Steel

Last updated

Steel is an alloy of iron and carbon with improved strength and fracture resistance compared to other forms of iron. Because of its high tensile strength and low cost, steel is one of the most commonly manufactured materials in the world. Steel is used in buildings, as concrete reinforcing rods, in bridges, infrastructure, tools, ships, trains, cars, bicycles, machines, electrical appliances, furniture, and weapons.

Contents

Iron is always the main element in steel, but many other elements may be present or added. Stainless steels, which are resistant to corrosion and oxidation, typically need an additional 11% chromium.

Iron is the base metal of steel. Depending on the temperature, it can take two crystalline forms (allotropic forms): body-centred cubic and face-centred cubic. The interaction of the allotropes of iron with the alloying elements, primarily carbon, gives steel and cast iron their range of unique properties. In pure iron, the crystal structure has relatively little resistance to the iron atoms slipping past one another, and so pure iron is quite ductile, or soft and easily formed. In steel, small amounts of carbon, other elements, and inclusions within the iron act as hardening agents that prevent the movement of dislocations.

The carbon in typical steel alloys may contribute up to 2.14% of its weight. Varying the amount of carbon and many other alloying elements, as well as controlling their chemical and physical makeup in the final steel (either as solute elements, or as precipitated phases), impedes the movement of the dislocations that make pure iron ductile, and thus controls and enhances its qualities. These qualities include the hardness, quenching behaviour, need for annealing, tempering behaviour, yield strength, and tensile strength of the resulting steel. The increase in steel's strength compared to pure iron is possible only by reducing iron's ductility.

Steel was produced in bloomery furnaces for thousands of years, but its large-scale, industrial use began only after more efficient production methods were devised in the 17th century, with the introduction of the blast furnace and production of crucible steel. This was followed by the Bessemer process in England in the mid-19th century, and then by the open-hearth furnace. With the invention of the Bessemer process, a new era of mass-produced steel began. Mild steel replaced wrought iron. The German states were the major steel producers in Europe in the 19th century. [1] American steel production was centered in Pittsburgh, Bethlehem, Pennsylvania, and Cleveland until the late 20th century.

Further refinements in the process, such as basic oxygen steelmaking (BOS), largely replaced earlier methods by further lowering the cost of production and increasing the quality of the final product. Today more than 1.6 billion tons of steel is produced annually. Modern steel is generally identified by various grades defined by assorted standards organisations. The modern steel industry is one of the largest manufacturing industries in the world, but also one of the most energy and greenhouse gas emission intense industries, contributing 8% of global emissions. [2] However, steel is also very reusable: it is one of the world's most-recycled materials, with a recycling rate of over 60% globally. [3]

The steel cable of a colliery winding tower Steel wire rope.JPG
The steel cable of a colliery winding tower

The noun steel originates from the Proto-Germanic adjective stahliją or stakhlijan 'made of steel', which is related to stahlaz or stahliją 'standing firm'. [4]

The carbon content of steel is between 0.02% and 2.14% by weight for plain carbon steel (iron-carbon alloys). Too little carbon content leaves (pure) iron quite soft, ductile, and weak. Carbon contents higher than those of steel make a brittle alloy commonly called pig iron. Alloy steel is steel to which other alloying elements have been intentionally added to modify the characteristics of steel. Common alloying elements include: manganese, nickel, chromium, molybdenum, boron, titanium, vanadium, tungsten, cobalt, and niobium. [5] Additional elements, most frequently considered undesirable, are also important in steel: phosphorus, sulfur, silicon, and traces of oxygen, nitrogen, and copper.

Plain carbon-iron alloys with a higher than 2.1% carbon content are known as cast iron. With modern steelmaking techniques such as powder metal forming, it is possible to make very high-carbon (and other alloy material) steels, but such are not common. Cast iron is not malleable even when hot, but it can be formed by casting as it has a lower melting point than steel and good castability properties. [5] Certain compositions of cast iron, while retaining the economies of melting and casting, can be heat treated after casting to make malleable iron or ductile iron objects. Steel is distinguishable from wrought iron (now largely obsolete), which may contain a small amount of carbon but large amounts of slag.

Material properties

Origins and production

An iron-carbon phase diagram showing the conditions necessary to form different phases FeC-phase-diagram--multilingual.svg
An iron-carbon phase diagram showing the conditions necessary to form different phases
An incandescent steel workpiece in a blacksmith's art Blacksmithing at the 2015 Fort Ross Festival - Fort Ross State Historic Park - Jenner, California - Sarah Stierch.jpg
An incandescent steel workpiece in a blacksmith's art

Iron is commonly found in the Earth's crust in the form of an ore, usually an iron oxide, such as magnetite or hematite. Iron is extracted from iron ore by removing the oxygen through its combination with a preferred chemical partner such as carbon which is then lost to the atmosphere as carbon dioxide. This process, known as smelting, was first applied to metals with lower melting points, such as tin, which melts at about 250 °C (482 °F), and copper, which melts at about 1,100 °C (2,010 °F), and the combination, bronze, which has a melting point lower than 1,083 °C (1,981 °F). In comparison, cast iron melts at about 1,375 °C (2,507 °F). [6] Small quantities of iron were smelted in ancient times, in the solid-state, by heating the ore in a charcoal fire and then welding the clumps together with a hammer and in the process squeezing out the impurities. With care, the carbon content could be controlled by moving it around in the fire. Unlike copper and tin, liquid or solid iron dissolves carbon quite readily.[ citation needed ]

All of these temperatures could be reached with ancient methods used since the Bronze Age. Since the oxidation rate of iron increases rapidly beyond 800 °C (1,470 °F), it is important that smelting take place in a low-oxygen environment. Smelting, using carbon to reduce iron oxides, results in an alloy (pig iron) that retains too much carbon to be called steel. [6] The excess carbon and other impurities are removed in a subsequent step.[ citation needed ]

Other materials are often added to the iron/carbon mixture to produce steel with the desired properties. Nickel and manganese in steel add to its tensile strength and make the austenite form of the iron-carbon solution more stable, chromium increases hardness and melting temperature, and vanadium also increases hardness while making it less prone to metal fatigue. [7]

To inhibit corrosion, at least 11% chromium can be added to steel so that a hard oxide forms on the metal surface; this is known as stainless steel. Tungsten slows the formation of cementite, keeping carbon in the iron matrix and allowing martensite to preferentially form at slower quench rates, resulting in high-speed steel. The addition of lead and sulfur decrease grain size, thereby making the steel easier to turn, but also more brittle and prone to corrosion. Such alloys are nevertheless frequently used for components such as nuts, bolts, and washers in applications where toughness and corrosion resistance are not paramount. For the most part, however, p-block elements such as sulfur, nitrogen, phosphorus, and lead are considered contaminants that make steel more brittle and are therefore removed from steel during the melting processing. [7]

Properties

Fe-C phase diagram for carbon steels, showing the A0, A1, A2 and A3 critical temperatures for heat treatments Steel Fe-C phase diagram-en.png
Fe-C phase diagram for carbon steels, showing the A0, A1, A2 and A3 critical temperatures for heat treatments

The density of steel varies based on the alloying constituents but usually ranges between 7,750 and 8,050 kg/m3 (484 and 503 lb/cu ft), or 7.75 and 8.05 g/cm3 (4.48 and 4.65 oz/cu in). [8]

Even in a narrow range of concentrations of mixtures of carbon and iron that make steel, several different metallurgical structures, with very different properties can form. Understanding such properties is essential to making quality steel. At room temperature, the most stable form of pure iron is the body-centred cubic (BCC) structure called alpha iron or α-iron. It is a fairly soft metal that can dissolve only a small concentration of carbon, no more than 0.005% at 0 °C (32 °F) and 0.021 wt% at 723 °C (1,333 °F). The inclusion of carbon in alpha iron is called ferrite. At 910 °C, pure iron transforms into a face-centred cubic (FCC) structure, called gamma iron or γ-iron. The inclusion of carbon in gamma iron is called austenite. The more open FCC structure of austenite can dissolve considerably more carbon, as much as 2.1%, [9] (38 times that of ferrite) carbon at 1,148 °C (2,098 °F), which reflects the upper carbon content of steel, beyond which is cast iron. [10] When carbon moves out of solution with iron, it forms a very hard, but brittle material called cementite (Fe3C).[ citation needed ]

When steels with exactly 0.8% carbon (known as a eutectoid steel), are cooled, the austenitic phase (FCC) of the mixture attempts to revert to the ferrite phase (BCC). The carbon no longer fits within the FCC austenite structure, resulting in an excess of carbon. One way for carbon to leave the austenite is for it to precipitate out of solution as cementite, leaving behind a surrounding phase of BCC iron called ferrite with a small percentage of carbon in solution. The two, ferrite and cementite, precipitate simultaneously producing a layered structure called pearlite, named for its resemblance to mother of pearl. In a hypereutectoid composition (greater than 0.8% carbon), the carbon will first precipitate out as large inclusions of cementite at the austenite grain boundaries until the percentage of carbon in the grains has decreased to the eutectoid composition (0.8% carbon), at which point the pearlite structure forms. For steels that have less than 0.8% carbon (hypoeutectoid), ferrite will first form within the grains until the remaining composition rises to 0.8% of carbon, at which point the pearlite structure will form. No large inclusions of cementite will form at the boundaries in hypoeutectoid steel. [11] The above assumes that the cooling process is very slow, allowing enough time for the carbon to migrate.[ citation needed ]

As the rate of cooling is increased the carbon will have less time to migrate to form carbide at the grain boundaries but will have increasingly large amounts of pearlite of a finer and finer structure within the grains; hence the carbide is more widely dispersed and acts to prevent slip of defects within those grains, resulting in hardening of the steel. At the very high cooling rates produced by quenching, the carbon has no time to migrate but is locked within the face-centred austenite and forms martensite. Martensite is a highly strained and stressed, supersaturated form of carbon and iron and is exceedingly hard but brittle. Depending on the carbon content, the martensitic phase takes different forms. Below 0.2% carbon, it takes on a ferrite BCC crystal form, but at higher carbon content it takes a body-centred tetragonal (BCT) structure. There is no thermal activation energy for the transformation from austenite to martensite.[ clarification needed ] There is no compositional change so the atoms generally retain their same neighbors. [12]

Martensite has a lower density (it expands during the cooling) than does austenite, so that the transformation between them results in a change of volume. In this case, expansion occurs. Internal stresses from this expansion generally take the form of compression on the crystals of martensite and tension on the remaining ferrite, with a fair amount of shear on both constituents. If quenching is done improperly, the internal stresses can cause a part to shatter as it cools. At the very least, they cause internal work hardening and other microscopic imperfections. It is common for quench cracks to form when steel is water quenched, although they may not always be visible. [13]

Heat treatment

There are many types of heat treating processes available to steel. The most common are annealing, quenching, and tempering.

Annealing is the process of heating the steel to a sufficiently high temperature to relieve local internal stresses. It does not create a general softening of the product but only locally relieves strains and stresses locked up within the material. Annealing goes through three phases: recovery, recrystallization, and grain growth. The temperature required to anneal a particular steel depends on the type of annealing to be achieved and the alloying constituents. [14]

Quenching involves heating the steel to create the austenite phase then quenching it in water or oil. This rapid cooling results in a hard but brittle martensitic structure. [12] The steel is then tempered, which is just a specialized type of annealing, to reduce brittleness. In this application the annealing (tempering) process transforms some of the martensite into cementite, or spheroidite and hence it reduces the internal stresses and defects. The result is a more ductile and fracture-resistant steel. [15]

Production

Iron ore pellets used in the production of steel LightningVolt Iron Ore Pellets.jpg
Iron ore pellets used in the production of steel

When iron is smelted from its ore, it contains more carbon than is desirable. To become steel, it must be reprocessed to reduce the carbon to the correct amount, at which point other elements can be added. In the past, steel facilities would cast the raw steel product into ingots which would be stored until use in further refinement processes that resulted in the finished product. In modern facilities, the initial product is close to the final composition and is continuously cast into long slabs, cut and shaped into bars and extrusions and heat treated to produce a final product. Today, approximately 96% of steel is continuously cast, while only 4% is produced as ingots. [16]

The ingots are then heated in a soaking pit and hot rolled into slabs, billets, or blooms. Slabs are hot or cold rolled into sheet metal or plates. Billets are hot or cold rolled into bars, rods, and wire. Blooms are hot or cold rolled into structural steel, such as I-beams and rails. In modern steel mills these processes often occur in one assembly line, with ore coming in and finished steel products coming out. [17] Sometimes after a steel's final rolling, it is heat treated for strength; however, this is relatively rare. [18]

History

Ancient

Bloomery smelting during the Middle Ages in the 5th to 15th centuries Bas fourneau.png
Bloomery smelting during the Middle Ages in the 5th to 15th centuries

Steel was known in antiquity and was produced in bloomeries and crucibles. [19] [20]

The earliest known production of steel is seen in pieces of ironware excavated from an archaeological site in Anatolia (Kaman-Kalehöyük) which are nearly 4,000 years old, dating from 1800 BC. [21] [22]

Steel was produced in Celtic Europe from around 800 BC, [23] high-carbon steel was produced in Britain from 490-375 BC, [24] [25] and ultrahigh-carbon steel was produced in the Netherlands from the 2nd-4th centuries AD. [26] The Roman author Horace identifies steel weapons such as the falcata in the Iberian Peninsula, while Noric steel was used by the Roman military. [27]

The reputation of Seric iron of South Asia (wootz steel) grew considerably in the rest of the world. [20] Metal production sites in Sri Lanka employed wind furnaces driven by the monsoon winds, capable of producing high-carbon steel. Large-scale Wootz steel production in India using crucibles occurred by the sixth century BC, the pioneering precursor to modern steel production and metallurgy. [19] [20]

The Chinese of the Warring States period (403221 BC) had quench-hardened steel, [28] while Chinese of the Han dynasty (202 BCAD 220) created steel by melting together wrought iron with cast iron, thus producing a carbon-intermediate steel by the 1st century AD. [29] [30]

There is evidence that carbon steel was made in Western Tanzania by the ancestors of the Haya people as early as 2,000 years ago by a complex process of "pre-heating" allowing temperatures inside a furnace to reach 1300 to 1400 °C. [31] [32] [33] [34] [35] [36]

Wootz and Damascus

Evidence of the earliest production of high carbon steel in South Asia is found in Kodumanal in Tamil Nadu, the Golconda area in Andhra Pradesh and Karnataka, regions of India, as well as in Samanalawewa and Dehigaha Alakanda, regions of Sri Lanka. [37] This came to be known as Wootz steel, produced in South India by about the sixth century BC and exported globally. [38] [39] The steel technology existed prior to 326 BC in the region as they are mentioned in literature of Sangam Tamil, Arabic, and Latin as the finest steel in the world exported to the Romans, Egyptian, Chinese and Arab worlds at that time – what they called Seric Iron. [40] A 200 BC Tamil trade guild in Tissamaharama, in the South East of Sri Lanka, brought with them some of the oldest iron and steel artifacts and production processes to the island from the classical period. [41] [42] [43] The Chinese and locals in Anuradhapura, Sri Lanka had also adopted the production methods of creating Wootz steel from the Chera Dynasty Tamils of South India by the 5th century AD. [44] [45] In Sri Lanka, this early steel-making method employed a unique wind furnace, driven by the monsoon winds, capable of producing high-carbon steel. [46] [47] Since the technology was acquired from the Tamilians from South India, [48] the origin of steel technology in India can be conservatively estimated at 400500 BC. [38] [47]

The manufacture of Wootz steel and Damascus steel, famous for its durability and ability to hold an edge, may have been taken by the Arabs from Persia, who took it from India. It was originally created from several different materials including various trace elements, apparently ultimately from the writings of Zosimos of Panopolis.[ citation needed ] In 327 BC, Alexander the Great was rewarded by the defeated King Porus, not with gold or silver but with 30 pounds of steel. [49] A recent study has speculated that carbon nanotubes were included in its structure, which might explain some of its legendary qualities, though, given the technology of that time, such qualities were produced by chance rather than by design. [50] Natural wind was used where the soil containing iron was heated by the use of wood. The ancient Sinhalese managed to extract a ton of steel for every 2 tons of soil, [46] a remarkable feat at the time. One such furnace was found in Samanalawewa and archaeologists were able to produce steel as the ancients did. [46] [51]

Crucible steel, formed by slowly heating and cooling pure iron and carbon (typically in the form of charcoal) in a crucible, was produced in Merv by the 9th to 10th century AD. [39] In the 11th century, there is evidence of the production of steel in Song China using two techniques: a "berganesque" method that produced inferior, inhomogeneous steel, and a precursor to the modern Bessemer process that used partial decarburization via repeated forging under a cold blast. [52]

Modern

A Bessemer converter in Sheffield, England Bessemer Converter Sheffield.jpg
A Bessemer converter in Sheffield, England

Since the 17th century, the first step in European steel production has been the smelting of iron ore into pig iron in a blast furnace. [53] Originally employing charcoal, modern methods use coke, which has proven more economical. [54] [55] [56]

Processes starting from bar iron

In these processes, pig iron made from raw iron ore was refined (fined) in a finery forge to produce bar iron, which was then used in steel-making. [53]

The production of steel by the cementation process was described in a treatise published in Prague in 1574 and was in use in Nuremberg from 1601. A similar process for case hardening armor and files was described in a book published in Naples in 1589. The process was introduced to England in about 1614 and used to produce such steel by Sir Basil Brooke at Coalbrookdale during the 1610s. [57]

The raw material for this process were bars of iron. During the 17th century, it was realized that the best steel came from oregrounds iron of a region north of Stockholm, Sweden. This was still the usual raw material source in the 19th century, almost as long as the process was used. [58] [59]

Crucible steel is steel that has been melted in a crucible rather than having been forged, with the result that it is more homogeneous. Most previous furnaces could not reach high enough temperatures to melt the steel. The early modern crucible steel industry resulted from the invention of Benjamin Huntsman in the 1740s. Blister steel (made as above) was melted in a crucible or in a furnace, and cast (usually) into ingots. [59] [60]

Processes starting from pig iron

An open hearth furnace in the Museum of Industry in Brandenburg, Germany Siemens Martin Ofen Brandenburg.jpg
An open hearth furnace in the Museum of Industry in Brandenburg, Germany
White-hot steel pouring out of an electric arc furnace in Brackenridge, Pennsylvania Allegheny Ludlum steel furnace.jpg
White-hot steel pouring out of an electric arc furnace in Brackenridge, Pennsylvania

The modern era in steelmaking began with the introduction of Henry Bessemer's process in 1855, the raw material for which was pig iron. [61] His method let him produce steel in large quantities cheaply, thus mild steel came to be used for most purposes for which wrought iron was formerly used. [62] The Gilchrist-Thomas process (or basic Bessemer process) was an improvement to the Bessemer process, made by lining the converter with a basic material to remove phosphorus.

Another 19th-century steelmaking process was the Siemens-Martin process, which complemented the Bessemer process. [59] It consisted of co-melting bar iron (or steel scrap) with pig iron.

These methods of steel production were rendered obsolete by the Linz-Donawitz process of basic oxygen steelmaking (BOS), developed in 1952, [63] and other oxygen steel making methods. Basic oxygen steelmaking is superior to previous steelmaking methods because the oxygen pumped into the furnace limited impurities, primarily nitrogen, that previously had entered from the air used, [64] and because, with respect to the open hearth process, the same quantity of steel from a BOS process is manufactured in one-twelfth the time. [63] Today, electric arc furnaces (EAF) are a common method of reprocessing scrap metal to create new steel. They can also be used for converting pig iron to steel, but they use a lot of electrical energy (about 440 kWh per metric ton), and are thus generally only economical when there is a plentiful supply of cheap electricity. [65]

Industry

Steel production (in million tons) by country as of 2007 Steel production by country map.PNG
Steel production (in million tons) by country as of 2007

The steel industry is often considered an indicator of economic progress, because of the critical role played by steel in infrastructural and overall economic development. [66] In 1980, there were more than 500,000 U.S. steelworkers. By 2000, the number of steelworkers had fallen to 224,000. [67]

The economic boom in China and India caused a massive increase in the demand for steel. Between 2000 and 2005, world steel demand increased by 6%. Since 2000, several Indian [68] and Chinese [69] steel firms have expanded to meet demand, such as Tata Steel (which bought Corus Group in 2007), Baosteel Group and Shagang Group. As of 2017, though, ArcelorMittal is the world's largest steel producer. [70]

In 2005, the British Geological Survey stated China was the top steel producer with about one-third of the world share; Japan, Russia, and the United States were second, third, and fourth, respectively, according to the survey. [71] The large production capacity of steel results also in a significant amount of carbon dioxide emissions inherent related to the main production route.

At the end of 2008, the steel industry faced a sharp downturn that led to many cut-backs. [72]

In 2021, it was estimated that around 7% of the global greenhouse gas emissions resulted from the steel industry. [73] [74] Reduction of these emissions are expected to come from a shift in the main production route using cokes, more recycling of steel and the application of carbon capture and storage or carbon capture and utilization technology.

Recycling

Steel is one of the world's most-recycled materials, with a recycling rate of over 60% globally; [3] in the United States alone, over 82,000,000 metric tons (81,000,000 long tons; 90,000,000 short tons) were recycled in the year 2008, for an overall recycling rate of 83%. [75]

As more steel is produced than is scrapped, the amount of recycled raw materials is about 40% of the total of steel produced - in 2016, 1,628,000,000 tonnes (1.602×109 long tons; 1.795×109 short tons) of crude steel was produced globally, with 630,000,000 tonnes (620,000,000 long tons; 690,000,000 short tons) recycled. [76]

Contemporary

Bethlehem Steel in Bethlehem, Pennsylvania was one of the world's largest manufacturers of steel before its closure in 2003. Bethlehem Steel.jpg
Bethlehem Steel in Bethlehem, Pennsylvania was one of the world's largest manufacturers of steel before its closure in 2003.

Carbon

Modern steels are made with varying combinations of alloy metals to fulfill many purposes. [7] Carbon steel, composed simply of iron and carbon, accounts for 90% of steel production. [5] Low alloy steel is alloyed with other elements, usually molybdenum, manganese, chromium, or nickel, in amounts of up to 10% by weight to improve the hardenability of thick sections. [5] High strength low alloy steel has small additions (usually < 2% by weight) of other elements, typically 1.5% manganese, to provide additional strength for a modest price increase. [77]

Recent Corporate Average Fuel Economy (CAFE) regulations have given rise to a new variety of steel known as Advanced High Strength Steel (AHSS). This material is both strong and ductile so that vehicle structures can maintain their current safety levels while using less material. There are several commercially available grades of AHSS, such as dual-phase steel, which is heat treated to contain both a ferritic and martensitic microstructure to produce a formable, high strength steel. [78] Transformation Induced Plasticity (TRIP) steel involves special alloying and heat treatments to stabilize amounts of austenite at room temperature in normally austenite-free low-alloy ferritic steels. By applying strain, the austenite undergoes a phase transition to martensite without the addition of heat. [79] Twinning Induced Plasticity (TWIP) steel uses a specific type of strain to increase the effectiveness of work hardening on the alloy. [80]

Carbon Steels are often galvanized, through hot-dip or electroplating in zinc for protection against rust. [81]

Alloy

Forging a structural member out of steel Alcator C-Mod superstructure forging 1.jpg
Forging a structural member out of steel
Cor-Ten rust coating Rust-AH-2022.jpg
Cor-Ten rust coating

Stainless steels contain a minimum of 11% chromium, often combined with nickel, to resist corrosion. Some stainless steels, such as the ferritic stainless steels are magnetic, while others, such as the austenitic, are nonmagnetic. [82] Corrosion-resistant steels are abbreviated as CRES.

Alloy steels are plain-carbon steels in which small amounts of alloying elements like chromium and vanadium have been added. Some more modern steels include tool steels, which are alloyed with large amounts of tungsten and cobalt or other elements to maximize solution hardening. This also allows the use of precipitation hardening and improves the alloy's temperature resistance. [5] Tool steel is generally used in axes, drills, and other devices that need a sharp, long-lasting cutting edge. Other special-purpose alloys include weathering steels such as Cor-ten, which weather by acquiring a stable, rusted surface, and so can be used un-painted. [83] Maraging steel is alloyed with nickel and other elements, but unlike most steel contains little carbon (0.01%). This creates a very strong but still malleable steel. [84]

Eglin steel uses a combination of over a dozen different elements in varying amounts to create a relatively low-cost steel for use in bunker buster weapons, and Hadfield steel (after Sir Robert Hadfield) or manganese steel contains 12–14% manganese which, when abraded, strain-hardens to form a very hard skin which resists wearing. Uses of this particular alloy include tank tracks, bulldozer blade edges, and cutting blades on the jaws of life. [85]

Standards

Most of the more commonly used steel alloys are categorized into various grades by standards organizations. For example, the Society of Automotive Engineers has a series of grades defining many types of steel. [86] The American Society for Testing and Materials has a separate set of standards, which define alloys such as A36 steel, the most commonly used structural steel in the United States. [87] The JIS also defines a series of steel grades that are being used extensively in Japan as well as in developing countries.

Uses

A roll of steel wool Steel-wool.jpg
A roll of steel wool

Iron and steel are used widely in the construction of roads, railways, other infrastructure, appliances, and buildings. Most large modern structures, such as stadiums and skyscrapers, bridges, and airports, are supported by a steel skeleton. Even those with a concrete structure employ steel for reinforcing. It sees widespread use in major appliances and cars. Despite the growth in usage of aluminium, steel is still the main material for car bodies. Steel is used in a variety of other construction materials, such as bolts, nails and screws and other household products and cooking utensils. [88]

Other common applications include shipbuilding, pipelines, mining, offshore construction, aerospace, white goods (e.g. washing machines), heavy equipment such as bulldozers, office furniture, steel wool, tool, and armour in the form of personal vests or vehicle armour (better known as rolled homogeneous armour in this role).

Historical

A carbon steel knife Carbon steel knife.jpg
A carbon steel knife

Before the introduction of the Bessemer process and other modern production techniques, steel was expensive and was only used where no cheaper alternative existed, particularly for the cutting edge of knives, razors, swords, and other items where a hard, sharp edge was needed. It was also used for springs, including those used in clocks and watches. [59]

With the advent of faster and cheaper production methods, steel has become easier to obtain and much cheaper. It has replaced wrought iron for a multitude of purposes. However, the availability of plastics in the latter part of the 20th century allowed these materials to replace steel in some applications due to their lower fabrication cost and weight. [89] Carbon fiber is replacing steel in some cost insensitive applications such as sports equipment and high-end automobiles.

Long

A steel bridge The viaduct La Polvorilla, Salta Argentina.jpg
A steel bridge
A steel pylon suspending overhead power lines Steel tower.jpg
A steel pylon suspending overhead power lines

Flat carbon

Weathering (COR-TEN)

Stainless

A stainless steel gravy boat Sauce boat.jpg
A stainless steel gravy boat

Low-background

Steel manufactured after World War II became contaminated with radionuclides by nuclear weapons testing. Low-background steel, steel manufactured prior to 1945, is used for certain radiation-sensitive applications such as Geiger counters and radiation shielding.

See also

Related Research Articles

<span class="mw-page-title-main">Alloy</span> Mixture or metallic solid solution composed of two or more elements

An alloy is a mixture of chemical elements of which at least one is a metal. Unlike chemical compounds with metallic bases, an alloy will retain all the properties of a metal in the resulting material, such as electrical conductivity, ductility, opacity, and luster, but may have properties that differ from those of the pure metals, such as increased strength or hardness. In some cases, an alloy may reduce the overall cost of the material while preserving important properties. In other cases, the mixture imparts synergistic properties to the constituent metal elements such as corrosion resistance or mechanical strength.

<span class="mw-page-title-main">Heat treating</span> Process of heating something to alter it

Heat treating is a group of industrial, thermal and metalworking processes used to alter the physical, and sometimes chemical, properties of a material. The most common application is metallurgical. Heat treatments are also used in the manufacture of many other materials, such as glass. Heat treatment involves the use of heating or chilling, normally to extreme temperatures, to achieve the desired result such as hardening or softening of a material. Heat treatment techniques include annealing, case hardening, precipitation strengthening, tempering, carburizing, normalizing and quenching. Although the term heat treatment applies only to processes where the heating and cooling are done for the specific purpose of altering properties intentionally, heating and cooling often occur incidentally during other manufacturing processes such as hot forming or welding.

<span class="mw-page-title-main">Martensite</span> Type of steel crystalline structure

Martensite is a very hard form of steel crystalline structure. It is named after German metallurgist Adolf Martens. By analogy the term can also refer to any crystal structure that is formed by diffusionless transformation.

<span class="mw-page-title-main">Cementite</span> Compound of iron and carbon

Cementite (or iron carbide) is a compound of iron and carbon, more precisely an intermediate transition metal carbide with the formula Fe3C. By weight, it is 6.67% carbon and 93.3% iron. It has an orthorhombic crystal structure. It is a hard, brittle material, normally classified as a ceramic in its pure form, and is a frequently found and important constituent in ferrous metallurgy. While cementite is present in most steels and cast irons, it is produced as a raw material in the iron carbide process, which belongs to the family of alternative ironmaking technologies. The name cementite originated from the theory of Floris Osmond and J. Werth, in which the structure of solidified steel consists of a kind of cellular tissue, with ferrite as the nucleus and Fe3C the envelope of the cells. The carbide therefore cemented the iron.

<span class="mw-page-title-main">Austenite</span> Metallic, non-magnetic allotrope of iron or a solid solution of iron, with an alloying element

Austenite, also known as gamma-phase iron (γ-Fe), is a metallic, non-magnetic allotrope of iron or a solid solution of iron with an alloying element. In plain-carbon steel, austenite exists above the critical eutectoid temperature of 1000 K (727 °C); other alloys of steel have different eutectoid temperatures. The austenite allotrope is named after Sir William Chandler Roberts-Austen (1843–1902). It exists at room temperature in some stainless steels due to the presence of nickel stabilizing the austenite at lower temperatures.

<span class="mw-page-title-main">Bainite</span> Plate-like microstructure in steels

Bainite is a plate-like microstructure that forms in steels at temperatures of 125–550 °C. First described by E. S. Davenport and Edgar Bain, it is one of the products that may form when austenite is cooled past a temperature where it is no longer thermodynamically stable with respect to ferrite, cementite, or ferrite and cementite. Davenport and Bain originally described the microstructure as being similar in appearance to tempered martensite.

<span class="mw-page-title-main">Bulat steel</span> Steel alloy known in Russia from medieval times

Bulat is a type of steel alloy known in Russia from medieval times; it was regularly mentioned in Russian legends as the material of choice for cold steel. The name булат is a Russian transliteration of the Persian word fulad, meaning steel. This type of steel was used by the armies of nomadic peoples. Bulat steel was the main type of steel used for swords in the armies of Genghis Khan. Bulat Steel is generally agreed to be a Russian name for wootz steel, the production method of which has been lost for centuries, and the bulat steel used today makes use of a more recently developed technique.

<span class="mw-page-title-main">Crucible steel</span> Type of steel

Crucible steel is steel made by melting pig 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 and 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, turning both into steel. 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 and a lower-carbon steel. 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. The steel was often much higher in carbon content and in quality in comparison with other methods of steel production of the time because of the use of fluxes. The steel was usually worked very little and at relatively low temperatures to avoid any decarburization, hot short crumbling, or excess diffusion of carbon; just enough hammering to form the shape of a sword. With a carbon content close to that of cast iron, it usually required no heat treatment after shaping other than air cooling to achieve the correct hardness, relying on composition alone. The higher-carbon steel provided a very hard edge, but the lower-carbon steel helped to increase the toughness, helping to decrease the chance of chipping, cracking, or breaking.

<span class="mw-page-title-main">High-strength low-alloy steel</span> Type of alloy steel

High-strength low-alloy steel (HSLA) is a type of alloy steel that provides better mechanical properties or greater resistance to corrosion than carbon steel. HSLA steels vary from other steels in that they are not made to meet a specific chemical composition but rather specific mechanical properties. They have a carbon content between 0.05 and 0.25% to retain formability and weldability. Other alloying elements include up to 2.0% manganese and small quantities of copper, nickel, niobium, nitrogen, vanadium, chromium, molybdenum, titanium, calcium, rare-earth elements, or zirconium. Copper, titanium, vanadium, and niobium are added for strengthening purposes. These elements are intended to alter the microstructure of carbon steels, which is usually a ferrite-pearlite aggregate, to produce a very fine dispersion of alloy carbides in an almost pure ferrite matrix. This eliminates the toughness-reducing effect of a pearlitic volume fraction yet maintains and increases the material's strength by refining the grain size, which in the case of ferrite increases yield strength by 50% for every halving of the mean grain diameter. Precipitation strengthening plays a minor role, too. Their yield strengths can be anywhere between 250–590 megapascals (36,000–86,000 psi). Because of their higher strength and toughness HSLA steels usually require 25 to 30% more power to form, as compared to carbon steels.

<span class="mw-page-title-main">Wootz steel</span> Type of crucible steel

Wootz steel, also known as Seric steel, is a crucible steel characterized by a pattern of bands and high carbon content. These bands are formed by sheets of microscopic carbides within a tempered martensite or pearlite matrix in higher-carbon steel, or by ferrite and pearlite banding in lower-carbon steels. It was a pioneering steel alloy developed in southern India in the mid-1st millennium BC and exported globally.

<span class="mw-page-title-main">Carbon steel</span> Steel in which the main interstitial alloying constituent is carbon

Carbon steel is a steel with carbon content from about 0.05 up to 2.1 percent by weight. The definition of carbon steel from the American Iron and Steel Institute (AISI) states:

<span class="mw-page-title-main">Quenching</span> Rapid cooling of a workpiece to obtain certain material properties

In materials science, quenching is the rapid cooling of a workpiece in water, gas, oil, polymer, air, or other fluids to obtain certain material properties. A type of heat treating, quenching prevents undesired low-temperature processes, such as phase transformations, from occurring. It does this by reducing the window of time during which these undesired reactions are both thermodynamically favorable and kinetically accessible; for instance, quenching can reduce the crystal grain size of both metallic and plastic materials, increasing their hardness.

<span class="mw-page-title-main">Tempering (metallurgy)</span> Process of heat treating used to increase the toughness of iron-based alloys

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

<span class="mw-page-title-main">Hardenability</span> Depth to which a metal is hardened after being submitted to a thermal treatment

Hardenability is the depth to which a steel is hardened after putting it through a heat treatment process. It should not be confused with hardness, which is a measure of a sample's resistance to indentation or scratching. It is an important property for welding, since it is inversely proportional to weldability, that is, the ease of welding a material.

Ferroalloy refers to various alloys of iron with a high proportion of one or more other elements such as manganese (Mn), aluminium (Al), or silicon (Si). They are used in the production of steels and alloys. The alloys impart distinctive qualities to steel and cast iron or serve important functions during production and are, therefore, closely associated with the iron and steel industry, the leading consumer of ferroalloys. The leading producers of ferroalloys in 2014 were China, South Africa, India, Russia and Kazakhstan, which accounted for 84% of the world production. World production of ferroalloys was estimated as 52.8 million tonnes in 2015.

Cryogenic hardening is a cryogenic treatment process where the material is cooled to approximately −185 °C (−301 °F), usually using liquid nitrogen. It can have a profound effect on the mechanical properties of certain steels, provided their composition and prior heat treatment are such that they retain some austenite at room temperature. It is designed to increase the amount of martensite in the steel's crystal structure, increasing its strength and hardness, sometimes at the cost of toughness. Presently this treatment is being used on tool steels, high-carbon, high-chromium steels and in some cases to cemented carbide to obtain excellent wear resistance. Recent research shows that there is precipitation of fine carbides in the matrix during this treatment which imparts very high wear resistance to the steels.

In metallurgy and materials science, annealing is a heat treatment that alters the physical and sometimes chemical properties of a material to increase its ductility and reduce its hardness, making it more workable. It involves heating a material above its recrystallization temperature, maintaining a suitable temperature for an appropriate amount of time and then cooling.

<span class="mw-page-title-main">Isothermal transformation diagram</span>

Isothermal transformation diagrams are plots of temperature versus time. They are generated from percentage transformation-vs time measurements, and are useful for understanding the transformations of an alloy steel at elevated temperatures.

<span class="mw-page-title-main">Austempering</span>

Austempering is heat treatment that is applied to ferrous metals, most notably steel and ductile iron. In steel it produces a bainite microstructure whereas in cast irons it produces a structure of acicular ferrite and high carbon, stabilized austenite known as ausferrite. It is primarily used to improve mechanical properties or reduce / eliminate distortion. Austempering is defined by both the process and the resultant microstructure. Typical austempering process parameters applied to an unsuitable material will not result in the formation of bainite or ausferrite and thus the final product will not be called austempered. Both microstructures may also be produced via other methods. For example, they may be produced as-cast or air cooled with the proper alloy content. These materials are also not referred to as austempered.

<span class="mw-page-title-main">Mangalloy</span> Alloy steel containing around 13% manganese

Mangalloy, also called manganese steel or Hadfield steel, is an alloy steel containing an average of around 13% manganese. Mangalloy is known for its high impact strength and resistance to abrasion once in its work-hardened state.

References

  1. R., Allen. "(1979). International Competition in Iron and Steel, 1850-1913". JSTOR. Cambridge university. JSTOR   2120336 . Retrieved November 13, 2020.
  2. "Decarbonization in steel | McKinsey". www.mckinsey.com. Retrieved 2022-05-20.
  3. 1 2 Hartman, Roy A. (2009). "Recycling". Encarta . Archived from the original on 2008-04-14.
  4. Harper, Douglas. "steel". Online Etymology Dictionary .
  5. 1 2 3 4 5 Ashby, Michael F. & Jones, David R.H. (1992) [1986]. Engineering Materials 2 (with corrections ed.). Oxford: Pergamon Press. ISBN   0-08-032532-7.
  6. 1 2 Smelting. Encyclopædia Britannica. 2007.
  7. 1 2 3 "Alloying of Steels". Metallurgical Consultants. 2006-06-28. Archived from the original on 2007-02-21. Retrieved 2007-02-28.
  8. Elert, Glenn. "Density of Steel" . Retrieved 2009-04-23.
  9. Sources differ on this value so it has been rounded to 2.1%, however the exact value is rather academic because plain-carbon steel is very rarely made with this level of carbon. See:
  10. Smith & Hashemi 2006 , p. 363.
  11. Smith & Hashemi 2006 , pp. 365–372.
  12. 1 2 Smith & Hashemi 2006 , pp. 373–378.
  13. "Quench hardening of steel". keytometals.com. Archived from the original on 2009-02-17. Retrieved 2009-07-19.
  14. Smith & Hashemi 2006 , p. 249.
  15. Smith & Hashemi 2006 , p. 388.
  16. Smith & Hashemi 2006 , p. 361
  17. Smith & Hashemi 2006 , pp. 361–362.
  18. Bugayev et al. 2001 , p. 225
  19. 1 2 Davidson 1994, p. 20.
  20. 1 2 3 Srinivasan, S.; Ranganathan, S. (1994). "The Sword in Anglo-Saxon England: Its Archaeology and Literature". Bangalore: Department of Metallurgy, Indian Institute of Science. ISBN   0-85115-355-0. Archived from the original on 2018-11-19.
  21. Akanuma, H. (2005). "The significance of the composition of excavated iron fragments taken from Stratum III at the site of Kaman-Kalehöyük, Turkey". Anatolian Archaeological Studies. 14. Tokyo: Japanese Institute of Anatolian Archaeology: 147–158.
  22. "Ironware piece unearthed from Turkey found to be oldest steel". The Hindu. Chennai, India. 2009-03-26. Archived from the original on 2009-03-29. Retrieved 2022-08-13.
  23. Wells, Peter (1995). "Resources and Industry". In Green, Miranda (ed.). The Celtic World. Routledge. p. 218. ISBN   9781135632434. Plainer (1968) has shown that already at the start of the Iron Age, around 800 BC, some (Celtic) smiths had mastered the techniques for producing fine, hard and sharp-cutting edges by carburising blades to make the iron-carbon alloy, steel.
  24. "East Lothian's Broxmouth fort reveals edge of steel". BBC News. 15 January 2014.
  25. An Inherited Place: Broxmouth Hillfort and the South-East Scottish Iron Age. Society of Antiquaries of Scotland. 2013. ISBN   978-1-908332-05-9.
  26. Godfrey, Evelyne; et al. (2004). "A Germanic ultrahigh carbon steel punch of the Late Roman-Iron Age". Journal of Archaeological Science. 31 (8): 1117–1125. Bibcode:2004JArSc..31.1117G. doi:10.1016/j.jas.2004.02.002.
  27. "Noricus ensis", Horace, Odes, i. 16.9
  28. Wagner, Donald B. (1993). Iron and Steel in Ancient China: Second Impression, With Corrections. Leiden: E.J. Brill. p. 243. ISBN   90-04-09632-9.
  29. Needham, Joseph (1986). Science and Civilization in China: Volume 4, Part 3, Civil Engineering and Nautics. Taipei: Caves Books, Ltd. p. 563.
  30. Gernet, Jacques (1982). A History of Chinese Civilization. Cambridge: Cambridge University Press. p. 69. ISBN   0-521-49781-7.
  31. Schmidt, Peter; Avery, Donald (1978). "Complex Iron Smelting and Prehistoric Culture in Tanzania". Science. 201 (4361): 1085–1089. Bibcode:1978Sci...201.1085S. doi:10.1126/science.201.4361.1085. JSTOR   1746308. PMID   17830304. S2CID   37926350.
  32. Schmidt, Peter; Avery, Donald (1983). "More Evidence for an Advanced Prehistoric Iron Technology in Africa". Journal of Field Archaeology. 10 (4): 421–434. doi:10.1179/009346983791504228.
  33. Schmidt, Peter (1978). Historical Archaeology: A Structural Approach in an African Culture. Westport, CT: Greenwood Press.
  34. Avery, Donald; Schmidt, Peter (1996). "Preheating: Practice or illusion". The Culture and Technology of African Iron Production. Gainesville: University of Florida Press. pp. 267–276.
  35. Schmidt, Peter (2019). "Science in Africa: A history of ingenuity and invention in African iron technology". In Worger, W; Ambler, C; Achebe, N (eds.). A Companion to African History. Hoboken, NJ: Wiley Blackwell. pp. 267–288.
  36. Childs, S. Terry (1996). "Technological history and culture in western Tanzania". In Schmidt, P. (ed.). The Culture and Technology of African Iron Production. Gainesville, FL: University of Florida Press.
  37. Wilford, John Noble (1996-02-06). "Ancient Smelter Used Wind To Make High-Grade Steel". The New York Times.
  38. 1 2 Srinivasan, Sharada; Ranganathan, Srinivasa (2004). India's Legendary Wootz Steel: An Advanced Material of the Ancient World. National Institute of Advanced Studies. OCLC   82439861. Archived from the original on 2019-02-11. Retrieved 2014-12-05.
  39. 1 2 Feuerbach, Ann (2005). "An investigation of the varied technology found in swords, sabres and blades from the Russian Northern Caucasus" (PDF). IAMS. 25: 27–43 (p. 29). Archived from the original (PDF) on 2011-04-30.
  40. Srinivasan, Sharada (1994). "Wootz crucible steel: a newly discovered production site in South India". Papers from the Institute of Archaeology. 5: 49–59. doi: 10.5334/pia.60 .
  41. Hobbies – Volume 68, Issue 5 – p. 45. Lightner Publishing Company (1963)
  42. Mahathevan, Iravatham (24 June 2010). "An epigraphic perspective on the antiquity of Tamil". The Hindu . Archived from the original on 1 July 2010. Retrieved 31 October 2010.
  43. Ragupathy, P (28 June 2010). "Tissamaharama potsherd evidences ordinary early Tamils among population". Tamilnet. Retrieved 31 October 2010.
  44. Needham, Joseph (1986). Science and Civilization in China: Volume 4, Part 1, Civil Engineering and Nautics (PDF). Taipei: Caves Books, Ltd. p. 282. ISBN   0-521-05802-3. Archived from the original (PDF) on 2017-07-03. Retrieved 2017-08-04.
  45. Manning, Charlotte Speir. Ancient and Mediæval India. Volume 2. ISBN   978-0-543-92943-3.
  46. 1 2 3 Juleff, G. (1996). "An ancient wind powered iron smelting technology in Sri Lanka". Nature . 379 (3): 60–63. Bibcode:1996Natur.379...60J. doi:10.1038/379060a0. S2CID   205026185.
  47. 1 2 Coghlan, Herbert Henery. (1977). Notes on prehistoric and early iron in the Old World. Oxprint. pp. 99–100
  48. Manning, Charlotte Speir. Ancient and Medieval India. Volume 2. ISBN   978-0-543-92943-3.
  49. Durant, Will (1942). The Story of Civilization, Our Oriental Heritage. Simon and Schuster. p. 529. ISBN   0-671-54800-X.
  50. Sanderson, Katharine (2006-11-15). "Sharpest cut from nanotube sword". Nature News. doi: 10.1038/news061113-11 . S2CID   136774602.
  51. Wayman, M.L. & Juleff, G. (1999). "Crucible Steelmaking in Sri Lanka". Historical Metallurgy. 33 (1): 26.
  52. Hartwell, Robert (1966). "Markets, Technology and the Structure of Enterprise in the Development of the Eleventh Century Chinese Iron and Steel Industry". Journal of Economic History . 26: 53–54. doi:10.1017/S0022050700061842. S2CID   154556274.
  53. 1 2 Tylecote, R.F. (1992) A history of metallurgy 2nd ed., Institute of Materials, London. pp. 95–99 and 102–105. ISBN   0-901462-88-8.
  54. Raistrick, A. (1953) A Dynasty of Ironfounders.
  55. Hyde, C.K. (1977) Technological Change and the British iron industry. Princeton
  56. Trinder, B. (2000) The Industrial Revolution in Shropshire. Chichester.
  57. Barraclough 1984, pp. 48–52.
  58. King, P.W. (2003). "The Cartel in Oregrounds Iron: trading in the raw material for steel during the eighteenth century". Journal of Industrial History. 6 (1): 25–49.
  59. 1 2 3 4 "Iron and steel industry". Encyclopædia Britannica. 2007.
  60. Barraclough, K.C. (1984) Steel before Bessemer: II Crucible Steel: the growth of technology. The Metals Society, London.
  61. Swank, James Moore (1892). History of the Manufacture of Iron in All Ages. Burt Franklin Publisher. ISBN   0-8337-3463-6.
  62. Bessemer process. Vol. 2. Encyclopædia Britannica. 2005. p. 168.
  63. 1 2 Sherman, Zander (4 September 2019). "How my great-grandfather's Dofasco steel empire rose and fell, and his descendants with it". The Globe and Mail Inc.
  64. Basic oxygen process. Encyclopædia Britannica. 2007.
  65. Fruehan & Wakelin 1998, pp. 48–52.
  66. "Steel Industry". Archived from the original on 2009-06-18. Retrieved 2009-07-12.
  67. " Congressional Record V. 148, Pt. 4, April 11, 2002 to April 24, 2002 ". United States Government Printing Office.
  68. Chopra, Anuj (February 12, 2007). "India's steel industry steps onto world stage". Cristian Science Monitor. Retrieved 2009-07-12.
  69. "Worldsteel | World crude steel output decreases by -2.8% in 2015". Archived from the original on 2017-02-02. Retrieved 2016-12-26.
  70. "Top Steelmakers in 2017" (PDF). World Steel Association. Archived from the original (PDF) on August 23, 2018. Retrieved August 22, 2018.
  71. "Long-term planning needed to meet steel demand". The News. 2008-03-01. Archived from the original on 2010-11-02. Retrieved 2010-11-02.
  72. Uchitelle, Louis (2009-01-01). "Steel Industry, in Slump, Looks to Federal Stimulus". The New York Times. Retrieved 2009-07-19.
  73. Rossi, Marcello (2022-08-04). "The Race to Remake the $2.5 Trillion Steel Industry With Green Steel". Singularity Hub. Retrieved 2022-08-06.
  74. "Global Steel Industry's GHG Emissions". Global Efficiency Intelligence. 6 January 2021. Retrieved 2022-08-06.
  75. Fenton, Michael D (2008). "Iron and Steel Scrap". In United States Geological Survey (ed.). Minerals Yearbook 2008, Volume 1: Metals and Minerals. Government Printing Office. ISBN   978-1-4113-3015-3.
  76. The World Steel Association (2018-03-01). "Steel and raw materials" (PDF).
  77. "High strength low alloy steels". Schoolscience.co.uk. Archived from the original on 2020-09-21. Retrieved 2007-08-14.
  78. "Dual-phase steel". Intota Expert Knowledge Services. Archived from the original on 2011-05-25. Retrieved 2007-03-01.
  79. Werner, Ewald. "Transformation Induced Plasticity in low alloyed TRIP-steels and microstructure response to a complex stress history". Archived from the original on December 23, 2007. Retrieved 2007-03-01.
  80. Mirko, Centi; Saliceti Stefano. "Transformation Induced Plasticity (TRIP), Twinning Induced Plasticity (TWIP) and Dual-Phase (DP) Steels". Tampere University of Technology. Archived from the original on 2008-03-07. Retrieved 2007-03-01.
  81. Galvanic protection. Encyclopædia Britannica. 2007.
  82. "Steel Glossary". American Iron and Steel Institute (AISI). Retrieved 2006-07-30.
  83. "Steel Interchange". American Institute of Steel Construction Inc. (AISC). Archived from the original on 2007-12-22. Retrieved 2007-02-28.
  84. "Properties of Maraging Steels". Archived from the original on 2009-02-25. Retrieved 2009-07-19.
  85. Tweedale, Geoffrey, ed. (1987). Sheffield Steel and America: A Century of Commercial and Technological Independence. Cambridge University Press. pp. 57–62.
  86. Bringas, John E. (2004). Handbook of Comparative World Steel Standards: Third Edition (PDF) (3rd. ed.). ASTM International. p. 14. ISBN   0-8031-3362-6. Archived from the original (PDF) on 2007-01-27.
  87. Steel Construction Manual, 8th Edition, second revised edition, American Institute of Steel Construction, 1986, ch. 1 pp. 1–5
  88. Ochshorn, Jonathan (2002-06-11). "Steel in 20th Century Architecture". Encyclopedia of Twentieth Century Architecture. Retrieved 2010-04-26.
  89. Venables, John D.; Girifalco, Louis A.; Patel, C. Kumar N.; McCullough, R.L.; Marchant, Roger Eric; Kukich, Diane S. (2007). Materials science. Encyclopædia Britannica.

Bibliography

Further reading