20 The Organisation of Pottery Production

The technical features of the vessels are related to a particular social reality and the way the production is organised, so that the study of pottery technology allows us to approach the production structure. In this sense, one of the issues which have been fully addressed in ceramic studies is the organisation of production and its relationship with the social structure. An overview regarding some studies which consider this topic in ceramic analysis was made, for example, by W. Longacre (1999). In this sense, many scholars have tried to establish a number of markers defining types of pottery production organisation. However, these markers were often drawn from ethnographic analogies according to processual theoretical and conceptual frameworks.

The aim of this research strategy was to establish broad categories or “black boxes” by means of a number of indicators related to specific forms of organising pottery production. Ultimately, archaeologists can use these general models to interpret their archaeological record. Examples of such categories can be found in Rice (1984b), Sinopoli (1991) or García Rosselló (2008), although we prefer to emphasise here an example of one of these general models applied to the management of raw materials. In this case, five strategies of clay procurement (Bishop et al., 1982) are categorised according to indicators involving the variability degree of pottery pastes (Arnold, 2000):

1. Non-discriminatory strategy: Various clays are used without a clear preference for a particular type of material. In these cases, the quality of the raw material is unknown and possibly irrelevant for the potter. This type of behaviour often takes place in production processes in which potters test the several resources available in the territory or when dispersed habitats prevail. This strategy of raw material procurement promotes very heterogeneous ceramic assemblages characterised by a high degree of variability.

2. Discriminatory strategy: Potters prefer to exploit one specific clay source, the quality of the material is known and the ceramic pastes are quite homogeneous and less variable than those produced by means of non-discriminatory strategies. This strategy of raw material procurement can be associated with a settlement pattern in which both people and pottery production are more concentrated or centralised in specific areas. It can be also related, as occurs in Tikul (Yucatán, Mexico), to a management of the clay sources restricted to a few specialists who select, extract and prepare the clay before it is distributed among all the potters in the community (Arnold, 1999).

3. Specialised strategy: Different types of clays are exploited depending on their specific properties and according to the different types of vessels produced and their functions. The quality and physical properties of the material is well-known and the ceramic assemblage shows higher heterogeneity, although this variability can be organised according to typological and morphometric parameters.

4. Composite strategy: This strategy involves the mixture of clays from different sources or deposits. Although these mixtures can be made by means of a more or less well-established proportion, they usually promote the presence of very heterogeneous ceramic assemblages.

5. Importation of clays: In this case, the clays used in the pottery production come from sources distant from the place of manufacture. This strategy results in ceramic assemblages that are very heterogeneous compositionally; however, they can be organised according to the origin of the raw materials.

As can be seen, many processual archaeologists tried to establish evolutionary models by means of transcultural and universal correspondences in order to interpret the technological variability and change, as well as the features of the archaeological ceramics, in terms of organisation and structure of production. In these kinds of perspectives the variability and changes recorded in the pottery are seen as indicators of technological development. Furthermore, it is understood that technological progress itself determines the traits of the pottery, its manufacturing process and the organisation of production. According to this trend, technology causes changes in the society and determines how it is organised. Thus, the technology becomes the cause and the backbone of the forms of social organization (Colomer, 2005; García Rosselló, 2008; Harry and Bubemyre, 2002; Kingery, 1984; Rice, 1984b).

One of the main criticisms made of this theoretical framework relates to the way the relationships between these broad transcultural categories and the specific reality observed through the artefacts are established. Critics highlight the weakness of the links created between the manufacturing process and the organisation of production in evolutionary terms, as well as using western concepts such as specialisation, efficiency, scale of production, etc. Thus, P. Rice (1984a) pointed out that such generalisations of the patterns of change and continuity in the organisation of pottery production cannot be made. This is because different societies may exhibit different motivations in the way they develop ceramic production, even if they use the same technological choices. These kinds of interpretations are ethnocentric. In this sense, P. Rice (1991) has come to argue that the concept of specialisation and other similar categories used to interpret the ceramic record must be related, for instance, to specific forms of understanding gender relations based on western and modern preconceptions. In addition to these constraints, it must be noted that these interpretative models were often designed from ethnography through indicators of the organisation of production that are very difficult to identify in the archaeological record, since they are not based on materiality (García Rosselló, 2008).

In spite of the fact that these models are excessively reductionist and generalist they may allow us to undertake more or less general categorisations regarding the way production is organised. However, we cannot forget that pottery production is always embedded in political, social and economic dynamics that are specific to each society. It is within these particular historical and social processes in which a certain way of organising pottery production has meaning and makes sense. In consequence, it should be within this framework where, ultimately, technological facts and the social role of the practices related to ceramic production must be interpreted. In this sense, other studies (e.g., Courty and Roux, 1999; Roux, 2003; García Rosselló, 2008) propose that the adoption of a specific technology is strongly determined by the forms of organisation and the social needs of each group. Thus, this perspective goes beyond evolutionary viewpoints and does not minimise the importance of the social contexts in shaping the materiality and the organisation of production. In this case, the technology is not only a cause but also a consequence of a certain social order.

Some ethnographic studies (e.g., Arnold, 1984; Gosselain, 1994a, 1996, 2000; Livingstone-Smith, 2000; Sillar and Tite, 2000) pointed out the presence of production patterns that are shared at the group level related to the use of certain concepts, raw materials, recipes and forming methods as well as decorative and firing strategies. In short, it is documented the collective use of certain chaînes opératoires to achieve specific ceramic products as a result of the sociocultural interaction existing among potters. For example, there may be a common understanding regarding the concept of symmetry or the spatial organisation of the vessel that makes it possible to carry out certain designs on the ceramics (Arnold, 1984; Shepard, 1971). This interaction enables individuals to develop particular styles by means the repetition of specific designs in certain parts of the vessels according to their type, form, function, etc. (Jones, 2004; Sofaer, 2006).

This repetition of the characteristics of the pottery in time and space, as well as the development of certain technological traditions, originates in social interactions that are determined by the way in which the production is organised in each society. In this sense, rather than referring to a series of ideal patterns regarding the social structure of production, it may be more useful to approach the features of the vessels to determine the degree of interaction that existed between potters, as well as between them and the users in particular social contexts. There is no doubt that human practices and their products are strongly determined by the context, where they interact with other technologies that are specific to each culture. Thus, the intensity and the degree of variability or regularity of the technological actions and the end products can be only understood within certain social and symbolic contexts (Prieto, 1999).

Finally, some authors (e.g., Arnold, 1999; García Rosselló, 2008; Jones, 2004; Sigaut, 1994) support a halfway position between both trends. In this sense, it is argued that a vision in which pottery, technology and technological change are inactive is neither convenient nor acceptable. In this way, then, not all the features of pottery production are exclusively determined by the ideational sphere and specific forms of social organisation. Thus, it was already mentioned that the raw materials available in the environment and the material dimension of the artefacts (i.e., their physical properties) are also fully involved in and interact with the forms of organisation of pottery production, thus influencing the individuals’ lifestyle.

In this sense, the systematic rejection of any theoretical position can be harmful, since it would leave aside a large number of elements that should be considered in any explanation. Thus, greater dialogue and consensus can be achieved on the analysis of the organisation of production if environmental and functional parameters are combined with contextual and social factors (Livingstone-Smith, 2000). Furthermore, as already seen, social aspects cannot be separated from the physical dimension of the artefacts, since certain properties of the materials or the adoption of certain techniques may constrain or expand the development of specific social practices and forms of organisation. For example, although explained in simplistic terms, it can be said that the use of hand-made modelling techniques provide greater freedom to potters. These forming methods allow an individual to model a wide range of shapes that are not determined by external elements as happens, for instance, when the vessels are made using mould making techniques or the potter’s wheel. In this way, hand-made forming methods increase the number of opportunities for individuals to express their ideas, thus increasing their range of social action through materiality (Barley, 1994; Jones, 2004; Sigaut, 1994; Tite, 2008; West, 1992). Then, it is not surprising that, as happens among the Kabyle Berber ethnic group in the Maghreb (Balfet, 1984), hand-made modelling and manually created shapes are considered as a continuous process of expression and creation of forms.

Thus, the development of a particular material culture involves dynamic and multidimensional movements back and forth between technology, environment, materiality and the social and ideological context. This interaction leads to specific technological traditions and certain forms of organisation of the pottery production. Therefore, several theoretical trends can be combined in order to create an epistemological basis to address social complexity through the study of the material record. This approach can only be possible if Bourdieu’s theory of practice, the principles of the French School and the anthropology of techniques regarding the role of technology in culture, and, finally, a historical perspective are fully considered (Dietler and Herbich, 1998; Dobres and Hoffman, 1994; Pfaffenberger, 1992).

This strategy may help in the assessment of the role that pottery technology plays in specific social contexts and in the mutability of the structure. Also, it enables us to know which social phenomena and dynamics promote the production and reproduction of material culture (Dietler and Herbich, 1998). That is, we can approach the way in which ceramic technology is structured by social practices and how technology also simultaneously structures such practices. The first issue is directly related to the strategies used by individuals and people to communicate, interact and share their knowledge in specific social contexts in order to organise themselves to create certain ceramic assemblages. The second involves the role that materials, techniques and objects, as well as their features, play regarding the creation, consolidation and reproduction of certain social structures.

Social relations and the organisation of pottery production in past societies must be necessarily addressed through the indicators and connections readily visible in the material record and in particular contexts. Thus, the development of a proper theoretical and methodological framework is crucial in order to identify which features of the materials provide information about social processes and in which way they do it (Dietler and Herbich, 1998).

In this sense, the concept of chaîne opératoire involves a theoretical and methodological framework that includes several technological parameters useful for addressing the social practices and interactions underlying the features of the vessels. Furthermore, J. García Rosselló (2008) has summarised some of the indicators, many of them related to the chaîne opératoire, which may provide significant information to approximate the way the production and the consumption of ceramics is organised in society. These markers are: 1) the location of production centres, 2) the kind and quantity of the pottery produced, 3) the patterns of pottery distribution, 4) the organisation of the spaces involved in the production process, 5) raw material management, 6) the techniques used in the production, 7) the variability of fabrics and types, 8) the physical properties of the end products, 9) the waste volume and 10) the tools involved in the production.

The best option to deal with the organisation of production in ancient societies consists in using as many of these indicators as possible. Nevertheless, we must be aware that most of them can be only addressed from direct evidence obtained from the study of the production contexts, i.e., workshops or working areas. Unfortunately, the difficulty for identifying working areas related to prehistoric ceramics production is well-known, and usually the clearest proof is the finding of firing structures. In other cases, the digging of raw materials and tools involved in pottery production may suggest that this activity could have taken place in certain contexts (e.g., domestic areas) which could have also been used as working areas. Likewise, the development of the whole production process in a single space and close to the firing areas may explain, for instance, the occasional occurrence of charcoals and other components in ceramic pastes or fabrics (Neupert, 2000; Riederer, 2004).

This difficulty to record production areas could be related, in many cases, with the management of these spaces. There are ethnographic cases, such as the Bantu potters from South Africa, in which the artisans do not have a fixed place to carry out their production, thus it can be placed in different locations which involve both domestic spaces and outdoor areas or yards. In this way, all the equipment and materials involved in pottery production (e.g., tools, clay, water, etc.) can be easily transported, often inside the own vessels (Krause, 1984). In a similar way, it is not easy to identify production areas by the presence of tools, since there may exist a wide variability in the tools used to make a vessel (Vidal and Mallía, 2011), even within the same ethnic group. For instance, among the Konkomba potters from Northern Ghana (Albero et al., 2013) an opportunistic use of the tools is documented, so that a wide range of instruments are used for the same purpose.

In short, as some scholars have pointed out (Costin, 2000; Tite, 2008), the organisation of pottery production usually must be inferred through pottery technology itself, from elements such as the degree of standardisation, the use of certain techniques and materials and their relation to certain equipment or strategies of space management, the potter’s skills, the patterns of pottery distribution, etc. In the case of archaeological ceramics, given the data provided by the archaeometric analysis of fabrics and the comprehensive study of the depositional context of the vessels, the organisation of production can be addressed by paying attention to the location of production centres, the kind and quantity of the pottery produced, the patterns of pottery distribution, the raw material management, the techniques used in the production, the variability of fabrics and types and, finally, the physical properties of the end products. Unfortunately, it is impossible to directly address the management of the spaces involved in the production, the volume of waste generated and the tools involved in pottery making when the working areas have not been accurately identified. Anyway, some inferences regarding these aspects can be indirectly carried out through the study of pottery fabrics.

Summing up, the analysis of the social organisation of pottery production can, on the one hand, be developed by means of “black boxes” based on transcultural models established beforehand that can be compared with the archaeological record. On the other hand, a reference framework based on the contextual knowledge available regarding the socioeconomic organisation of the communities under study can be used. Several aspects (e.g., specialisation and standardisation degree, potter’s skills) that are useful to approach the organisation of production from paste studies can be addressed in these particular contextual frameworks through the indicators listed above. In addition, the study of these parameters and the social contexts in which ceramics are framed represent a process that is both inductive and deductive, thus providing more strength and significance to the interpretations regarding the organisation of pottery production and its relation to the organisation of society.

20.1 Level of Specialisation in Pottery Production

Cultural evolutionary positions have related ceramic features to certain forms of more or less specialised production in order to determine the complexity of any specific society. However, we can go beyond such a definition of complexity and consider that the level of production specialisation is related to certain technological behaviours undertaken by potters and users as a response to given social needs. These relations between specialisation and social organisation are neither readily visible nor predictable, but the fact is that potters develop their production taking into account a given reality and the society as a whole (Barley, 1994; Harry and Bubemyre, 2002).

Issues regarding pottery specialisation can be archaeologically addressed through the study of the technological attributes of the pottery and its contexts of production and use, since they involve conscious or unconscious actions carried out by the potter and other individuals. In this sense, valid cultural assumptions about the organisation of production and its level of specialisation cannot be established until the features of the vessels have been fully addressed.

Ethnographically, the level of specialisation of pottery production is often identified by means of the time that the potters spend in their production, commonly expressed as full time or part time pottery productions. Furthermore, it should be considered whether the potters depend solely on ceramic production for their living or whether they carry out a variety of activities in order to diversify the means for providing the income needed for the production unit (Costin, 2000; García Rosselló, 2008).

Other authors (Vaughn and Neff, 2004) state that a specialised behaviour involves a level of pottery production that exceeds self-consumption needs. In this sense, C. Costin (1991, 2005) and M. Hagstrum (Costin and Hagstrum, 1995) define specialisation as a regular and permanent production system in which potters depend on the existence of exchange relations that go beyond the sphere of the household. Thus, potters need to distribute their pottery in order to subsist, while non-producers need to acquire certain ceramic assemblages to carry out certain activities.

In short, the elements that define craft specialisation are, first, an intensity and level of production exceeding the needs of the potters and their households. Second, the use of the vessels by individuals who are not involved in the manufacturing process but have access to the products by means of diverse exchange relations (e.g., trade, clientelism relations, etc.) and, in consequence, play a key role in the development of specialised productions. In this sense, in specialised pottery productions consumers and producers usually do not belong to the same household.

These definitions imply a huge variability regarding the kinds and levels of specialisation that can be identified. In addition, these perspectives entail constraints when they are applied to the study of specialisation in archaeological contexts. As said before, in these contexts we must necessarily approach the social organisation of production through the characteristics of the ceramic record itself. In this sense, specialisation involves a standardisation of the strategies of production that often coincide with the social regularisation of the resources used by a specific group.

This adjustment in pottery production is reflected in a reduction of the variability found in the end products manufactured from these resources and the development of certain potter’s skills. Thus, pottery assemblages that are quite homogeneous regarding the materials, techniques, fabrics and shapes used in their manufacture are considered a consequence of specialised behaviour. In contrast, heterogeneous ceramic assemblages with a high variability suggest the opposite phenomena (Clark, 2007; Rice, 1987). In short, the potter’s skills and the level of variability of the ceramic assemblages are key aspects that have to be fully addressed in order to study specialisation in pottery production.

Finally, it should be noted that potters producing hand-made vessels do not have certain mechanical devices available, such as the potter’s wheel or moulds, which allow an individual to accurately control certain aspects of the vessels (e.g., symmetry or wall thickness) that are considered characteristic of specialised pottery productions. However, we consider that the absence of these features in the vessels does not preclude the fact that certain experience, practice and knowledge and, therefore, relatively specialised potter’s skills are involved in the development of any pottery production. These assumptions are especially noticeable when hand-made forming methods are used, since potters spend a long time learning how to process the paste and model the pottery in an attempt to produce vessels with acceptable sizes and shapes according to a certain range of possibilities. In short, we should consider that the parameters discussed in this section are not universals and cannot be applied without some reflection. In this sense, the concept of specialisation may be contingent and valid for some cases but not for others, thus being a parameter unsuitable to address the organisation of production in many societies.

20.2 Level of Variability in Pottery Production

As we have seen, archaeological research has often focused on the technological variability of ceramics as a valid way to study the organisation of production and its level of specialisation. The variability of ceramic assemblages can be estimated by classifying the pottery in different categories and recording their differences and similarities according to certain spatial and temporal segments. This procedure helps indicate whether certain materials and techniques were repeatedly used in a routine way in pottery production and if they were developed with the aim of achieving ceramic products with specific properties.

Hence, the range of variability of the artefacts may change both in time and space and may increase with social distance. It is within this framework where information is constantly renewed and diverse material and cultural alternatives take place. In this sense, as explained in the previous sections, the study of the heterogeneity and homogeneity of ceramic assemblages ultimately allows us to approach certain political and social aspects underlying pottery production. The spaces in which production is carried out and the contexts of use of the artefacts are also an active part of technology and can influence the level of production as well as the variability observed in the archaeological record. For instance, different features can be present in a ceramic assemblage depending on whether the production and use of the vessels takes place either in public or private spaces (Arnold, 1999; Balfet, 1984; DeBoer, 1984; Harry and Bubemyre, 2002; Longacre, 1999; Rice, 1984b).

As just mentioned, specialisation in pottery production can be related to multiple kinds of social organisation, thus being a dynamic and multidimensional concept. Hence, the multiple ways the variability of ceramic assemblages may be related to the specialisation of pottery production (see, for instance, Arnold, 2000; Longacre, 1999) militates against any proposal that establishes straight connections between them. In fact, this same complexity explains why pottery should be considered from a multidimensional perspective which includes both micro and macro levels of analysis (e.g., archaeometric methods, macroscopic analysis, typological classifications, contextual studies, analysis of absorbed residues, use-wear traces, etc.) if a thorough interpretation of the organisation of production and the level of variability is to be reached.

According to B. Sillar (1997), the existence of a more or less specialised production within a potter community depends on how individuals perceive and use local resources, the knowledge transfer systems and the social organisation itself. Therefore, it is highly relevant to insert the inferences made regarding pottery production in a specific cultural context, since it makes it possible to interpret the level of variability of the vessels according to specific forms of social organisation. On many occasions the social factors are the only elements that can explain the variability observed in the material record and its significance, thus they must be integrated in our explanations of the role of pottery technology in society. For instance, an increase in the variability found in the ceramic assemblages may be related to the creation of new recipes and operational sequences as a result of changes in the learning context and knowledge transfer systems (Schiffer and Skibo, 1987). These alterations in learning contexts may be related, for example, with a breakdown in the organisation of pottery production as a result of certain social dynamics that affect society as a whole and promote changes in the social role of the vessels (Albero, 2011a).

The use of archaeometric approaches in pottery studies and the implementation of statistical analysis are absolutely necessary to accurately establish the variability degree in pottery production (Kvamme et al., 1996). However, some authors (Arnold, 2000; Longacre, 1999; Rice, 1991) argue that it may not be appropriate to measure variability with methods based on accurate metric systems. This is because it is usually very difficult to manufacture two identical vessels in hand-made pottery production, especially when the pottery is produced by individuals with scarcely specialised skills.In this kind of pottery production, the standardisation and the sharing of ideas can take place at low numerical resolution, since the specific know-how of each potter determines the form of the pottery, even if there are common criteria among all the individuals involved (Miller, 1985). Thus, certain variations exist among the products performed by a single potter but also and among the ones manufactured within a potter community. In this aspect, the macro/micro-scale chosen and the methods used to collect the data, as well as the statistical techniques applied, may be often excessively precise and rigid for hand-made pottery. Thus, the concept of variability and its categorisation can be a response to specific cultural logics. As a consequence, it is not always possible to pinpoint a direct correlation between the variability that researchers record in the chemical, mineralogical or textural composition of pottery fabrics and the socially meaningful practices that the potters develop. In this sense, we must be aware that a given technological choice can exhibit qualitative and quantitative differences along time and space, but it may still have the same cultural significance to the people who made it.

20.2.1 Low Variability

In spite of the constraints that the use of the concepts of specialisation and variability entail in the study of the organisation of pottery production, it is common to associate the presence of specialisation with a low variability in certain features of the vessels. However, the standardisation of the products can be related to two different kinds of phenomena that are not mutually exclusive (Longacre, 1999):

a) Standardisation that is the consequence of daily praxis, i.e., the application of certain technical gestures based on the repetition of habits and the experience gained over time since apprenticeship. In this sense, domestic pottery productions, as observed in north-east Ghana (Albero et al., 2013; Calvo et al., 2013), may have some level of standardisation in several aspects, such as the paste recipes or the shapes and sizes of the vessels. For instance, in these cases, the potter feels or intuitively determines when the clay is ready or what amount of raw material is needed to perform each kind of vessel. However, there may be slight differences in the ceramic bodies that permit distinguishing one potter from another, as well as the products performed by a given potter community.

b) Standardisation as a result of the conscious extra effort that potters invest to homogenise their production according to technical and social criteria. In these cases, the demands made by the users may promote a more stable production, thus influencing aspects such as the paste, the shape or the size of the vessels. In this sense, W. Longacre (1999) ethnographically documented that some potters recognise that the demand of the vessels largely determines the level of specialisation of their production.

In any case, a low technological variability is usually related to pottery productions which are exclusively developed by a small number of individuals within the community. In this way, both information and knowledge are well-established and controlled and they flow easily. In this case, highly skilled potters manufacture vessels that are going to be used by the rest of the community and even by people outside it (Harry and Bubemyre, 2002). A low variability at the regional level may also be related to seasonal or cyclical movements across the territory undertaken by specialised potters who develop an itinerant production. Moreover, standardisation also results when a potter gets married and moves to a different household, thus influencing the technological choices and ways of manufacture common to other individuals (Albero et al., 2013; Gosselain, 2008). Normally, these kinds of highly skilled potters often sell their products inside the production units or in local or regional markets (Calvo et al., 2013). Depending on the demand, pottery production can be carried out according to a previous demand, or conversely, the potter can have some stock available.

S. van der Leeuw (1984) stated that an increase in the level of production or the development of communal pottery firings may promote standardisation and a significant reduction of the variability of the fabrics and vessels produced within the group. Relatively standardised pottery pastes have a very similar physicochemical behaviour when heated, thus facilitating the control of the firing process and the achievement of appropriate products. Hence, the low variability of pastes allows potters to reduce the risk of failure during firing. That is the reason why many potters are often very conservative regarding the paste preparation process. According to Mommsen (2004), this strategy aims at preventing possible economic losses, especially those resulting from the failures that take place during the firing process.

In functionalist terms, it is considered that certain types of vessels, such as cooking pots, are generally more stable and less variable regarding paste, shape and decorative patterns. In these cases, the design is strongly determined by functional parameters (Albero and Puerta, 2011; Djordjevic, 2003; Fernández Navarro, 2008) and potters develop an extra effort to achieve highly standardised productions with specific characteristics. Nevertheless, there are many traditional societies, like the Shipibo-Conibo from Peru (DeBoer, 1984), in which the whole ceramic assemblage, including serving bowls, show a low degree of variability over time regarding their shape, size, colour and other features. In this case, the level of standardisation does not respond to the function of the vessels, but rather to the continuous practice explained before.

20.2.2 High Variability

The presence of a high compositional variability within the ceramic assemblage of a single archaeological site can be related to several reasons. It may be evidence of the coexistence of multiple production units within the site that have different ways of making pottery as well as the development of individual actions within certain production units that do not follow the general trend. Thus, O. Gosselain (2008) ethnographically documented in western Niger that an increase in the variability degree of the fabrics may indicate little connection between the way the ceramic paste is processed and the identification of the potters with particular social groups or specific production centres.

Thus, an increase in variability may indicate the absence of agents that control or influence the production of the potters, thus evidencing decentralised production systems that are not subject to any political, social or economic control. Thus, from a capitalist point of view, the vessels are not thought to be “put into circulation” when the strategy corresponds to domestic productions for self-consumption. Thus, in these cases, there is not any market demand that determines the need for manufacturing products with specific features; and there is no competition between different potters which may promote the production of more standardised ceramic assemblages.

Furthermore, a high variability of the ceramic assemblage associated with an archaeological site may be also the consequence of the regional distribution of certain ceramics. In this sense, the vessels are more or less intensely distributed across the territory by means of the contacts and exchanges that individuals carry out, thus increasing the variability of the ceramic record. For instance, in the ethnographic studies conducted in Ancash in Peru (Druc, 1996) or in north-east Ghana (Calvo et al., 2011, 2013) it was observed that family ties and certain exchange networks promote a wide distribution of the pottery. In these cases, the vessels are ultimately used in domestic contexts that can be located tens of kilometres from the production centre. Moreover, the presence of low standardised ceramic assemblages in the same archaeological site may be also related to the presence of several itinerant potters who carry their own raw materials along the territory. Such kinds of itinerant potters, who usually have highly specialised skills, are ethnographically well documented in places such as Mallorca (Llabrés, 1977), Ancash (Druc, 1996), the Andes (Sillar, 1997) and Cyprus (Cooper, 2000).

As several authors note (Arnold, 2000; García Rosselló, 2008; Livingstone-Smith, 2000, Roux, 2011), the spatial distribution of the settlements in the territory, their function and interrelation can influence the way pottery is produced, distributed and deposited, as well as its characteristics and degree of variability. For instance, a dispersed settlement pattern can help increase the variability of the ceramics found in a certain region. Similarly, the variability of the ceramic record also increases significantly when communities or groups coming from diverse settlements deposit their pottery in a single archaeological site. In this case, as often happens with sites that have a ritual or funerary function (e.g., Albero, 2011a), a number of individuals usually provide different pottery vessels which are related to diverse raw materials and fabrics, thus promoting heterogeneous ceramic assemblages. In short, to correctly interpret the degree of variability of the pottery, it is important to determine the origin of the vessels found in the archaeological record and to assess the way in which the space was used as well as how the settlement is distributed along the territory.

20.3 Potter’s Skills

Another aspect that has been usually associated with the organisation of pottery production and the level of specialisation, as well as with the properties and variability of the vessels, is the potter’s skills. In this way, potters play a significant role in the generation of an iterative process of actions that enable certain standardisation degree in the production as well as the transfer of technological knowledge (Budden and Sofaer, 2009). As seen in previous sections, the knowledge that potters acquire regarding their craft is an essential element of technology and plays an important role in the reproduction of material culture. Thus, the maintenance or change of any particular potter’s skills – i.e., certain knowledge and abilities – is related to specific social dynamics associated with particular learning strategies (Budden, 2008, 2009).

The level of specialisation of the potter’s skills is related to the regularity with which the artisans use specific techniques and procedures along the production process in order to develop a particular way of making pottery. In this way, an increase in the intensity and level of production also raises the repetition of certain technological actions and experiences of the individuals. In this sense, through continuous practice, the potters can develop certain skills or behavioural patterns that can lead to more standardised ceramic assemblages. Hence, differences in the features of the pottery are often considered related to certain production processes and they can ultimately reflect different potter’s skills (Arthur, 2003, Blackman, 1993; Budden, 2008, 2009; Budden and Sofaer, 2009; Costin and Hagstrum, 1995; Ortega et al., 2005; Roux, 2003). For example, research undertaken in San Nicolas (Philippines) focused on the domestic production developed by three generations of potters, recorded that the vessels which were shown to be more standardised statistically were made by the older artisans. Although these individuals are more familiar with the pottery production and have more expertise, there is also a strong formal resemblance between the products made by potters from different generations (Longacre, 1999).

In short, the characteristics of the vessels in terms of the materials and techniques used for their production, as well as the properties of the end products, are all key elements to address when considering the potter’s skill. The kind of paste processing technique or the quality of the forming methods, as well as the finishing surface treatments and the firing procedures may reveal important information about the potters themselves, their involvement in pottery production and the consistency of the learning system at play. In this sense, it is considered, for instance, that the firing of pottery requires certain specific knowledge and potter’s skills in order to make possible the maintenance and the reproduction in time and space of certain strategies and structures involved in this phase of pottery production (Sillar and Tite, 2000).

Taking these premises in mind, two main categories representing different potter’s skills can be established according to the manufacturing processes in practice and the kind of final product obtained. These general models are not universal: as always, the wide potential diversity present in potter’s skills as well as their influence by cultural peculiarities should be considered.

a) Low-skilled potters: These potters, who develop the whole chaîne opératoire by themselves, use raw materials hardly purified and processed. Moreover, the individuals are not very familiar with the properties of the raw materials, and pastes are often poorly mixed and homogenised. In addition, the joints of the coils are poorly sealed and uneven firings are frequent. The end products are usually slightly asymmetrical and there is little consensus regarding the fabrics and types obtained, which are often related to polyfunctionality phenomena. In many cases, pottery with such features may be associated with products performed by assistants or apprentices (Calvo et al., 2014; Vidal, 2011b). In short, all these aspects determine a high variability in the technological choices made during the production process, thus promoting a low standardisation of the ceramic assemblages.

b) Highly-skilled potters: These artisans use highly standardised raw materials and effectively prepare and homogenise the pastes. They also take care that the junctions of the coils are well sealed during the modelling stage. In addition, both pottery firing and surface treatments are usually quite uniform. The end products and fabrics obtained are compositionally homogeneous, symmetrical and standardised regarding their typology. Moreover, the physical properties of the vessels often respond to their different functions. In short, in this case, the variability of the ceramic assemblages is significantly reduced and is organised, according to typological and functional criteria. This approach can be useful since all these variables are also related, in addition to the potter’s skills, to the effort and time invested in the development of pottery production. However, it must be also considered, as we observed ethnographically, that highly skilled and specialised potters can adjust their manufacturing process and the type of end products performed according to the customers’ demand and the social or historical context in which their production takes place. In some cases, highly skilled potters who perform specialised productions can manufacture vessels of low quality for their personal use, can keep for themselves those products with faults that cannot be put into circulation, or they may simply not need to make a good quality vessel.

A productive system involving several crafts organised into guild-like associations – i.e., groups of individuals with a common internal structure and similar skills who carry out relatively specialised practices – would have favoured a collective understanding of the social order. In this sense, the interaction that takes place between artisans who share common skills provides some cohesion between the individuals of the community and, especially, among those who develop the same activity (Sofaer, 2006).

Furthermore, it must be taken into account that pottery production may involve one or more individuals when addressing the potter’s skills through the features of the vessels. In this way, a potter can develop the production alone or may require some assistance throughout the process (Crown, 2007, Vidal, 2008b). Thus, the perception and visualisation that the potters have of the entire manufacturing process can be significantly reduced in complex production systems in which the different phases of pottery production are developed by different individuals. In such systems, the individuals may carry out their actions exclusively in specific phases of the chaîne opératoire. In these situations, the standardisation of the features of the ceramic assemblages does not necessarily imply that all the individuals involved in the production process are related to highly specialised skills.

For instance, in Quinua (Peru) mould-made pottery production promotes an intense level of production as well as highly standardised products. However, no highly skilful potters are needed to model the vessels (Arnold, 1999). This is because the different phases of the chaîne opératoire are segmented and carried out by different individuals, facilitating a greater specialisation and efficiency in the manufacturing process and the consequent increase in the level of production. Segmentation of the production process gives place to a more complex form of organisation in which a delocalisation of the knowledge involved in the manufacturing process occurs. Actually, the knowledge that was previously concentrated in one highly skilled potter is now diffused across multiple individuals, thus giving place to the so-called aggregated skills. As we observed in Fez (Morocco), it is quite unlikely that a single potter accumulates all the knowledge required to develop the entire process of creating a vessel in highly specialised and intense production systems. In short, despite obtaining highly standardised products, such kinds of specialised pottery productions may be related to low-skilled potters. In these cases, these potters’ skills are a consequence of the limited interaction and knowledge transfer between the individuals in the whole production process.

Unfortunately, it is not possible to know only from the ceramic itself and without studying the working areas if a vessel was produced by means of a segmented strategy or, in contrast, by a single potter. Thus, it is obviously advisable to address the organisation of pottery production and the potter’s skills not only through the manufacturing techniques and the variability degree of the vessels, but also by approaching the segmentation degree of the manufacturing process. In this sense, there are systems of production that require, in addition to certain techniques and materials, the aggregation of more or less specialised individuals organised according to certain operational sequences. These sequences arrange the behaviours of the different individuals along the manufacturing process, giving place to specific chaînes opératories that allow the production of certain types of vessels (Arnold, 1999; Balfet, 1984).

As already seen, unilinear correlations between social complexity, specialisation of production, the degree of variability of the end products and the potters’ skills are not without problems. In this sense, certain more or less specialised behaviours do not necessarily materialise with respect to products with specific features. In fact, the interpretation of the organisation of pottery production and the features of the vessels in the archaeological record requires considering the multiple interactions developed by the potters in their social environment. Hence, there are many types of specialised pottery productions depending on the social ties existing between the potters that may result in well-standardised vessels.

For instance, consensus in the ways of organising the production may be easier to maintain in small groups of potters with relatively specialised skills. The reason is that in such contexts the information easily flows among all the individuals, thus promoting categories that are widely shared by the different potters. This consensus is materialised in similar end products that, in some cases, are even subject to certain controls, for instance by means of social prohibitions. These social mechanisms of control are developed in order to reduce innovation or competitiveness and, ultimately, prevent change and promote stability in the social organisation of production in which artisans are inserted as well as in the products manufactured (Dietler and Herbich, 1998; Rice, 1984a; Roux, 2003).

In other cases, high-skilled potters can develop their productions side by side with low-skilled potters within the same community. This form of organization can be seen, for instance, in the Terramara culture (Italy) since the Bronze Age (Brodà et al., 2009) and has been also proposed for the Talayotic culture in the Balearic Islands (Lull et al., 2008). In these cases, specialised potters are normally responsible for the manufacture of certain types of vessels or the use of certain techniques typical of specific potter skills.

On the one hand, low-skilled potters can produce smaller vessels related to serving wares that have a more limited life expectancy, require less time investment and are easier to achieve. On the other hand, as pointed out by experimental studies (Brodà et al., 2009), large and slender vessels shaped by means of the coiling method entail higher technical difficulty. As a consequence, the manufacture of these kinds of vessels requires highly skilled potters as well as greater effort and time investment. In addition, Clark (2007) notes that the firing is more complex in large vessels, especially when this process is undertaken in open firings, due to the uneven heat distribution along the ceramic body. In short, this kind of division in pottery production can be related to the maintenance of certain social order. In this line, it is argued that Inka political domination in Peru is materialised in the ceramics through the control of the production of large vessels used to store the surpluses produced (Dobres and Hoffman, 1994).

In cases in which low-skilled and high-skilled potters share infrastructure (e.g., firing structures) any time during the production process, the less specialised potters can use paste recipes, techniques and materials similar to the ones used by more specialised individuals. Such convergence in the technological choices is carried out in order to ensure the success of all products made by means of this collective strategy (Brodà et al., 2009; Budden and Sofaer, 2009).

This diversity of productive strategies characterised by a dichotomy in the types of pottery made by low-skilled and high-skilled potters has some interesting implications. For example, some authors (DeBoer, 1984, Mills, 1989) explained the technological change and the level of variability in the ceramic record considering the use life of the vessels and the frequency and intensity of production.

As previously mentioned, B. Mills (1989) distinguishes different rates of change in the vessels depending on their size, spatial location and frequency of use. These aspects determine the regularity of the pottery production and determine the level of variability of the end products. For example, large ceramics such as storage jars normally have a reduced mobility and are located in confined spaces, thus having a longer use life and thus being less regularly manufactured. A low variability degree in large vessels requires, therefore, highly specialised and experienced potters who produce these kinds of materials, as seen among Kusasi in Upper East Ghana (Calvo et al., 2013). These potters’ skills permit them to maintain a formal coherence between one vessel and another across time through daily practice.

In contrast, DeBoer (1984) argues that a greater variability should be expected in those vessels whose production is more frequent and continuous, i.e., smaller vessels used to serve, prepare and cook foodstuffs. Thus, regardless of a potter’s skills, an increase in the number of firings and the pottery produced also increases the chances for individuals to express themselves through the ceramics and therefore to experiment and introduce innovations in the manufacturing process. In these cases, the variability of the ceramic assemblages can be high and changes in the ceramics can take place more frequently.

Finally, we can note another model, observed for example in Pòrtol in Mallorca (Albero and Puerta, 2011), characterised by a specialisation among diverse potter communities or even within a single community. In these cases, high-skilled potters produce only certain types of vessels associated with specific functions and related to the use of certain techniques and materials. Such types of organisation regarding production may promote a significant social interdependence between the different potters of the community, since the set of ceramics used in everyday life involves the complementary use of all the kinds of vessels that are produced (Sillar, 1997).

We can conclude, then, by noting that it is not appropriate to determine the organisation of pottery production on the basis of the potters’skills alone. To this end, it should be considered that the level of specialisation of the potters along with the intensity of production and the variability of the ceramic assemblages within specific contexts also play an important role. Similarly, it is not possible to understand this situation by considering only the intensity of production, since this increase and the amount of production does not necessarily involve a higher standardisation of the ceramic record (Arnold, 1999).

Image

Figure 20.1: Conceptual map of the main variables involved in the study of the social organisation of pottery production and the possible archaeological evidence related to these parameters.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset