Pottery

We're all about Art! 
Let's get potted on Pottery.

 

NOTE: Use the slide out menu at left to link to and view the pages of the artists in the Pottery category

A History of Pottery
by Charles Platten Woodhouse.

Online Resources
The production of pottery is one of the most ancient arts. The oldest known body of pottery dates from the Jomon period (from about 10,500 to 400 BC) in Japan; and even the earliest Jomon ceramics exhibit a unique sophistication of technique and design.

Excavations in the Near East have revealed that primitive fired-clay vessels were made there more than 8,000 years ago. Potters were working in Iran by about 5500 BC, and earthenware was probably being produced even earlier on the Iranian high plateau. Chinese potters had developed characteristic techniques by about 5000 BC. In the New World many pre-Columbian American cultures developed highly artistic pottery traditions.

After general sections on basic pottery types and decorating techniques this article focuses on the development of Western pottery since the beginning of the Renaissance. For detailed treatment of ancient Western and non-Western pottery, see Chinese art and architecture; Egypt, ancient; Greek art; Islamic art and architecture; Japanese art and architecture; Korean art; Mesopotamia; Minoan art; Persian art and architecture; pre-Columbian art and architecture.

TYPES OF WARES
Pottery comprises three distinctive types of wares. The first type, earthenware, has been made following virtually the same techniques since ancient times; only in the modern era has mass production brought changes in materials and methods. Earthenware is basically composed of clay--often blended clays--and baked hard, the degree of hardness depending on the intensity of the heat. After the invention of glazing, earthenwares were coated with glaze to render them waterproof; sometimes glaze was applied decoratively. It was found that, when fired at great heat, the clay body became nonporous. This second type of pottery, called stoneware, came to be preferred for domestic use.
The third type of pottery is a Chinese invention that appeared when feldspathic material in a fusible state was incorporated in a stoneware composition. The ancient Chinese called decayed feldspar kaolin (meaning "high place," where it was originally found); this substance is known in the West as china clay. Petuntse, or china stone, a less decayed, more fusible feldspathic material, was also used in Chinese porcelain; it forms a white cement that binds together the particles of less fusible kaolin. Significantly, the Chinese have never felt that high-quality porcelain must be either translucent or white. Two types of porcelain evolved: "true" porcelain, consisting of a kaolin hard-paste body, extremely glassy and smooth, produced by high temperature firing, and soft porcelain, invariably translucent and lead glazed, produced from a composition of ground glass and other ingredients including white clay and fired at a low temperature. The latter was widely produced by 18th-century European potters.
It is believed that porcelain was first made by Chinese potters toward the end of the Han period (206 BC-AD 220), when pottery generally became more refined in body, form, and decoration. The Chinese made early vitreous wares (protoporcelain) before they developed their white vitreous ware (true porcelain) that was later so much admired by Europeans.

Regardless of time or place, basic pottery techniques have varied little except in ancient America, where the potter's wheel was unknown. Among the requisites of success are correct composition of the clay body by using balanced materials; skill in shaping the wet clay on the wheel or pressing it into molds; and, most important, firing at the correct temperature. The last operation depends vitally on the experience, judgment, and technical skill of the potter.

DECORATING TECHNIQUES
In the course of their long history potters have used many decorating techniques. Among the earliest, impressing and incising of wares are still favored. Ancient potters in Egypt, Mesopotamia, Greece, northern India, and the high regions of Central Asia (where primitive terra-cotta figures associated with religious cults were produced) frequently decorated wares with impressed or incised designs. A notable incising technique developed more recently was that of Korean potters working in the Koryo period (918-1392). These artisans began by ornamenting their celadon wares with delicately incised and impressed patterns and later developed elaborate inlaying by filling incised lines with colored slip (semiliquid clay). Black and white slip was used most effectively for inlaying colored porcelains. Decoration of this sort generally depends more on the skill of the artisan than on the complexity of the tools being used.

An especially popular type of decoration involved the sgraffito, or "scratched," technique used by Italian potters before the 15th century. This technique, which is thought to have reached Italy from the Near East, was probably derived from China, where it was first used during the Song (Sung) dynasty (960-1279). By the 16th century Italian potters working mainly in Padua and Bologna had developed great skill in sgraffito, which entailed the incising of designs on red or buff earthenware that had been coated with ordinary transparent lead glaze, usually toned yellow or, sometimes, brown, copper, or green. After firing, the wares were dipped into white clay slip so that a dark pattern could be cut on the surface. By cutting through the white slip, the artist produced a design on the exposed red or buff body. Pigments were also sometimes applied. After a further coating of lead glaze the ware was fired a second time.

A sound knowledge of glazes--both utilitarian and decorative--is vital to the potter. The origin of glazes and glazing techniques is unknown, but the fine lustrous glazes developed in China surely began with a simple glaze that served to cover earthenware and render it watertight. Chinese potters used two kinds of glazes, one composed basically of feldspar, and another produced by fusing silica of quartz or sand by means of a flux, generally of lead oxide.
Chinese potters regarded glazes and glazing techniques as having prime importance; under the Han emperors they made great efforts to improve this technology. The use of lead glaze increased, and wood ash was incorporated to impart a dullish brown or gray green coloring, somewhat blotchy and occasionally iridescent. These effects were entirely natural, as no coloring matter was added to the composition. Glazing techniques were modified under successive dynasties. Colored glazes were developed and used to brilliant effect by Tang (T'ang) and Song potters, and a great diversity of brightly hued wares appeared over the centuries.

Many connoisseurs feel that the pure white porcelain, called blanc de chine, which first appeared during the Ming dynasty, is the most serenely beautiful of all Chinese ceramics. Dehua (Te-hua) potters in Fujian (Fukien) province, working during the 17th century, produced their blanc de chine masterpieces in the purest white porcelain coated with a thick white glaze.

Salt glaze, used by English potters during the early 1700s, may well have been known to the Chinese but was not used by them. Near Eastern potters glazed wares in ancient times. Potters in Mesopotamia and Iran commonly used an alkaline glaze made of quartz mixed with sodium and potassium. An admixture of colored metallic oxides, mostly lead, was introduced later.

Painting on pottery and porcelain became richly colorful in many regions and periods. Decorative brush painting directly on the baked clay reached its zenith in China during the Ming dynasty (1368-1644), whose artists were highly skilled at painting in fired colors. For a long period Chinese ceramic artists had used only black or brown pigment to decorate wares that were then covered with clear glaze. It is believed that the appearance in China of 13th-century brush-decorated wares from Persia sparked a change. These works, painted in blue cobalt under the glaze, inspired the brushwork of the Chinese and the resulting so-called blue-and-white style.

Ming artists also excelled in painting over the glaze, using brilliant enamel colors. The overglaze technique, which evolved over two centuries, demanded correct preparation of the enamels, skill in application, and the proper (low) firing temperature. The overglaze enamel decorations executed during the reign of Chenghua (1465-87), which were never surpassed in China, incorporated flowers, foliage, and figure subjects against backgrounds of arabesques and scrollwork. Designs enclosed within dark blue outlines were filled in with brilliant color. Enamel decoration of superb quality was also executed in Japan during the Edo period (1615-1868) by celebrated artists and potters of the caliber of Kenzan, Kakiemon, and Ninsei.

In the ancient Aegean the potter's art developed continuously from the Neolithic period and through the periods of the Minoan and Mycenaean civilizations, culminating, in ancient Greece, in a unique type of painted pottery, which reached its height between the 6th and 4th centuries BC. The finest Greek pottery, especially Attic vases, was exquisitely proportioned and often decorated with finely painted relief work. Unlike artisans in Egypt, Mesopotamia, and Persia, the Attic potters did not apply heavy glaze to their wares. The unique gloss commonly seen on Attic pottery and similar wares made elsewhere in Greece still baffles those who have tried to determine its formula and method of application. Neither a glaze nor a varnish, it is more marked on some areas, such as those painted black, than on others. Some experts conjecture that it may be attributed to illite or a similar clay mineral in a weak solution that was thinly applied to the surface of wares or mixed into the black "paint" used by the artists.

In the Islamic world ceramic decorative art flowered with the creation of a great diversity of painted wares. Painted luster decoration on pottery originated in Mesopotamia and spread to ancient Egypt; later, under Islam in Persia, this type of decoration on white-glazed wares became incredibly brilliant. Islamic luster-painted wares were later imitated by Italian potters during the Renaissance.

MAJOR TRADITIONS IN THE WEST
After the fall of the ancient Roman Empire potters in Europe produced little other than repetitive utilitarian wares until the end of the Middle Ages.

Earthenware
A distinctive type of earthenware known as majolica, which was derived from Chinese porcelain, appeared in Italy during the last quarter of the 14th century. It is now believed that this type of painted earthenware was inspired by the Hispano-Moresque luster-decorated ware of Spanish origin introduced to Italy by Majorcan seagoing traders.
Majolica ware, whether thrown on the wheel or pressed into molds, was fired once to obtain a brown or buff body, then dipped in glaze composed of lead and tin oxide with a silicate of potash. The opaque glaze presented a surface that was suitable to receive decoration. A second firing after decoration fixed the white glaze to the body and the pigments to the glaze, so that the colors became permanently preserved. Frequently, the beauty of these wares was increased by dipping them in a translucent lead glaze composed of oxide of lead mixed with sand, potash, and salt. When certain luster pigments and enamels were used in all-over painting, wares had to be specially fired at low temperature. Application of metallic luster pigments required great skill because these colors were extremely volatile and needed special handling.

Luca della Robbia (see della Robbia, family) did not, as has been held, invent the enamel tin-glazing process; nevertheless, his work raised majolica production from a craft to high art in Italy. Not only did he use blue and white enamels in decorative work, but, as a sculptor, he also used the majolica technique to add brilliance to the surface of his productions. By the beginning of the 15th century Italian potters had abandoned the old familiar processes, and a revolution in style and techniques was under way. The severe style as followed principally in the school of Tuscany continued to the end of the 15th century, but rules and principles slackened until the inclusion of human figures in designs, previously frowned upon, was accepted. At the end of the 15th century Faenza became the thriving center of a reinvigorated pottery industry in Italy. A new, rich decorative style, known as istoriato, fired the imagination of potters, reaching its zenith in the workshops of Urbino.

In early 17th-century England attractive slipwares were produced, including the slip-decorated earthenware that was a speciality of the Toft family of potters. A kind of tin-glazed earthenware was also produced in the Netherlands, principally at Delft, beginning in the mid-17th century. Termed delftware, it was among the first European wares to be decorated with motifs inspired by Chinese and Japanese models.

Continental Porcelains
Eventually, European potters, who much admired the porcelain of the Far East, attempted to imitate it, but the formula remained elusive. Francesco de Medici, grand duke of Tuscany, produced an inferior type of soft-paste porcelain in his Florence workshop during the 16th century. In March 1709, Augustus II of Saxony announced that his ceramist Johann Bottger (1682-1719) had discovered how to make porcelain. The first European royal porcelain manufactory was consequently established at Meissen (see Meissen ware) near Dresden, Germany. Throughout the century following the discovery of the porcelain formula--when, despite the utmost precautions at Meissen, the secret leaked out--many rival factories were set up in Europe. Germany, Austria, Italy, France, and England soon had factories engaged in the production of wares much like those of Meissen.

Porcelain figures were first produced in Meissen as table ornaments; the earliest examples were formed as part of sweetmeat dishes. Many splendid wares issued from the royal factory, but none were more admired than the finely modeled and decorated porcelain figures imitated by almost every German, Austrian, Italian, and English factory of note. Widespread interest in figures of both pottery and porcelain has continued to the present. Johann Joachim Kandler (1706-75), a master modeler, was the most notable of the artisans engaged in this work at Meissen and rivaled the famous Franz Anton Bustelli (1723-63) of Nymphenburg (see Nymphenburg ware).

The methods used to produce porcelain figures as developed by Kandler imparted a new dimension to the art. German porcelain figures were usually produced from molds, which, in turn, were cast from an original master model made of wax, clay, or, occasionally, wood. The use of molds facilitated unlimited reproduction. Because the figures shrank during firing, allowances had to be made in their sizes; they were also provided with a small venthole in the back or base to permit excess heated air to escape. Because different factories placed these holes differently, their positions help determine the provenance and authenticity of given pieces. When considerable undercutting was necessary, porcelain figures were usually made in sections, using separate molds. Portions of elaborate groups and single figures were later joined by a specially trained assembler (known as a "repairer") who usually worked from a master model.

Europe's second hard-paste porcelain factory began operations at Vienna in 1717. In the late 1700s at the royal Sevres (see Sevres ware) factory in France, potters experimented until they developed a remarkably white, finely textured body. Sevres wares were painted in unique colors that no other European factory could duplicate. The bleu de roi and rose Pompadour of Sevres wares captivated all Europe and, with the products of Meissen and Vienna, inspired English potters.

English Wares
The finest English porcelain--both soft- and hard-paste--was made between about 1745 and 1775. The first English porcelain was probably produced at Chelsea (see Chelsea ware) under Charles Gouyn, but his successor Nicholas Sprimont, a Flemish silversmith who took over management in 1750, was responsible for the high-quality wares, especially the superb figures, for which the factory became famous. Factories at Worcester (see Worcester ware), Bow, and Derby also produced wares that rival those of the Continent.

Led by the ambitious, energetic, and enterprising Josiah Wedgwood and his successors at the Etruria factory, English potters in the late 18th and early 19th centuries became resourceful and inventive. Wedgwood's contributions consisted mainly of a much improved creamware, his celebrated jasperware, so-called black basalt, and a series of fine figures created by famous modelers and artists. After Wedgwood, other potters of the first half of the 19th century developed a number of new wares. Of these, Parian ware was the most outstanding and commercially successful.

The name of this ware was derived from Paros, the Greek island from which sculptors in ancient times obtained the creamy or ivory-tinted marble that Parian ware resembled. The first examples of this new product, described as "statuary porcelain," issued from Copeland and Garret's factory in 1842 and were immediately acclaimed. Two varieties of Parian ware were produced: statuary parian, used in the making of figures and reproductions of sculpture, and hard-paste, or standard, parian, from which hollowware was made. Statuary parian, incorporating a glassy frit, is classified as soft porcelain. Standard parian, with a greater proportion of feldspar in the composition but no frit, is hard porcelain. Early parian statuary was ivory-tinted due to the presence of iron in the feldspar devoid of iron silicate. Suitable deposits were eventually located in Sweden and Ireland. Both English and American potters either obtained details of the original formula or worked out their own, and the resulting production of Parian wares on both sides of the Atlantic was enormous.

Among the most beautiful and successful wares invented by 19th-century potters were those decorated in what came to be known in England as pate-sur-pate, a paste-on-paste technique devised sometime after 1870 by Marc-Louis Solon (1835-1913) of Minton's in England. Pate-sur-pate, involving both modeling and painting techniques, was stained Parian ware decorated with reliefs in translucent tinted or white slip, the colors being laid one upon the other. Solon was inspired by a Chinese celadon case decorated with embossed flowers that he had admired in the museum at Sevres, where he worked for a time. At first his slip painting on biscuit porcelain simply peeled off; he was successful, however, when he applied layers of slip to a damp surface. Minton wares decorated with pate-sur-pate became the most costly and coveted ceramic ornaments produced in England in the last quarter of the 19th century. Only a few English potters mastered Solon's complex technique, although the work of his pupil, Alboin Birks, rivaled that of the master.

20th-Century Developments
By the late 19th century, with the development of machinery and the introduction of new technologies, the age of mass production dawned and the potter's art consequently suffered. Western ceramic wares declined markedly in quality of materials and decoration. Florid designs, gaudy coloring, and inartistic shapes became fashionable, and the resulting decadence continued into the 20th century. Not until the 1930s were signs of revival in the form and decoration of ceramics discernible, principally in the productions of artist-potters who were active in Western Europe and the United States. Many of these artist-potters arrived at their innovations by way of continuous experiment with materials and techniques. Others sought inspiration from primitive types of Japanese pottery or in the forms of ancient American Indian traditions. Since the end of World War II the design and decoration of ceramics in both Europe and the United States, especially ornamental wares, has been largely influenced by individual artist-artisans. Commercial products, such as tablewares, have tended to reflect the styles and patterns developed by these potters, whose work has often shown striking originality.


Two Medieval London-type jugs from Longmarket
By John Cotter 

Two of the most significant medieval pots found on the Longmarket site are the subject of this note. Both are of considerable interest and beauty and although broken they are remarkable for their state of completeness and preservation. The reason for their excellent condition is that both vessels were thrown to the bottom of two separate cess-pits or latrines where they lay undisturbed for the next seven centuries. 

Both are glazed wheel-thrown jugs in what is known as London-type ware. The kilns for this `type' have never been discovered but there is little doubt that it was made somewhere in the London area, at a time when London potters were heavily influenced by French pottery styles. They have a fine sandy fabric, dull orange in colour with a pale grey core. A creamy white slip or liquid clay has been smeared all over the outside of both vessels including the handles and spout and partly inside the neck, but it stops a few centimetres above the base where it was deliberately cleaned away with a knife or perhaps a cloth or leather pad. 

The smaller of the two (on the left in the photo) is a baluster jug decorated in the Rouen style and dates to c. 1200-1250 A.D. This style consists of contrasting creamy white and dark red painted zones defined by applied strips in white clay with further details, such as studs and pellets, also executed in white clay (the excavators christened it the `rhubarb pot' due to this rhubarb- and-custard colour scheme). This is a variation of the chevron design which was popular both on London and Rouen (Normandy) jugs. On some jugs, such as this one, the design seems to echo motifs found in Norman architecture and perhaps also the wrought ironwork seen, for example, on church doors (hence the `studs' and `nails'). The handle has been plugged-in through the body wall in typical London fashion while a separate bridge spout has been applied to the front of the jug. A clear pitted glaze covers most of the upper two thirds of the body but the neck is only partially glazed while the handle and adjacent area are glaze-free. Glaze dribbles on the underside suggest that like most London jugs this example was fired upside-down. 

The second vessel (on the right in the photo) is a large rounded jug decorated in the North French style. This was more long-lived than the Rouen style being present on London excavations as early as c. 1200 and possibly still in circulation as late as c. 1340, albeit on a much reduced scale. Other imported and local wares found in the same cess-pit suggest however that this particular jug was discarded around 1275-1300 A.D. 

On this jug the plain paired vertical strips are of applied red-brown body clay which contrasts with the overall white slip background. The sinuous strips with their intervening smeared scales and the vertical diamond-rouletted strips are all in applied white clay. Like the Rouen-style jug glaze coverage is confined to the upper two thirds of the body and avoids the handle area. The glaze itself is pitted and mottled green; its uneven and patchy application over a background of contrasting red and white, plain and decorated details, gives the vessel a typical medieval exuberance which many people find charming, even beautiful, but which others find garish or repulsive. 

As with the first jug the handle on this one has also been plugged-in, it is also slightly facetted and at the top are two `ears' of applied clay - a typical North French feature. The spout is unfortunately missing but probably resembled that of the first jug. Under the base there are some large glaze splashes with a contact scar from the rim of a similar jug stacked upon it in the kiln, both, as usual, fired upside down. 

The wood-lined cess-pit which produced the Rouen-style jug lay in the back yard of a property owned around 1200 by a certain William Malemie, about whom nothing further is known. One is naturally tempted to associate the jug with William and certainly the dating does not preclude this, particularly if he lived for some years after 1200. There is some archaeological evidence however that William Malemie's back yard may have been encroached upon by his influential neighbour Terric the Goldsmith, or Terric's sons, early in the thirteenth century, but eventually the original Malemie boundary was re-established. Probably we will never know for certain to whom the jug belonged but in archaeological terms it is still quite an achievement to narrow it down to just one or two known families. 

Both the Malemie and Terric families undoubtedly belonged to Canterbury's wealthy artisan or merchant class, a fact reflected in their choice of quality French-style tablewares from London in preference to the less sophisticated wares of the Tyler Hill kilns located just outside Canterbury. The same is true for the unknown owner of the slightly later North French-style London jug. His stone-lined cess-tank yielded jugs from Tyler Hill, London, and some particularly fine whiteware jugs from the Saintonge area of south-west France. 

Neither of the Longmarket jugs is exactly paralleled in the published corpus of jugs in London-type ware, the forms and many of the decorative elements are, but the precise combinations as seen here appear to be entirely new. They are thus an important addition to the study of this ware as well as the study of medieval pottery as a whole.

Features On Site
Art Museums
Featured Artist

Medieval woman
throwing pottery on a wheel. English Book Plate.
Featured Artist

Two Medieval London-type jugs from Longmarket
pot on left c. 1200-1250
pot on right c. 1275-1300 A.D. See story below.
Featured Artist

Medieval Pottery excavated from a Carmelite friary at Esslingen Germany
The building of an extension to the Esslingen Technical College on a plot of land directly outside the medieval city centre on the Kiesstraße, known since the Middle Ages as "Auf dem Kies" (roughly translatable as "on the banks"), threatened an area known to have been the site of a Carmelite Friary, founded at the end of the 13th century. Con'd Below.
Featured Artist

Medieval Pottery excavated from a Carmelite friary at Esslingen Germany
The efforts of the Baden-Württemberg State Monuments Office to preserve this area as an archaeological "research park" for future generations were ultimately unsuccessful, but an agreement was reached which allowed the archaeologists time in which to rescue as much as possible. Accordingly, large scale excavations were undertaken in 1991-92. Con'd Below.
Featured Artist
Medieval Pottery excavated from a Carmelite friary at Esslingen Germany
It is hardly surprising that there were several unique vessels amongst the huge corpus of pottery rescued. Perhaps the most exciting find was a tripartite vessel in the "Swabian" red-painted Fineware made in Buoch, which was the most popular fineware in the mid-Neckar area in the 13th - 15th century. Although fragments of such vessels had been found elsewhere, we had never before been able to work out what kind of strange pot they belonged to. The actual function of this tree-bodied (each "globe" was connected with the others) vessel, with one spout and one upper opening, is however still something of a mystery. Con'd Below.
Featured Artist

Medieval Pottery excavated from a Carmelite friary at Esslingen Germany
The soft, moist sediments filling the stream also preserved a wealth of artefacts, many of wich were completely undamaged. Wooden lathe-turned plates, cooped beakers and carved spoons, wooden flutes, leather shoes and even a wicker fish-trap were all well preserved. 
Large amounts of pottery vessels were also preserved, representing a wide spectrum (cooking pots with lids, jugs, beakers, miniature vesels) of the typical pottery types found in the Esslingen area in the later medieval period. Con'd Below.
Featured Artist

Medieval Pottery excavated from a Carmelite friary at Esslingen Germany
The majority of the pottery vessels consisted of several dozen small unglazed bowls in a coarse grey ware. Many of these seemed to have been unused and were found stacked inside one another, which suggests that a large supply had been thrown out in one go at some time in the 15th century (after the fire?). Perhaps they were used for sharing out meals for the poor. End
Featured Artist

Examples of Pagan period Saxon pottery. 
c. 10th & 11th century

These Urns were used to store the ashes of the dead, which were then buried. The British Isles has large and diverse areas of clay that are suitable to make pottery. Broadly speaking, the area diagonally south of York and down to Cheshire has in various places clay deposits that are close to the surface. This enabled people from much, much earlier times and up to the Viking period to dig clay for pottery without having to go too deep. Clay is very heavy, and difficult to dig out. The rest of Britain by and large had to make do with 'costly' imports that could have come from a few miles down the road, or possibly several days travel away. Their only other alternatives were wooden vessels, or in other more remote areas, 'soft' soap-stone containers. Con'd Below.
Featured Artist

Examples of Pagan period Saxon pottery. 
c. 10th & 11th century

The Urns in the back row were used to store the ashes of the dead, which were then buried.Pottery was a very important method of producing cheap cooking pots, bowls, cups, lamps, bottles, jugs, etc.. It was also used for loom-weights, crucibles and moulds. In early pagan Anglo-Saxon times pottery 'urns' were used to hold ashes of people who had died and been cremated. These were then often buried in small 'barrows'. Many of these cremation urns were highly decorated. The vast majority of the early pottery though was simply made, probably within the village or on the farm, using methods such as coiling or making thumb-pots. Later on, as shown by excavated examples, there were specialist potters who made wheel thrown pottery in towns. This was then sold by the potter, or possibly by travelling merchants in the markets, although some pottery would still be home produced. Con'd Below.
Featured Artist

Examples of Pagan period Saxon and Viking pottery.
c. 10th & 11th century.

The pots were used for a variety of purposes, some for storage, some for cooking and some for eating and drinking from. Bowls would have been used for storage as well as cooking, eating and serving. Cups were generally in the form of handle-less beakers. A potter's tools were fairly simple. An animal rib or flat piece of wood for shaping the pot when throwing, knives for trimming, antler tines for piercing for spouts and bungs, perhaps a number of sheep's tibiae and metapodials ( elements of the bones in the foot of the animal ), as templates for rim profiles. Some carved bone and antler stamps were used with rouletting wheels for decorating the pottery. Con'd Below.
Featured Artist

Examples of Pagan period Saxon and Viking pottery.
c. 10th & 11th century.

Evidence would suggest that after about 900AD the potter's wheel as we would recognise it came back into fashion. The type of potter's wheel probably varied, anything from a small turn-table ( slow wheel ) to a large kick wheel. Two kinds of fast wheel may have been used. The first and most likely type to have been used in the Saxon period, is basically a cartwheel mounted horizontally on a pivot, the wheel being rotated by hand or with a stick. The pot was thrown on a disc or small platform fixed to the centre or nave of the wheel. The other type consisted of a lower wheel turned with the foot and an upper wheel head for throwing the pot, the two wheels being connected by a series of struts. Con'd Below.
Featured Artist

Examples of Pagan period Saxon and Viking pottery.
c. 10th & 11th century.

To make clay good enough for a pot, the Saxon potter would have to put in some back-breaking graft. After unearthing a large amount of clay, he would take his raw material, steep it in water and then beat it, usually with a large wooden 'spatula' until it was well mixed, although some potters may have worked it by treading it with bare feet. He would then remove any large stones and gravel from it. Next, he would carefully mix sand, crushed shell, grass, or even crushed pottery from broken fired pots in with it to help bind it together. Then he would have to wage it ( knead it like bread ), to ensure it was thoroughly mixed. The clay at this point would have to be made pliant enough by the addition of water or be left to dry some more. The potter would then take a ball of this clay of the correct size and consistency for the item he was making. This clay was formed into a pot, mainly by building it up from layers of rings which are smoothed together by hand ( coiling ) or, by about 900AD, on a wheel. Other methods may have included paddle and anvil techniques, with a pebble and spatula, thumb pots and moulding over wooden moulds ( this method was often used to make crucibles). Con'd Below.
Featured Artist

Examples of Pagan period Saxon and Viking pottery.
c. 10th & 11th century.

The item would then be left to dry gently. Features such as handles and spouts were usually added to the vessel when it had dried to a 'leather' hardness, or was firm enough not to distort when being handled. The simplest, and commonest, form of spout is the pinched spout made by pulling out the rim from inside with one finger whilst supporting the rim in position on the outside with two fingers. In this case this was done when the pot was first made. Tubular spouts were made either by throwing a small cylindrical shape, or by moulding clay around a forefinger, stick or bone. This was then smoothed onto the outside of the vessel once a hole had been made. Handles would be made by throwing, pulling or rolling out, and also applied by smoothing onto the outside of the pot. At this stage the bottom of the pot might be trimmed with a knife to give the familiar 'saggy bottom'. The 'saggy bottom' was we believe better for cooking with, as it helped to even out the differences of temperature in a cooking fire, which could easily crack a pot. Floors of the period weren't very flat themselves, so rounded bottom pots really didn't matter. Con'd Below.
Featured Artist

Examples of Pagan period Saxon and Viking pottery.
c. 10th & 11th century.

The container could then be worked over with a damp cloth or wet hands, which brings the finest clay particles to the surface, giving a smooth finish. The inside of the pot could also be burnished with a smooth pebble or bone to smear the clay particles over each other producing a more water tight vessel. It could also be decorated by painting with a slip ( a creamy mixture of fine clay and water ) of a different colour to the body. Sometimes slip painting amounted simply to vertical stripes of slip, sometimes it took the form of scrolls and swirls. Glazes were almost universally lead based, giving a greeny yellow colour, although copper or iron could be added to change the colour or add speckles of a different colour. These were added to the pot after an initial firing. The glaze could have been applied as a dry powder, although most was applied as a water based paste. Liquid glazes could be applied to the leather hard pot with a brush or by hand smearing, which accounts for the uneven thickness of many of the glazes from this period. The pot could also be dipped in a bath of glaze. It was then left to finally dry before it was fired to make it hard. Con'd Below.
Featured Artist

Examples of Pagan period Saxon and Viking pottery.
c. 10th & 11th century.

In the early period the pots were fired in a covered fire pit called a clamp. This did not always reach a very high temperature so the pots often did not fire very well. The fire that was built over the pots excluded most of the oxygen which fired the pottery black or charcoal-grey. By the later period firing was done in a simple kiln which was easier to control, guaranteeing a better and more even firing. In order to make a kiln the potter dug two shallow pits, one of them with a semi-permanent wall of clay or stone ( sometimes insulated with earth or turf ) with a simple domed roof built over it, possibly just of turf, but sometimes of clay. ( Turf is fine for a single firing, but if it becomes too roasted, breaks down into sand and minerals which just don't hold together ). This one became the kiln and was joined to the other pit by a small opening. The pots were stacked in the kiln, generally upside-down, sometimes one inside another, whichever way they packed most tightly. The loading could be done through the top of the kiln before it was sealed or through the flue/door opening. Con'd Below.
Featured Artist

Examples of Pagan period Saxon and Viking pottery.
c. 10th & 11th century.

The kiln was then sealed with wet clay leaving just the opening between the pits and a small flue opening. Some kilns had a raised central floor on which more pots were stacked, which allowed the hot air to circulate around the pots better. A hot fire was then built in the second pit in front of the opening. The potter would keep adding fuel slowly until the temperature was high enough to fire the pots, gauging its 'readiness' by the degree of luminosity of the items which glow whilst being fired. When this temperature had been reached the potter let the kiln cool down ( sometimes for a whole day ) until it was cool enough to remove the pots. Most would be hard and ready for use although some would have cracked if the clay and sand or shell had not been correctly mixed. With maintenance, a kiln of this type might last from five to ten years. Some pots would have been almost black due to a process known as reduction. This happens when oxygen is excluded from the kiln by clamping off any airways, and leaving it for a period of time. End

 


Copyright © 2007 Rideau Valley Artists Website.
All rights reserved. Revised: July 02, 2007 .