Bloomery

Last updated

A bloomery in operation. The bloom will eventually be drawn out of the bottom hole. Swieto Slaska piec p.jpg
A bloomery in operation. The bloom will eventually be drawn out of the bottom hole.

A bloomery is a type of metallurgical furnace once used widely for smelting iron from its oxides. The bloomery was the earliest form of smelter capable of smelting iron. Bloomeries produce a porous mass of iron and slag called a bloom. The mix of slag and iron in the bloom, termed sponge iron , is usually consolidated and further forged into wrought iron. Blast furnaces, which produce pig iron, have largely superseded bloomeries.

Contents

Process

An iron bloom just removed from the furnace: Surrounding it are pieces of slag that have been pounded off by the hammer. Iron bloom.jpg
An iron bloom just removed from the furnace: Surrounding it are pieces of slag that have been pounded off by the hammer.

A bloomery consists of a pit or chimney with heat-resistant walls made of earth, clay, or stone. Near the bottom, one or more pipes (made of clay or metal) enter through the side walls. These pipes, called tuyeres , allow air to enter the furnace, either by natural draught or forced with bellows or a trompe. An opening at the bottom of the bloomery may be used to remove the bloom, or the bloomery can be tipped over and the bloom removed from the top.

The first step taken before the bloomery can be used is the preparation of the charcoal and the iron ore. Charcoal is nearly pure carbon, which, when burned, both produces the high temperature needed for the smelting process and provides the carbon monoxide needed for reduction of the metal.

The ore is broken into small pieces and usually roasted in a fire, to make rock-based ores easier to break up, bake out some impurities, and (to a lesser extent) to remove any moisture in the ore. Any large impurities (as silica) in the ore can be removed as it is crushed. The desired particle size depends primarily on which of several ore types may be available, which will also have a relationship to the layout and operation of the furnace, of which a number of regional, historic/traditional forms exist. Natural iron ores can vary considerably in oxide form (Fe
2
O
3
/ Fe
3
O
4
/ FeO(OH)), and importantly in relative iron content. Since slag from previous blooms may have a high iron content, it can also be broken up and may be recycled into the bloomery with the new ore.

In operation, after the bloomery is heated typically with a wood fire, shifting to burning sized charcoal, iron ore and additional charcoal are introduced through the top. Again, traditional methods vary, but normally smaller charges of ore are added at the start of the main smelting sequence, increasing to larger amounts as the smelt progresses. Overall, a typical ratio of total charcoal to ore added is in a roughly one-to-one ratio. Inside the furnace, carbon monoxide from the incomplete combustion of the charcoal reduces the iron oxides in the ore to metallic iron without melting the ore; this allows the bloomery to operate at lower temperatures than the melting temperature of the ore. As the desired product of a bloomery is iron that is easily forgeable, it requires a low carbon content. The temperature and ratio of charcoal to iron ore must be carefully controlled to keep the iron from absorbing too much carbon and thus becoming unforgeable. Cast iron occurs when the iron absorbs 2% to 4% carbon. Because the bloomery is self-fluxing, the addition of limestone is not required to form a slag.

The small particles of iron produced in this way fall to the bottom of the furnace, where they combine with molten slag, often consisting of fayalite, a compound of silicon, oxygen, and iron mixed with other impurities from the ore. The hot liquid slag, running to the bottom of the furnace, cools against the base and lower side walls of the furnace, effectively forming a bowl still containing fluid slag. As the individual iron particles form, they fall into this bowl and sinter together under their own weight, forming a spongy mass referred to as the bloom. Because the bloom is typically porous, and its open spaces can be full of slag, the extracted mass must be beaten with heavy hammers to both compress voids and drive out any molten slag remaining. This process may require several additional heating and compaction cycles, working at high 'welding' temperatures. Iron treated this way is said to be wrought (worked), and the resulting iron, with reduced amounts of slag, is called wrought iron or bar iron. Because of the creation process, individual blooms can often have differing carbon contents between the original top and bottom surfaces, differences that will also be somewhat blended together through the flattening, folding, and hammer-welding sequences. Producing blooms coated in steel (higher carbon) by manipulating the charge of and air flow to the bloomery is also possible. [1]

As the era of modern commercial steelmaking began, the word "bloom" was extended to another sense referring to an intermediate-stage piece of steel, of a size comparable to many traditional iron blooms, that was ready to be further worked into billet.

History

A drawing of a simple bloomery and bellows. PSM V38 D159 Persian method of smelting iron.jpg
A drawing of a simple bloomery and bellows.
Bloomery smelting during the Middle Ages, as depicted in the De Re Metallica by Georgius Agricola, 1556 Bas fourneau.png
Bloomery smelting during the Middle Ages, as depicted in the De Re Metallica by Georgius Agricola, 1556

The onset of the Iron Age in most parts of the world coincides with the first widespread use of the bloomery. While earlier examples of iron are found, their high nickel content indicates that this is meteoric iron. Other early samples of iron may have been produced by accidental introduction of iron ore in copper-smelting operations. Iron appears to have been smelted in the Middle East as early as 3000 BC, but coppersmiths, not being familiar with iron, did not put it to use until much later. In the West, iron began to be used around 1200 BC. [2]

East Asia

China has long been considered the exception to the general use of bloomeries. The Chinese are thought to have skipped the bloomery process completely, starting with the blast furnace and the finery forge to produce wrought iron; by the fifth century BC, metalworkers in the southern state of Wu had invented the blast furnace and the means to both cast iron and to decarburize the carbon-rich pig iron produced in a blast furnace to a low-carbon, wrought iron-like material. Recent evidence, however, shows that bloomeries were used earlier in ancient China, migrating in from the west as early as 800 BC, before being supplanted by the locally developed blast furnace. Supporting this theory was the discovery of "more than ten" iron-digging implements found in the tomb of Duke Jing of Qin (d. 537 BCE), whose tomb is located in Fengxiang County, Shaanxi (a museum exists on the site today). [3]

Sub-Saharan Africa

The earliest records of bloomery-type furnaces in East Africa are discoveries of smelted iron and carbon in Nubia n ancient Sudan dated at least to the seventh to the sixth century BC. The ancient bloomeries that produced metal tools for the Nubians and Kushites produced a surplus for sale. All traditional sub-Saharan African iron-smelting processes are variants of the bloomery process. [4] There is considerable discussion about the origins of iron metallurgy in Africa. Smelting in bloomery type furnaces in West Africa and forging of tools appeared in the Nok culture of central Nigeria by at least 550 BC and possibly several centuries earlier. [5] [6] Also, evidence indicates iron smelting with bloomery-style furnaces dated to 750 BC in Opi (Augustin Holl 2009) and Lejja dated to 2,000 BC (Pamela Eze-Uzomaka 2009), both sites in the Nsukka region of southeast Nigeria in what is now Igboland. [7] [8] [6] The site of Gbabiri, in the Central African Republic, has also yielded evidence of iron metallurgy, from a reduction furnace and blacksmith workshop, with earliest dates of 896–773 and 907–796 BC, respectively. [6]

South Asia

During a hydroelectric plant project, in the southern foothills of the Central Highlands, Samanalawewa, in Sri Lanka, a wind-driven furnace was found in an excavation site. Such furnaces were powered by the monsoon winds and have been dated to 300 BC using radiocarbon-dating techniques. These ancient Lankan furnaces might have produced the best-quality steel for legendary Damascus swords as referred in earlier Syrian records. [9] Field trials using replica furnaces confirmed that this furnace type uses a wind-based air-supply principle that is distinct from either forced or natural draught, and show also that they are capable of producing high-carbon steel. [10]

Wrought iron was used in the construction of monuments such as the iron pillar of Delhi, built in the third century AD during the Gupta Empire. The latter was built using a towering series of disc-shaped iron blooms. Similar to China, high-carbon steel was eventually used in India, although cast iron was not used for architecture until modern times. [11]

Early to Medieval Europe

A Catalan furnace, with tuyere and bellows on the right PSM V38 D175 A blomary fire.jpg
A Catalan furnace, with tuyere and bellows on the right

Early European bloomeries were relatively small, primarily due to the mechanical limits of human-powered bellows and the amount of force possible to apply with hand-driven sledge hammers. Those known archaeologically from the pre-Roman Iron Age tend to be in the 2 kg range, produced in low shaft furnaces. Roman-era production often used furnaces tall enough to create a natural draft effect (into the range of 200 cm tall), and increasing bloom sizes into the range of 10–15 kg. [12] Contemporary experimenters had routinely made blooms using Northern European-derived "short-shaft" furnaces with blown air supplies in the 5–10 kg range [13] The use of waterwheels, spreading around the turn of the first millennium and used to power more massive bellows, allowed the bloomery to become larger and hotter, with associated trip hammers allowing the consolidation forging of the larger blooms created. Progressively larger bloomeries were constructed in the late 14th century, with a capacity of about 15 kg on average, though exceptions did exist. European average bloom sizes quickly rose to 300 kg, where they levelled off until the demise of the bloomery.

As a bloomery's size is increased, the iron ore is exposed to burning charcoal for a longer time. When combined with the strong air blast required to penetrate the large ore and charcoal stack, this may cause part of the iron to melt and become saturated with carbon in the process, producing unforgeable pig iron, which requires oxidation to be reduced into cast iron, steel, and iron. This pig iron was considered a waste product detracting from the largest bloomeries' yield, and early blast furnaces, identical in construction, but dedicated to the production of molten iron, were not built until the 14th century. [14] [15]

Bloomery type furnaces typically produced a range of iron products from very low-carbon iron to steel containing around 0.2–1.5% carbon. The master smith had to select pieces of low-carbon iron, carburize them, and pattern-weld them together to make steel sheets. Even when applied to a noncarburized bloom, this pound, fold, and weld process resulted in a more homogeneous product and removed much of the slag. The process had to be repeated up to 15 times when high-quality steel was needed, as for a sword. The alternative was to carburize the surface of a finished product. Each welding's heat oxidises some carbon, so the master smith had to make sure enough carbon was in the starting mixture. [16] [17]

In England and Wales, despite the arrival of the blast furnace in the Weald in about 1491, bloomery forges, probably using waterpower for the hammer and the bellows, were operating in the West Midlands region beyond 1580. In Furness and Cumberland, they operated into the early 17th century and the last one in England (near Garstang) did not close until about 1770. [18] [19]

One of the oldest-known blast furnaces in Europe has been found in Lapphyttan in Sweden, carbon-14 dated to be from the 12th century. [20] [ full citation needed ] The oldest bloomery in Sweden, also found in the same area, has been carbon-14 dated to 700 BCE. [21]

Bloomeries survived in Spain and southern France as Catalan forges into the mid-19th century, [22] and in Austria as the Stückofen  [ fr ] to 1775.

The Americas

Iron smelting was unknown in pre-Columbian America.

Excavations at L'Anse aux Meadows, Newfoundland, have found considerable evidence for the processing of bog iron and the production of iron in a bloomery by the Norse. [23] The cluster of Viking Age (c.1000–1022 AD) at L'Anse aux Meadows are situated on a raised marine terrace, between a sedge peat bog and the ocean. Estimates from the smaller amount of slag recovered archaeologically suggest 15 kg of slag was produced during what appears to have been a single smelting attempt. By comparing the iron content of the primary bog iron ore found in the purpose built 'furnace hut' with the iron remaining in that slag, an estimated 3 kg iron bloom was produced. At a yield of at best 20% from what is a good iron rich ore, this suggests the workers processing the ore had not been particularly skilled. [23] This supports the idea that iron processing knowledge was widespread and not restricted to major centers of trade and commerce. [23] Archaeologists also found 98 nail, and importantly, ship rivet fragments, at the site as well as considerable evidence for woodworking – which points to boat or possibly ship repairs being undertaken at the site. [23] [24] (An important consideration remains that a potential 3 kg raw bloom most certainly does not make enough refined bar to manufacture the 3 kg of recovered nails and rivets.)

A view of the bloomeries (Catalan forges) at Mission San Juan Capistrano, the oldest (circa 1790s) existing facilities of their kind in California Mission San Juan Capistrano 4-5-05 100 6559.JPG
A view of the bloomeries (Catalan forges) at Mission San Juan Capistrano, the oldest (circa 1790s) existing facilities of their kind in California

In the Spanish colonization of the Americas, bloomeries or "Catalan forges" were part of "self-sufficiency" at some of the missions, encomiendas , and pueblos . As part of the Franciscan Spanish missions in Alta California, the "Catalan forges" at Mission San Juan Capistrano from the 1790s are the oldest existing facilities of their kind in the present day state of California. The bloomeries' sign proclaims the site as being "part of Orange County's first industrial complex".

The archaeology at Jamestown Virginia (circa 1610–1615[ citation needed ]) had recovered the remains of a simple short-shaft bloomery furnace, likely intended as yet another "resource test" like the one in Vinland much earlier. The English settlers of the Thirteen Colonies were prevented by law from manufacture; for a time, the British sought to situate most of the skilled artisanry at domestic locations. In fact, this was one of the problems that led to the revolution.[ citation needed ] The Falling Creek Ironworks was the first in the United States. The Neabsco Iron Works is an example of the early Virginian effort to form a workable American industry.

The earliest iron forge in colonial Pennsylvania was Thomas Rutter's bloomery near Pottstown, founded in 1716. [25] In the Adirondacks, New York, new bloomeries using the hot blast technique were built in the 19th century. [26]

See also

Related Research Articles

<span class="mw-page-title-main">Smelting</span> Use of heat and a reducing agent to extract metal from ore

Smelting is a process of applying heat and a chemical reducing agent to an ore to extract a desired base metal product. It is a form of extractive metallurgy that is used to obtain many metals such as iron, copper, silver, tin, lead and zinc. Smelting uses heat and a chemical reducing agent to decompose the ore, driving off other elements as gases or slag and leaving the metal behind. The reducing agent is commonly a fossil fuel source of carbon, such as carbon monoxide from incomplete combustion of coke—or, in earlier times, of charcoal. The oxygen in the ore binds to carbon at high temperatures as the chemical potential energy of the bonds in carbon dioxide is lower than that of the bonds in the ore.

<span class="mw-page-title-main">Pig iron</span> Iron alloy

Pig iron, also known as crude iron, is an intermediate good used by the iron industry in the production of steel. It is developed by smelting iron ore in a blast furnace. Pig iron has a high carbon content, typically 3.8–4.7%, along with silica and other constituents of dross, which makes it brittle and not useful directly as a material except for limited applications.

<span class="mw-page-title-main">Wrought iron</span> Iron alloy with a very low carbon content

Wrought iron is an iron alloy with a very low carbon content in contrast to that of cast iron. It is a semi-fused mass of iron with fibrous slag inclusions, which give it a wood-like "grain" that is visible when it is etched, rusted, or bent to failure. Wrought iron is tough, malleable, ductile, corrosion resistant, and easily forge welded, but is more difficult to weld electrically.

<span class="mw-page-title-main">Blacksmith</span> Person who creates wrought iron or steel products by forging, hammering, bending, and cutting

A blacksmith is a metalsmith who creates objects primarily from wrought iron or steel, but sometimes from other metals, by forging the metal, using tools to hammer, bend, and cut. Blacksmiths produce objects such as gates, grilles, railings, light fixtures, furniture, sculpture, tools, agricultural implements, decorative and religious items, cooking utensils, and weapons. There was an historical distinction between the heavy work of the blacksmith and the more delicate operation of a whitesmith, who usually worked in gold, silver, pewter, or the finishing steps of fine steel. The place where a blacksmith works is called variously a smithy, a forge or a blacksmith's shop.

<span class="mw-page-title-main">Blast furnace</span> Type of furnace used for smelting to produce industrial metals

A blast furnace is a type of metallurgical furnace used for smelting to produce industrial metals, generally pig iron, but also others such as lead or copper. Blast refers to the combustion air being supplied above atmospheric pressure.

<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">William Kelly (inventor)</span> American businessman

William Kelly, born in Pittsburgh, Pennsylvania, was an American inventor. He is credited with being one of the inventors of modern steel production, through the process of injecting air into molten iron, which he experimented with in the early 1850s. A similar process was discovered independently by Henry Bessemer and patented in 1855. Due to a financial panic in 1857, a company that had already licensed the Bessemer process was able to purchase Kelly's patents, and licensed both under a single scheme using the Bessemer name. Kelly's role in the invention of the process is much less known.

<span class="mw-page-title-main">Bog iron</span> Form of iron ore deposited in bogs

Bog iron is a form of impure iron deposit that develops in bogs or swamps by the chemical or biochemical oxidation of iron carried in solution. In general, bog ores consist primarily of iron oxyhydroxides, commonly goethite.

<span class="mw-page-title-main">Wealden iron industry</span>

The Wealden iron industry was located in the Weald of south-eastern England. It was formerly an important industry, producing a large proportion of the bar iron made in England in the 16th century and most British cannon until about 1770. Ironmaking in the Weald used ironstone from various clay beds, and was fuelled by charcoal made from trees in the heavily wooded landscape. The industry in the Weald declined when ironmaking began to be fuelled by coke made from coal, which does not occur accessibly in the area.

<span class="mw-page-title-main">Puddling (metallurgy)</span> Step in the manufacture of iron

Puddling is the process of converting pig iron to bar (wrought) iron in a coal fired reverberatory furnace. It was developed in England during the 1780s. The molten pig iron was stirred in a reverberatory furnace, in an oxidizing environment to burn the carbon, resulting in wrought iron. It was one of the most important processes for making the first appreciable volumes of valuable and useful bar iron without the use of charcoal. Eventually, the furnace would be used to make small quantities of specialty steels.

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

A finery forge is a forge used to produce wrought iron from pig iron by decarburization in a process called "fining" which involved liquifying cast iron in a fining hearth and removing carbon from the molten cast iron through oxidation. Finery forges were used as early as the 3rd century BC in China. The finery forge process was replaced by the puddling process and the roller mill, both developed by Henry Cort in 1783–4, but not becoming widespread until after 1800.

<span class="mw-page-title-main">Ironsand</span> A type of sand with heavy concentrations of iron

Ironsand, also known as iron-sand or iron sand, is a type of sand with heavy concentrations of iron. It is typically dark grey or blackish in colour.

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

A Walloon forge is a type of finery forge that decarbonizes pig iron into wrought iron.

<span class="mw-page-title-main">Ferrous metallurgy</span> Metallurgy of iron and its alloys

Ferrous metallurgy is the metallurgy of iron and its alloys. The earliest surviving prehistoric iron artifacts, from the 4th millennium BC in Egypt, were made from meteoritic iron-nickel. It is not known when or where the smelting of iron from ores began, but by the end of the 2nd millennium BC iron was being produced from iron ores in the region from Greece to India, and sub-Saharan Africa. The use of wrought iron was known by the 1st millennium BC, and its spread defined the Iron Age. During the medieval period, smiths in Europe found a way of producing wrought iron from cast iron, in this context known as pig iron, using finery forges. All these processes required charcoal as fuel.

<span class="mw-page-title-main">Cornwall Iron Furnace</span> Historic district in Pennsylvania, United States

Cornwall Iron Furnace is a designated National Historic Landmark that is administered by the Pennsylvania Historical and Museum Commission in Cornwall, Lebanon County, Pennsylvania in the United States. The furnace was a leading Pennsylvania iron producer from 1742 until it was shut down in 1883. The furnaces, support buildings and surrounding community have been preserved as a historical site and museum, providing a glimpse into Lebanon County's industrial past. The site is the only intact charcoal-burning iron blast furnace in its original plantation in the western hemisphere. Established by Peter Grubb in 1742, Cornwall Furnace was operated during the Revolution by his sons Curtis and Peter Jr. who were major arms providers to George Washington. Robert Coleman acquired Cornwall Furnace after the Revolution and became Pennsylvania's first millionaire. Ownership of the furnace and its surroundings was transferred to the Commonwealth of Pennsylvania in 1932.

<span class="mw-page-title-main">Iron metallurgy in Africa</span> Iron production in Africa

Iron metallurgy in Africa developed within Africa; though initially assumed to be of external origin, this assumption has been rendered untenable; archaeological evidence has increasingly supported an indigenous origin. Some recent studies date the inception of iron metallurgy in Africa between 3000 BCE and 2500 BCE. Archaeometallurgical scientific knowledge and technological development originated in numerous centers of Africa; the centers of origin were located in West Africa, Central Africa, and East Africa; consequently, as these origin centers are located within inner Africa, these archaeometallurgical developments are thus native African technologies.

<span class="mw-page-title-main">Archaeometallurgical slag</span> Artefact of ancient iron production

Archaeometallurgical slag is slag discovered and studied in the context of archaeology. Slag, the byproduct of iron-working processes such as smelting or smithing, is left at the iron-working site rather than being moved away with the product. As it weathers well, it is readily available for study. The size, shape, chemical composition and microstructure of slag are determined by features of the iron-working processes used at the time of its formation.

Experimental archaeometallurgy is a subset of experimental archaeology that specifically involves past metallurgical processes most commonly involving the replication of copper and iron objects as well as testing the methodology behind the production of ancient metals and metal objects. Metals and elements used primarily as alloying materials, such as tin, lead, and arsenic, are also a part of experimental research.

Kemondo Iron Age Sites or KM2 and KM3 are Early Iron Age complex industrial archaeological sites in Kemondo ward, Bukoba Rural District, Kagera Region, Tanzania, excavated by a team led by archaeologist Peter Schmid in the late 1970s and 1980s. The excavations aimed at better understanding the iron smelting process and its ritual aspects in East Africa. At the KM2 and KM3 sites, Schmidt tested the hypothesis that the high combustion temperature of furnaces, discovered to be between 1,350–1,400 °C (2,460–2,550 °F), was caused by the preheating of air blasts. Preheating has been suggested to be a distinct feature of African Early Iron Age smelting techniques by ethnographic observations of the Haya people of northwestern Tanzania.

<span class="mw-page-title-main">Metallurgical furnace</span> Device used to heat, melt, or otherwise process metals

A metallurgical furnace, often simply referred to as a furnace when the context is known, is an industrial furnace used to heat, melt, or otherwise process metals. Furnaces have been a central piece of equipment throughout the history of metallurgy; processing metals with heat is even its own engineering specialty known as pyrometallurgy.

References

Bloomery iron furnace along Bloomery Pike (West Virginia Route 127) near Bloomery, West Virginia, United States. Bloomery Iron Furnace Bloomery WV 2013 09 03 06.jpg
Bloomery iron furnace along Bloomery Pike (West Virginia Route 127) near Bloomery, West Virginia, United States.
  1. Thornton, Jonathan; Williams, Skip; Shugar, Aaron. "The Rockbridge Bloomery – Reports: Smelting Enriched Bog Ore in a Low Shaft Bloomery". The Smelter's Art: Experimental Iron Production at The Rockbridge Bloomery. Washington and Lee University. Retrieved 21 August 2023.
  2. "The History of Forging - Now and Then". Canton Drop Forge.
  3. "The Earliest Use of Iron in China" by Donald B. Wagner in Metals in Antiquity, by Suzanne M. M. Young, A. Mark Pollard, Paul Budd and Robert A. Ixer (BAR International Series, 792), Oxford: Archaeopress, 1999, pp. 1–9.
  4. Cline, W. W. (1937) Mining and Metallurgy in Negro Africa.Meroe become the iron smelting center of East Africa Menasha, Wisconsin: George Banta
  5. Eggert, Manfred (2014). "Early iron in West and Central Africa". In Breunig, P (ed.). Nok: African Sculpture in Archaeological Context. Frankfurt, Germany: Africa Magna. pp. 51–59.
  6. 1 2 3 Eggert, Manfred (2014). "Early iron in West and Central Africa". In Breunig, P (ed.). Nok: African Sculpture in Archaeological Context. Frankfurt, Germany: Africa Magna. pp. 53–54. ISBN   9783937248462.
  7. Eze–Uzomaka, Pamela. "Iron and its influence on the prehistoric site of Lejja". Academia.edu. University of Nigeria, Nsukka, Nigeria. Retrieved 12 December 2014.
  8. Holl, Augustin F. C. (6 November 2009). "Early West African Metallurgies: New Data and Old Orthodoxy". Journal of World Prehistory. 22 (4): 415–438. doi:10.1007/s10963-009-9030-6. S2CID   161611760.
  9. Hoyland, Robert G. (2006). "Medieval Islamic swords and swordmaking: Kindi's treatise 'On swords and their kinds' (edition, translation, and commentary)" . Retrieved 9 October 2022 via Google Scholar.
  10. Juleff, Gill (January 1996). "An ancient wind-powered iron smelting technology in Sri Lanka". Nature. 379 (6560): 60–63. doi:10.1038/379060a0. ISSN   1476-4687. S2CID   205026185.
  11. Ranganathan, Srinivasa; Srinivasan, Sharada (1997). "Metallurgical Heritage of India". Golden Jubilee Souvenir, Indian Institute of Science. University of Illinois, Department of Materials Science and Engineering. pp. 29–36. Retrieved 30 October 2019.
  12. Radomir Pliener, Iron in Archaeology - the European Bloomery Smelters, chapter XII, 2000
  13. Darrell Markewitz, "If you don't get any IRON - Towards an Effective Method for Small Iron Smelting Furnaces", EXARC Journal 2012-1
  14. Douglas Alan Fisher, The Epic of Steel, Harper & Row 1963, p. 26–29
  15. Blast furnace, theory and practice, American Institute of Mining, Metallurgical, and Petroleum Engineers, Gordon and Breach Science 1969, pp. 4–5
  16. "Some Aspects of the Metallurgy and Production of European Armor". Archived from the original on 22 April 2002. Retrieved 14 July 2012.
  17. Alan R. Williams, Methods of manufacture of swords in medieval Europe, Gladius 1977, p. 70–77
  18. H. R. Schubert, History of the British Iron and Steel Industry (1957).
  19. R. F. Tylecote, History of Metallurgy (1991).
  20. "The blast furnace in earlier times". Jernkontoret. 21 December 2018. Archived from the original on 26 December 2023.
  21. Magnusson, G. (2015). Järnet och Sveriges medeltida modernisering. Jernkontoret, Stockholm.
  22. "Bloomery process". Encyclopædia Britannica . Retrieved 15 July 2017. The final version of this kind of bloomery hearth survived in Spain until the 19th century.
  23. 1 2 3 4 Bowles, G.; Bowker, R.; Samsonoff, N. (2011). "Viking expansion and the search for bog iron". Platforum. 12: 25–37.
  24. Lewis-Simpson, Shannon (2000). Vinland Revisited: The Norse World at the Turn of the First Millennium. St. John's, Newfoundland: Historic Sites Association of Newfoundland and Labrador. ISBN   0-919735-07-X.
  25. Bolles, Albert Sidney (1878). Industrial history of the United States, from the earliest settlements to the present time: being a complete survey of American industries, embracing agriculture and horticulture; including the cultivation of cotton, tobacco, wheat; the raising of horses, neat-cattle, etc.; all the important manufactures, shipping and fisheries, railroads, mines and mining, and oil; also a history of the coal-miners and the Molly Maguires; banks, insurance, and commerce; trade-unions, strikes, and eight-hour movement; together with a description of Canadian industries. Norwich, Connecticut: Henry Bill Publishing Company. p. 193.
  26. Gordon C. Pollard, "Experimentation in 19th century bloomery production: evidence from the Adirondacks of New York", Historical Metallurgy 32(1) (1998), 33–40.