bessemer, but the process could not be brought to commercial

fruition until the 20th century, when large tonnages of cheap, high-purity oxygen became available.

Commercial advantages include high production rates, less labour, and steel with a low nitrogen content.

Development of the BOP was initiated in Switzerland by Robert Durrer in the late 1940s. After experimenting

with a 2.5-ton pilot unit, Durrer worked with engineers at the Voest company at Linz, Austria, who set up a

commercially operating 35-ton converter in 1952. A second unit began operation within a year at Donawitz,

also in Austria. Consequently, the BOP was first known as the LD (Linz-Donawitz) process. Within 40 years,

virtually all of the steel in Japan and more than half of the steel worldwide was produced by the BOP.

A typical top-blown basic oxygen furnace is a vertical cylindrical vessel with a closed bottom and an

open upper cone through which a water-cooled oxygen lance can be raised and lowered. The vessel is lined

with a refractory such as magnesite and is mounted on trunnions so that it can be tilted for charging and also

for tapping liquid steel. A charge typically consisting of 70–75 percent molten blast-furnace iron (containing

approximately 4 percent carbon), 25–30 percent scrap metal, and lime and other fluxes is fed into the furnace.

The lance is lowered into the vessel, and oxygen is injected into the bath at supersonic velocities with flow

rates that can exceed 800 cubic m (28,000 cubic feet) per minute. The duration of the oxygen “blow,” normally

close to 20 minutes, is varied to reduce the carbon in the steel to the required level. The steel is then tapped

into a ladle at temperatures close to 1,600° C (2,900° F), and appropriate ferroalloys and deoxidizers are

44
added to meet the required steel composition. “Heats” of steel, ranging in size from 30 to 360 tons, can be

produced in 30 to 45 minutes.

Another, though less common, oxygen steelmaking system is a bottom-blown process known as the

Q-BOP (quick-quiet BOP) in North America and the OBM (from the German, Oxygen bodenblasen

Maxhuette, or “oxygen bottom-blowing furnace”) in Europe. In this system, oxygen is injected with lime

through nozzles, or tuyeres, located in the bottom of the vessel. The tuyeres consist of two concentric tubes:

oxygen and lime are introduced through the inner tube, and a hydrocarbon such as natural gas is injected

through the outer annulus. The endothermic (heat-absorbing) decomposition of the hydrocarbon near the

molten bath cools the tuyeres and protects the adjacent refractory. Yet another variation, which has found

wide application in top-blown furnaces, is the injection of inert gases into the molten bath through permeable

blocks in the bottom of the vessel for the purpose of enhancing chemical reactions.

45
Item No. 13: Electric Arc Furnace

Electric Arc Furnaces (EAFs) are a central part of the production route that is an alternative to the

dominant BF-BOF route. EAFs are used to produce carbon steels and alloy steels primarily by recycling

ferrous scrap. In an EAF scrap and/or manufactured iron units – such as DRI, pig iron, iron carbide – is

melted and converted into high quality steel by using high-power electric arcs formed between a cathode and

one (for DC) or three (for AC) anodes (Worrell et al., 2010. p. 14). Scrap is by far the he most important

resource, accounting for about 80% of all electric arc furnace metal feedstock (IEA, 2007. p. 128).

The iron units are loaded in a basket together with limestone – for slag formation – and charged into

the furnace. The main task of most modern EAFs is to convert the solid raw materials to liquid crude steel as

fast as possible and then refine further in subsequent secondary steelmaking processes. Nevertheless, if

time is available, almost any metallurgical operation may be performed during flat bath operation period (after

melting), which is usually performed as a pre-treatment to the secondary steelmaking operations. Oxygen

and coal powder injection are common treatment operations. (Worrell, et al., 2010. p. 14; Steeluniversity.org)

Total theoratical energy required to melt the scrap and to superheat it to the typical tap temperatures

requires around 350 – 370 kWh/t-steel. This energy can be provided by the electric arc, from fossil fuel

injection or oxidation of the scrap feedstock. In practice, the energy use is highly dependent on product mix,

local material and energy costs and is unique to the specific furnace operation. Factors such as raw material

composition, power input rates and operating practices – such as post-combustion, scrap preheating – can

greatly influence the balance. Actual electricity use in EAFs are reported to range between 300 – 550 kWh/t,

and in 1999 reached an average level of 425 kWh/t. Based on 2005 Figures, IEA estimated that reducing the

average electricity consumption from 425 kWh to 350 kWh for every ton of steel produced in EAFs worldwide,

0.1 EJ of energy can be saved annually (IEA, 207. pp. 130-131).

46
An electric arc furnace (EAF) is a furnace that heats charged material by means of an electric arc.

Industrial arc furnaces range in size from small units of approximately one ton capacity (used

in foundries for producing cast ironproducts) up to about 400 ton units used for secondary steelmaking. Arc

furnaces used in research laboratories and by dentistsmay have a capacity of only a few dozen grams.

Industrial electric arc furnace temperatures can be up to 1,800 °C (3,272 °F), while laboratory units can

exceed 3,000 °C (5,432 °F).

Arc furnaces differ

from induction furnaces in that

the charge material is directly

exposed to an electric arc and

the current in the furnace

terminals passes through the

charged material.

47
Item No. 14: Iron-Carbon Phase Diagram

In refining steel from iron ore, the quantity of carbon used must be carefully controlled in order for

the steel to have the desired properties. The reason for the strong relationship between steel properties and

carbon content can be understood by examining the iron–carbon phase diagram.

Figure 3.4 presents a commonly accepted iron–carbon phase diagram. One of the unique features

of this diagram is that the abscissa extends only to 6.7%, rather than 100%. This is a matter of convention.

In an iron-rich material, each carbon atom bonds with three iron atoms to form iron carbide, also called

cementite. Iron carbide is 6.7% carbon by weight. Thus, on the phase diagram, a carbon weight of 6.7%

corresponds to 100% iron carbide. A complete iron–carbon phase diagram should extend to 100% carbon.

However, only the iron-rich portion, as shown on Figure 3.4, is of practical significance (Callister, 2006). In

fact, structural steels have a maximum carbon content of less than 0.3%, so only a very small portion of the

phase diagram is significant for civil engineers.

48
The left side of Figure 3.4 demonstrates that pure iron goes through two transformations as

temperature increases. Pure iron below 912°C has a BCC crystalline structure called ferrite. At 912°C, the

ferrite undergoes a polymorphic change to a FCC structure called austenite. At 1394°C, another polymorphic

change occurs, returning the iron to a BCC structure. At 1539°C, the iron melts into a liquid. The high and

low temperature ferrites are identified as and ferrite, respectively. Since ferrite occurs only at very high

temperatures, it does not have practical significance for this book.

Carbon goes into solution with ferrite at temperatures between 400°C and 912°C. However, the

solubility limit is very low, with a maximum of 0.022% at 727°C. At temperatures below 727°C and to the

right of the solubility limit line, ferrite and iron carbide coexist as two phases. From 727°C to 1148°C, the

solubility of carbon in the austenite increases from 0.77% to 2.11%. The solubility of carbon in austenite is

greater than in a ferrite because of the crystalline structure of the austenite.

At 0.77% carbon and 727°C, a eutectoid reaction occurs; that is, a solid phase change occurs when

either the temperature or carbon content changes. At 0.77% carbon, and above 727°C, the carbon is in

solution as an interstitial element, within the FCC structure of the austenite. A temperature drop to below

727°C, which happens slowly enough to allow the atoms to reach an equilibrium condition, results in a two-

phase material, a ferrite and iron carbide. The ferrite will have 0.022% carbon in solution, and the iron carbide

will have a carbon content of 6.7%. The ferrite and iron carbide will form as thin plates, a lamellae structure.

This eutectoid material is called pearlite.

At carbon contents less than the eutectoid composition, 0.77% carbon, hypoeutectoid alloys are

formed. Consider a carbon content of 0.25%. Above approximately 860°C, solid austenite exists with carbon

in solution. The austenite consists of grains of uniform material that were formed when the steel was cooled

from a liquid to a solid. Under equilibrium temperature drop from 860°C to 727°C, ferrite is formed and

accumulates at the grain boundaries of the austenite. This is a proeutectoid ferrite. At temperatures slightly

49
above 727°C, the ferrite will have 0.022% carbon in solution and austenite will have 0.77% carbon. When

the temperature drops below 727°C, the austenite will transform to pearlite. The resulting structure consists

of grains of pearlite surrounded by a skeleton of ferrite.

When the carbon content is greater than the eutectoid composition, 0.77% carbon, hypereutectoid

alloys are formed. Iron carbide forms at the grain boundaries of the austenite at temperatures above 727°C.

The resulting microstructure consists of grains of pearlite surrounded by a skeleton of iron carbide.

The lever rule for the analysis of phase diagrams can be used to determine the phases and

constituents of steel.

50
Item No. 15: Heat Treatment of Steel - Annealing

Properties of steel can be altered by applying a variety of heat treatments. For example, steel can

be hardened or softened by using heat treatment; the response of steel to heat treatment depends upon its

alloy composition. Common heat treatments employed for steel include annealing, normalizing, hardening,

and tempering. The basic process is to heat the steel to a specific temperature, hold the temperature for a

specified period of time, then cool the material at a specified rate. The temperatures used for each of the

treatment types are shown in Figure 3.7.

51
The objectives of annealing are to refine the grain, soften the steel, remove internal stresses, remove

gases, increase ductility and toughness, and change electrical and magnetic properties. Four types of

annealing can be performed, depending on the desired results of the heat treatment:

Full annealing requires heating the steel to about 50°C above the austenitic temperature line and

holding the temperature until all the steel transforms into either austenite or austenite–cementite, depending

on the carbon content. The steel is then cooled at a rate of about 20°C per hour in a furnace to a temperature

of about 680°C, followed by natural convection cooling to room temperature. Due to the slow cooling rate,

the grain structure is a coarse pearlite with ferrite or cementite, depending on the carbon content. The slow

cooling rate ensures uniform properties of the treated steel. The steel is soft and ductile.

Process annealing is used to treat work-hardened parts made with low carbon steel (i.e., less than

0.25 percent carbon). The material is heated to about 700°C and held long enough to allow recrystallization

of the ferrite phase. By keeping the temperature below 727°C, there is not a phase shift between ferrite and

austenite, as occurs during full annealing. Hence, the only change that occurs is refinement of the size,

shape, and distribution of the grain structure.

Stress relief annealing is used to reduce residual stresses in cast, welded, and cold-worked parts

and cold-formed parts. The material is heated to 600 to 650°C, held at temperature for about one hour, and

then slowly cooled in still air.

Spheroidization is an annealing process used to improve the ability of high carbon (i.e., more than

0.6 percent carbon) steel to be machined or cold worked. It also improves abrasion resistance. The cementite

is formed into globules (spheroids) dispersed throughout the ferrite matrix.

52
Item No. 16: Heat Treatment of Steel – Normalizing

Normalizing is similar to annealing, with a slight difference in the temperature and the rate of cooling.

Steel is normalized by heating to about 60°C (110°F) above the austenite line and then cooling under natural

convection. The material is then air cooled. Normalizing produces a uniform, fine-grained microstructure.

However, since the rate of cooling is faster than that used for full annealing, shapes with varying thicknesses

results in the normalized parts having less uniformity than could be achieved with annealing. Since structural

plate has a uniform thickness, normalizing is an effective process and results in high fracture toughness of

the material.

53
Item No. 17: Heat Treatment of Steel – Hardening

Steel is hardened by heating it to a temperature above the transformation range and holding it until

austenite is formed. The steel is then quenched (cooled rapidly) by plunging it into, or spraying it with, water,

brine, or oil. The rapid cooling “locks” the iron into a BCC structure, marten site, rather than allowing the

transformation to the ferrite FCC structure. Marten site has a very hard and brittle structure. Since the cooling

occurs more rapidly at the surface of the material being hardened, the surface of the material is harder and

more brittle than the interior of the element, creating nonhomogeneous characteristics. Due to the rapid

cooling, hardening puts the steel in a state of strain. This strain sometimes causes steel pieces with sharp

angles or grooves to crack immediately after hardening. Thus, hardening must be followed by tempering.

54
Item No. 18: Heat Treatment of Steel – Tempering

The predominance of martensite in quench-hardened steel results in an undesirable brittleness.

Tempering is performed to improve ductility and toughness. Martensite is a somewhat unstable structure.

Heating causes carbon atoms to diffuse from martensite to produce a carbide precipitate and formation of

ferrite and cementite. After quenching, the steel is cooled to about 40°C then reheated by immersion in either

oil or nitrate salts. The steel is maintained at the elevated temperature for about two hours and then cooled

in still air.

In addition, Tempering is a process of heat treating, which is used to increase the toughness of iron-

based alloys. Tempering is usually performed after hardening, to reduce some of the excess hardness, and

is done by heating the metal to some temperature below the critical point for a certain period of time, then

allowing it to cool in still air. The exact temperature determines the amount of hardness removed, and

depends on both the specific composition of the alloy and on the desired properties in the finished product.

55
For instance, very hard tools are often tempered at low temperatures, while springs are tempered at much

higher temperatures.

Item No. 19: Steel Alloys

Alloy metals can be used to alter the characteristics of steel. By some counts, there are as many as

250,000 different alloys of steel produced. Of these, as many as 200 may be used for civil engineering

applications. Rather than go into the specific characteristics of selected alloys, the general effect of different

alloying agents will be presented. Alloy agents are added to improve one or more of the following properties:

1. hardenability

2. corrosion resistance

3. machinability

4. ductility

5. strength

Common alloy agents, their typical percentage range, and their effects are summarized in Table 3.1.

56
By altering the carbon and alloy content and by using different heat treatments, steel can be produced

with a wide variety of characteristics. These are classified as follows:

1. Low alloy

■ Low carbon Plain High strength–low alloy

■ Medium carbon Plain Heat treatable

57
■ High carbon Plain Tool

2. High Alloy

■ Tool

■ Stainless

Steels used for construction projects are predominantly low- and medium-carbon plain steels.

Stainless steel has been used in some highly corrosive applications, such as dowel bars in concrete

pavements and steel components in swimming pools and drainage lines. The Specialty Steel Industry of

North America, SSINA, promotes the use of stainless steel for structural members where corrosion resistance

is an important design consideration (SSINA, 1999).

The use and control of alloying agents is one of the most significant factors in the development of

steels with better performance characteristics. The earliest specification for steel used in building and bridge

construction, published in 1900, did not contain any chemical requirements. In 1991 ASTM published the

specification which controls content of 10 alloying elements in addition to carbon (Hassett, 2003).

58
Item No. 20: Structural Steel Grades

Structural steel is used in hot-rolled structural shapes, plates, and bars. Structural steel is used for

various types of structural members, such as columns, beams, bracings, frames, trusses, bridge girders, and

other structural applications.

Due to the widespread use of steel in many applications, there are a wide variety of systems for

identifying or designating steel, based on grade, type and class. Virtually every country with an industrial

capacity has specifications for steel. In the United States, there are several associations that write

specifications for steel, such as the Society of Automotive Engineers, SAE, the American Iron and Steel

Institute, AISI, and the American Society for Testing and Materials, ASTM. The most widely used designation

59
system was developed cooperatively by SAE and AISI based on chemical composition. This system uses a

letter to identify the broad class of alloys, and a five-digit number to define specific alloys within the class.

Several grades of structural steel are produced in the United States. Table 3.2 is a summary of

selected information from various sources. The American Institute of Steel Construction, AISC, Manual for

Steel Construction is an excellent reference on the types of steel used for structural applications. However,

the best sources of information for structural steels are the various ASTM specifications. Of particular note is

the fact that additional requirements are frequently included, dependent on the geometry of the product made

with a particular steel.

60
Historically, dating back to 1900, only two types of structural steel were used in the United States:

A7 for bridges and A9 for buildings. The specifications for these materials were very similar and in 1938, they

61
were combined into a single specification, A7. The specification for A7 and A9 were limited to requirements

for the tensile strength and yield point only; there were no chemical specifications. The chemical composition,

particularly carbon content, became an issue during the 1950s, as welding gained favor for making structural

connections. By 1964, AISC adopted five grades of steel for structural applications. The 1999 AISC Load

and Resistance Factor Design Specification for Structural Steel Buildings, 1999, identifies 15 different ASTM

steel designations for structural applications.

Item No. 21: Steel Sectional Shapes

Figure 3.9 illustrates structural cross-sectional shapes commonly used in structural applications.

These shapes are produced in different sizes and are designated with the letters W, HP, M, S, C, MC, and

L. W shapes are doubly symmetric wide-flange shapes whose flanges are substantially parallel. HP shapes

are also wide-flange shapes whose flanges and webs are of the same nominal thickness and whose depth

and width are essentially the same. The S shapes are doubly symmetric shapes whose inside flange surfaces

have approximately 16.67% slope. The M shapes are doubly symmetric shapes that cannot be classified as

W, S, or HP shapes. C shapes are channels with inside flange surfaces having a slope of approximately

16.67%. MC shapes are channels that cannot be classified as C shapes. L shapes are angle shapes with

either equal or unequal legs. In addition to these shapes, other structural sections are available, such as tee,

sheet piling, and rail, as shown in Figure 3.9.

The W, M, S, HP, C, and MC shapes are designated by a letter, followed by two numbers separated

by an The letter indicates the shape, while the two numbers indicate the nominal depth and the weight per

linear unit length. For example, W means W shape with a nominal depth of 44 in. and a weight of 335 lb/linear

62
foot. An angle is designated with the letter L, followed by three numbers that indicate the leg dimensions and

thickness in inches, such as L Dimensions of these structural shapes are controlled by ASTM A6/A6M.

W shapes are commonly used as beams and columns, HP shapes are used as bearing piles, and S

shapes are used as beams or girders. Composite sections can also be formed by welding different shapes

to use in various structural applications. Sheet piling sections are connected to each other and are used as

retaining walls.

63
Item No. 22: Cold-Formed Steel

Cold-formed steel is used for structural framing of floors, walls and roofs as well as interior partitions

and exterior curtain wall applications. The thickness of cold formed steel framing members ranges from 18

mils (0.0179 inches, 0.455 mm) to 118 mils (0.1180 inches, 3.000 mm). Cold-formed steel was formerly

known as “light gauge” steel; however, the reference nomenclature “gauge” became obsolete with the

adoption of a Universal Designator System for all generic cold-formed steel framing members in 2000.

Cold-formed steel used for steel framing members is predominately manufactured from scrap steel

using either electric arc or basic oxygen furnaces to cast slabs. The slabs are passed through a machine with

a series of rollers that reduce the slab into thin sheets of the desired thicknesses, strengths and other physical

properties. The sheets are sent thought a hot-dipped galvanizing process and then rolled into coils that weigh

approximately 13 tons.

The primary method of manufacturing steel framing members is roll-forming. At the roll-former the

coils are slit into the required width and fed through a series of dies, to form the stud, joist, angle or other

cold-formed member, as shown in Figure 3.10. Steel framing members may be manufactured with holes in

the member webs to facilitate utility runs during construction. The fabrication of all cold-formed steel

construction materials is governed by industry standards including the American Iron and Steel Institute’s

Specification for the Design of Cold-Formed Steel Framing Members (NASPEC) and ASTM.

64
Item No. 23: Cold-Formed Steel Grades

Structural and non-structural cold-formed steel members are manufactured from sheet steel in

compliance with ASTM A1003/A1003M, but limited to the material types and grades listed in Table 3.7. While

multiple grades are acceptable for the different steel types, the North American Standard for Cold-Formed

Steel Framing recognizes two yield strengths 33 and 55 ksi (AISI S201-07).

The large deformations caused by the cold-forming process results in local strainhardening at the

corners. The plastic deformation of the steel results instrain-hardening at the bends that increases the yield

strength, tensile (ultimate) strength and hardness but reduces ductility. Strain hardening can almost double

the yield strength and increase the tensile strength by 40% (Karren and Winter, 1967).

65
Item No. 24: Cold-Formed Steel Shapes

A wide variety of shapes can be produced by cold-forming and manufacturers have developed a

wide range of products to meet specific applications. Figure 3.11 shows the common shapes of typical cold-

formed steel framing members. Figure 3.12 shows common shapes for profiled sheets and trays used for

roofing and wall cladding and for load bearing deck panels.

For common applications, such as structural studs, industry organizations, such as the Steel Framing

Alliance (SFA) and the Steel Stud Manufacturers Association (SSMA) have developed standard shapes and

nomenclature to promote uniformity of product availability across the industry. Figure 3.11 shows the generic

shapes covered by the Universal Designator System. The designator consists of four sequential codes. The

first code is a three or four-digit number indicating the member web depth in 1/100 inches. The second is a

single letter indicating the type of member, as follows:

S = Stud or joist framing member with stiffening lips

L = Angle or L-header

F = Furring channels

U = Cold-rolled channel

T = Track section

66
67
The third is a three-digit numeral indication flange width in 1/100 inches followed by a dash. The

fourth is a two or three-digit numeral indicating the base steel thickness in 1/1000 inch (mils).

As an example, the designator system for a 6'', C-shape with 1-5/8''(1.62'') flanges and made with

0.054''thick steel is 600S162-54.

Item No. 25: Rebar

Rebar (short for reinforcing bar), known when massed as reinforcing steel or reinforcement steel, is

a steel bar or mesh of steel wires used as a tension device in reinforced concrete and

reinforced masonry structures to strengthen and aid the concrete under tension. Concrete is strong under

compression, but has weak tensile strength. Rebar significantly increases the tensile strength of the structure.

Rebar's surface is often deformed to promote a better bond with the concrete.

A tied rebar beam cage. This will be embedded inside

cast concrete to increase its tensile strength.

The most common type of rebar is carbon steel, typically consisting of hot-rolled round bars with

deformation patterns. Other readily available types include stainless steel, and composite bars made of glass

68
fiber, carbon fiber, or basalt fiber. The steel reinforcing bars may also be coated in an epoxy resin designed

to resist the effects of corrosion mostly in saltwater environments, but also land based constructions. Bamboo

has been shown to be a viable alternative to reinforcing steel in concrete construction.[citation needed] These

alternate types tend to be more expensive or may have lesser mechanical properties and are thus more often

used in specialty construction where their physical characteristics fulfill a specific performance requirement

that carbon steel does not provide. Steel and concrete have similar coefficients of thermal expansion, so a

concrete structural member reinforced with steel will experience minimal differential stress as the

temperature changes.

Steel has a thermal expansion coefficient nearly equal to that of modern concrete. If this were not

so, it would cause problems through additional longitudinal and perpendicular stresses at temperatures

different from the temperature of the setting. Although rebar has ribs that bind it mechanically to the concrete,

it can still be pulled out of the concrete under high stresses, an occurrence that often accompanies a larger-

scale collapse of the structure. To prevent such a failure, rebar is either deeply embedded into adjacent

structural members (40–60 times the diameter), or bent and hooked at the ends to lock it around the concrete

and other rebar. This first approach increases the friction locking the bar into place, while the second makes

use of the high compressive strength of concrete.

Common rebar is made of unfinished tempered steel, making it susceptible to rusting. Normally the

concrete cover is able to provide a pH value higher than 12 avoiding the corrosion reaction. Too little concrete

cover can compromise this guard through carbonation from the surface, and salt penetration. Too much

concrete cover can cause bigger crack widths which also compromises the local guard. As rust takes up

greater volume than the steel from which it was formed, it causes severe internal pressure on the surrounding

concrete, leading to cracking, spalling, and, ultimately, structural failure. This phenomenon is known as oxide

jacking. This is a particular problem where the concrete is exposed to salt water, as in bridges where salt is

69
applied to roadways in winter, or in marine applications. Uncoated, corrosion-resistant

low carbon/chromium (microcomposite), silicon bronze, epoxy-coated, galvanized, or stainless steel rebars

may be employed in these situations at greater initial expense, but significantly lower expense over the

service life of the project. Extra care is taken during the transport, fabrication, handling, installation, and

concrete placement process when working with epoxy-coated rebar, because damage will reduce the long-

term corrosion resistance of these bars. Even damaged bars have shown better performance than uncoated

reinforcing bars, though issues from debonding of the epoxy coating from the bars and corrosion under the

epoxy film have been reported. These bars are used in over 70,000 bridge decks in the USA.

The requirements for deformations are found in standard product specifications for steel bar

reinforcing, such as ASTM A615 and ASTM A706, and dictate lug spacing and height.

Fibre-reinforced plastic rebar is also used in high-corrosion environments. It is available in many forms, such

as spirals for reinforcing columns, common rods, and meshes. Most commercially available rebar is made

from unidirectional fibers set in a thermoset polymer resin, and is often referred to as FRP.

Some special construction such as research and manufacturing facilities with very sensitive electronics may

require the use of reinforcement that is non-conductive to electricity, and medical imaging equipment rooms

may require non-magnetic properties to avoid interference. FRP rebar, notably glass fibre types have low

electrical conductivity and are non-magnetic which is commonly used for such needs. Stainless steel rebar

with low magnetic permeability is available and is sometimes used to avoid magnetic interference issues.

Reinforcing steel can also be displaced by impacts such as earthquakes, resulting in structural

failure. The prime example of this is the collapse of the Cypress Street Viaduct in Oakland, California as a

result of the 1989 Loma Prieta earthquake, causing 42 fatalities. The shaking of the earthquake caused

70
rebars to burst from the concrete and buckle. Updated building designs, including more circumferential rebar,

can address this type of failure.

There are many methods of preparing metals and alloys for use; in this chapter we only have room for briefly

describing a few of the more important ones that are used in the production of both metals and alloys. Before

starting, we must recognise that metallurgists look on these not only as ways of shaping materials but also

as ways of controlling their microstructure and, consequently, their properties.

Item No. 26: Forming of Metals – Castings

Most common metals can be produced by melting and casting into moulds. The cast may be of the

shape and dimensions required for the component, or a prism of material may be produced for further

processing.

The general processes taking place during the solidification of molten pure metals and metallic

solutions have been described in Chapter 1. Solidification of alloys often gives rise to compositional variations

from place to place in a casting and on a microstructural scale within the dendrites. When intended for further

processing, little attempt is made to control grain size, and the metal often solidifies to a rather coarse grain

structure containing a number of casting defects, such as porosity, compositional variations and shrinkage.

These are not disastrous because further processing will rectify them. Shaped castings need more care. To

ensure that the desired mechanical properties are achieved the castings are normally degassed, the grain

size is carefully controlled by one or more of the means and compositional variations minimised by attention

to solidification patterns within the mould.

71
Item No. 27: Forming of Metals – Hot Working

The working of metals and alloys by rolling, forging, extrusion etc. (Fig. 9.2) depends upon plasticity,

which is usually much greater at high temperatures, i.e. temperatures above the metals’ recrystallisation

temperature. This allows all the common metals to be heavily deformed, especially in compression, without

breaking. For structural steel members, the most usual method is by hot rolling between simple cylindrical or

shaped rolls at temperatures around 1000°C or higher. After rolling, the members are left to cool naturally

72
and end up with annealed microstructures and grain sizes that depend on the extent of the deformation, the

maximum temperature used and the cooling rate. It follows that all these process variables need to be

controlled to give products with consistent properties. Another feature of hotworking processes is that

exposure to air at high temperatures causes a heavy film of oxide to form on the surface. Thus steel sections

delivered ‘as rolled’ are covered with iron oxide (mill scale) and need to be shot-blasted or sand-blasted

before receiving any protective coating. For many structural steels further heat treatment is required, as we

shall see in Chapter 11.

Many familiar articles, e.g. engine crankshafts, are forged into shape. This involves placing a hot

blank into one half of a shaped mould and then impressing the other half of the mould onto the blank (Fig.

9.2). This can be done under impact using such methods as drop forging and die stamping or more slowly

using large hydraulic presses.

Many metals can be extruded. This has the advantage that very long lengths with complex sections

can be produced. Aluminium glazing bars are a familiar example. One disadvantage of hot forming arises

from the contraction of the article on cooling and from such problems as oxidation. These and other factors

conspire to limit the precision of the product. In some cases the tolerances are acceptable, but to meet more

demanding tolerances further cold forming or machining is required.

73
74
Item No. 28: Forming of Metals – Cold Working

Because of their ductility at room temperature many metals and alloys can be cold worked, that is to

say, shaped at temperatures below their recrystallisation temperature. As we discussed in Chapter 8, this

creates an immense number of dislocations and, as a consequence, the metal work-hardens and its yield

point is raised. Indeed, for pure metals and some alloys it is the only way of increasing the yield strength.

There are many cold-working processes. Rolling is extensively used to produce sheet material, while

high-strength wire, as used for pre-stressing strands and cables, is cold drawn by pulling through a tapered

die. Metal sheets can be shaped into cups, bowls or motor-car body panels by deep drawing or stretch

forming (Fig. 9.2). Some metals can be cold extruded.

Clearly there comes a limit beyond which the ductility is exhausted and the metal will fracture. If

further cold work is required the metal must be annealed by heating it to a temperature where recrystallisation

occurs, when the original ductility is restored and further working is possible.

Cold working using well designed tools and careful control is capable of delivering to demanding

tolerances. From the metallurgist’s point of view, control over rolling and annealing schedules is a very

effective way of controlling the grain size of the product.

75
Item No. 29: Joining – Welding

The design and fabrication of joints between metallic structural components are obviously crucial

factors in ensuring the success of the structure. Design engineers have Codes of Practice etc. to help them

in their task, but some understating of the processes involved and relevant materials’ behaviour is also

important. Although adhesive bonding is being used increasingly for joining metal parts, the commonest

methods are still welding, brazing, soldering or by mechanical fasteners, such as rivets and bolts.

All welding involves essentially the same sequence of operations at the joint. The material is heated

locally to its melting temperature, additional metal may or may not be added and the joint is then allowed to

cool naturally. Some protection to the weld to avoid oxidation of the metal when molten and during cooling is

often provided by a slag layer (which is knocked-off when the weld has cooled) or by working in an

atmosphere of an inert gas such as argon. Whatever the material or process all welds should comply with

the two following ideal requirements:

76
1. There should be complete continuity between the parts to be joined, and every part of the joint

should be indistinguishable from the parent metal. In practice this is not always achieved,

although welds giving satisfactory performance can be made.

2. Any additional joining metal should have metallurgical properties that are no worse than those

of the parent metal. This is largely the concern of the supplier of welding consumables, though

poor welding practice can significantly affect the final product.

The weld itself is a small and rapidly formed casting. However, during welding a temperature gradient

is created in the parent material which results in a heat affected zone (HAZ) surrounding the weld. This

gradient ranges from the melting temperature at the fusion zone to ambient temperature at some distant

point. In the regions that have been exposed to high temperature and fast cooling rates, metallurgical

changes can occur. The quality of the joint is therefore affected by both the structure and properties of the

weld metal and the structure and properties of that part of the parent material that has undergone a significant

thermal cycle (the HAZ).

Both of these are significantly affected by the rate of cooling after welding – the slower the rate of

cooling the closer the structure is to equilibrium. Cooling occurs principally by conduction in the parent metal

and, since the thermal conductivity is a constant, the controlling factor is the thermal mass, i.e. the thickness

and size of the material to be welded. The greater the thermal mass the faster the cooling rate. Responses

to rapid cooling differ markedly from metal to metal, not only from say aluminium to steel but also from one

steel to another. Structural steels are designed to be weldable, i.e. they are capable of being welded without

serious loss of performance in the weld and HAZ. Nevertheless, the job must be carried out with thought,

care and skill, with due allowance made if the welding is being carried out in difficult conditions such as on a

construction site in poor weather. Most jobs are best carried out by welding specialists.

77
Item No. 30: Joining – Brazing, Soldering and Gluing

Brazing and soldering, and in some cases gluing, involve joining by means of a thin film of a material

that has a melting temperature lower than that of the parent material and which, when melted, flows into the

joint, often by capillary action, to form a thin film which subsequently solidifies. A sound well-brazed or

soldered joint should have a strength that is not too different from that of the parent material. Quite high forces

are needed to break a film of liquid provided the film is thin enough and the same applies to thin solid films.

This is not quite the whole explanation but is a very significant part of it. The rest is associated with the

behaviour of materials under complex stress conditions, biaxial and triaxial, and is beyond the scope of this

chapter.

78
Although it may seem strange to say so, gluing works in a very similar way. Thin layers of modern

adhesives bond well to the substrate material and are strong in shear. Design of joints to be made by gluing,

soldering or brazing should avoid potential failure by peeling and aim to use the adhesive in shear.

Soldering

Brazing

Gluing

79
Item No. 20: Joining – Bolting and Riveting

Some materials (such as cast iron) do not lend themselves to joining by welding. Even with materials

that can be welded (such as structural steel) it may not possible to weld prefabricated elements on a

construction site owing to difficulties of access and working conditions for both the welder and the welding

equipment. Gluing and brazing may be valid options but require thought about the joint design. Bolting and

riveting are by far the most common ways of making joints in such circumstances. Both rely on friction. A

tightened bolt forces the two members together and the friction between nut and bolt at the threads holds it

in place. In riveting, the hot rivet is hammered into prepared holes and the end hammered flat as a ‘head’ on

the surface of the sheet; as it cools it contracts and develops a tensile stress that effectively locks the

members together. High-strength friction grip bolts used in structural steelwork combine both aspects; the

nut is tightened to place the bolt into tension and this tensile pre-stress acts in the same way as the tensile

stress in a rivet.

Having produced and formed a metal, it is necessary to ensure that it performs well during service.

A major consideration in this is corrosion. This involves loss of material from the metal’s surface and can be

divided into two processes: dry oxidation and wet corrosion.

Item No. 31: Iron Ore

Earth's most important iron ore deposits are found in sedimentary rocks. They formed from chemical

reactions that combined iron and oxygen in marine and fresh waters. The two most important minerals in

these deposits are iron oxides: hematite (Fe2O3) and magnetite (Fe3O4). These iron ores have been mined

to produce almost every iron and steel object that we use today - from paper clips to automobiles to the steel

beams in skyscrapers.

80
Iron ores are rocks and minerals from which metallic iron can be economically extracted.

The ores are usually rich in iron oxides and vary in colour from dark grey, bright yellow, or deep purple to

rusty red. The iron is usually found in the form of magnetite (Fe

3O4,72.4%Fe), hematite (Fe2O3, 69.9% Fe), goethite (FeO(OH), 62.9% Fe), limonite (FeO(OH)·n(H2O),

55% Fe) or siderite (FeCO3, 48.2% Fe).

Ores containing very high quantities of hematite or magnetite (greater than about 60% iron) are

known as "natural ore" or "direct shipping ore", meaning they can be fed directly into iron-making blast

furnaces. Iron ore is the raw material used to make pig iron, which is one of the main raw materials to

make steel—98% of the mined iron ore is used to make steel. Indeed, it has been argued that iron ore is

"more integral to the global economy than any other commodity, except perhaps oil".

Metallic iron is virtually unknown on the surface of the Earth except as iron-

nickel alloys from meteorites and very rare forms of deep mantle xenoliths. Iron meteorites themselves are

thought to have originated from stellar bodies larger than 1,000 km in diameter. The origin of iron can be

ultimately traced to formation through nuclear fusion in stars and most of the iron is thought to have originated

in dying stars that are large enough to collapse or explode as supernovae. Although iron is the fourth-most

abundant element in the Earth's crust, comprising about 5%, the vast majority is bound in silicate or more

rarely carbonate minerals (for more information, see iron cycle). The thermodynamic barriers to separating

pure iron from these minerals are formidable and energy intensive, therefore all sources of iron used by

human industry exploit comparatively rarer iron oxide minerals, primarily hematite.

Prior to the industrial revolution, most iron was obtained from widely available goethite or bog ore,

for example during the American Revolution and the Napoleonic Wars. Prehistoric societies used laterite as

a source of iron ore. Historically, much of the iron ore utilized by industrialized societies has been mined from

predominantly hematite deposits with grades of around 70% Fe. These deposits are commonly referred to

81
as "direct shipping ores" or "natural ores". Increasing iron ore demand, coupled with the depletion of high-

grade hematite ores in the United States, after World War II led to development of lower-grade iron ore

sources, principally the utilization of magnetite and taconite.

Iron-ore mining methods vary by the type of ore being mined. There are four main types of iron-ore

deposits worked currently, depending on the mineralogy and geology of the ore deposits. These are

magnetite, titanomagnetite, massive hematite and pisolitic ironstone deposits.

Furthermore, the primary use of iron ore is in the production of iron. Most of the iron produced is then

used to make steel. Steel is used to make automobiles, locomotives, ships, beams used in buildings,

furniture, paper clips, tools, reinforcing rods for concrete, bicycles, and thousands of other items. It is the

most-used metal by both tonnage and purpose.

Iron Ore: A specimen of oolitic hematite iron ore.

The specimen shown is about two inches (five centimeters) across.

82

You might also like