Svoboda | Graniru | BBC Russia | Golosameriki | Facebook
Academia.eduAcademia.edu
The Retrospective Methods Network Newsletter Winter 2014/2015 №9 Edited by Frog Helen F. Leslie-Jacobsen and Joseph S. Hopkins Published by Folklore Studies / Dept. of Philosophy, History, Culture and Art Studies University of Helsinki, Helsinki 1 Tradition. Ed. Frog, Anna-Leena Siikala & Eila Stepanova. Studia Fennica Folkloristica 20. Helsinki: Finnish Literature Society. Pp. 82–119. Uther, Hans-Jörg. 2004. The Types of International Folktales: A Classification and Bibliography, Based on the System of Antti Aarne and Stith Thompson: Tales of the Stupid Ogre, Anecdotes and Jokes, and Formula Tables. Helsinki: Suomalainen Tiedeakatemia. Venditti, Chris et al. 2006. “Detecting the Nodedensity Artifact in Phylogeny Reconstruction”. Systematic biology 55(4): 637–643. Vialou, Denis. 1987. L’Art des Cavernes: Les Sanctuaires de la Préhistoire. Paris: Le Rocher. Vinson, Julien. 1883. Le folk-lore du pays basque. Les littératures populaires de toutes les nations 15. Paris: Maisonneuve. Wagner, Peter J., & Douglas H. Erwin. 1995. “Phylogenetic Patterns as Tests of Speciation Models”. In New Approaches to Studying Speciation in the Fossil Record. Ed. D.H. Erwin & R.L. Anstey. New York: Columbia University Press. Pp. 87–122 Webster, Andrea J., Robert Payne & Mark Pagel. 2003. “Molecular Phylogenies Link Rates of Evolution and Speciation”. Science 301: 478. Webster Wentworth. 1879. Basque Legends: Collected, Chiefly in the Labourland. London: Griffith & Farran. Wichmann, S., et al. 2011. “Correlates of Reticulation in Linguistic Phylogenies”. Language Dynamics and Change 1: 205–240. Wissler, Clark, & D.C. Duvall. 1908. Mythology of the Blackfoot Indians. Anthropological Papers of the American Museum of Natural History 2.1. New York: American Museum of Natural History. Ralston, William Ralston Shedden. 1873. Russian Folk-Tales. London: Smith, Elder and Co. Rangel, T.F., et al. 2010. “SAM: A Comprehensive Application for Spatial Analysis in Macroecology”. Ecography 33: 46–50. Richter, Fr von. 1889. “Litauische Marchen: Der einaugige Riese”. Zeitschrift fur Volkskunde 1: 87– 93. Rohrich, L. 1976. Sage und Märchen: Erzählforschung heute. Freiburg: Herder. Ronquist, F., & J.P. Huelsenbeck. 2003. “MRBAYES 3: Bayesian Phylogenetic Mixed Models”. Bioinformatics 19: 1572–1574. Ross, R.M., et al. 2013. “Population Structure and Cultural Geography of a Folktale in Europe”. Proceedings of the Royal Society B: Biological Sciences 280: 1756. Simms, Stephen Chapman. 1903. “Traditions of the Crow”. Field Columbian Museum, Publication 85. Anthropological Series 2.6. Chicago: Field Columbian Museum. Pp. 281–324. Spence, Lewis. 1914. The Myths of the North American Indian. New York: Thomas Y. Crowell Company. won Stephanowissel, Karadschitsch, with W. Tochter. 1854. Volksmarchen der Serben. Berlin. von Stier, Gaals Georg. 1857. Ungarische Volksmarchen. Pesth: Heckenast. von Sydow, Carl Wilhelm. 1948 [1965]. “Folktale Studies and Philology: Some Points of View”. In The Study of Folklore. Ed. A. Dundes. Englewood Cliffs: Prentice Hall. Pp. 219–242. von Sydow, Carl Wilhelm. 1927. “Folksagorforskningen”. Folkminnen och Folktankar 14: 105– 137. Tehrani Jamshid J. 2013. “The Phylogeny of Little Red Riding Hood”. PLoS ONE 8.11: e78871. Thompson, Stith. 1961. The Types of the Folktale: A Classification and Bibliography: Antti Aarne’s Verzeichnis der Märchentypen Translated and Enlarged. 2nd rev. edn. FF Communications 3. Helsinki: Academia Scientiarum Fennica. Tolley, Clive. 2012. “On the Trail of Þórr’s Goats”. In Mythic Discourses: Finno-Ugrian Studies in Oral Internet Sites Glottolog. http://glottolog.org/. WALS = World Atlas of Language Structures. http://wals.info/. De situ linguarum fennicarum aetatis ferreae, Pars I Frog and Janne Saarikivi, University of Helsinki Abstract: This article is the first part of a series that employs a descendant historical reconstruction methodology to reverse-engineer areas where Finnic languages were spoken especially during the Iron Age (500 BC – AD 1150/1300). This opening article of the series presents a heuristic cartographic model of estimated locations of groups speaking Finnic languages and their neighbours in ca. AD 1000. The aim of this article is to provide the first of three maps of the Uralic-speaking peoples in Northwest Europe in three approximated periods: ca. AD 1000, AD 1 and a map indicating the linguistic Urheimats of reconstructed intermediate proto-languages within the Uralic language family (Proto- Finnic, Proto-Sámi, Proto-Mordvin, etc.). For reasons of length, the aim of providing three different maps which involve different materials and present different issues has required presenting the investigation as a series. The present article is only the first part 64 the status quo in current research and for the purposes of future discussion. Articles of the present series are especially important owing to fundamental changes in understanding of the history of West Uralic languages across roughly the past decade. This investigation employs the methodology of descendant historical reconstruction. Focus here is on the spread of languages as an intergenerationally transmitted dynamic and evolving social and historical phenomenon, not on the specific features of lexicon or linguistic structures. Modelling language distribution and change across the Iron Age has highlighted the need for a developed discussion and presentation of the methodology applied here. As the issues of methodology are dependent on the empirical evidence as a frame of reference for discussion, it was decided that the outcomes of the study should be presented first. Therefore the methodological framework will only be briefly outlined here along with introductions to relevant terms and concepts with a concentrated presentation discussed in relation to the model of all three maps to follow in a later issue of this journal. The presentation offered here opens with some general remarks about language and culture along with an outline of methodological points and issues. The body of the article introduces the groups indicated on Map 2 (e.g. F13: Satakunta) with comments on sources, points about language and other distinctive details, and also remarks on etymologies of names in those cases where these are possible to establish with a reasonable degree of certainty. In organizing this information, it was decided to present first those groups and cultures neighbouring the Finnic peoples to reduce repetition. Introductions to Germanic and Slavic groups in particular provide frames of reference for subsequent groups. Additional probable Uralic groups speaking non-Finnic languages will be presented following the Finnic groups. Ethnonyms and toponyms from Old Norse and Old Russian sources are given in standardized forms for orthographic convenience and consistency (with exceptions noted). In the case of Old Norse, this means that forms are Icelandic/West Norse, although of this work and it focuses on the first map for the period ca. AD 1000. The goal of presenting maps for these three periods developed as a satellite of the Viking Age in Finland (VAF) project. The VAF set out to examine cultures and historical changes during the Viking Age in territories of Finland and Karelia1 as well as the discourse surrounding them as viewed from the perspectives of different disciplines (see Ahola et al. 2014b). Written sources with detailed information about groups inhabiting these areas in general significantly post-date the Viking Age. Each discipline is thus heavily reliant on findings of other disciplines when developing a perspective on culture formations in that period, and they generally frame the Viking Age in relation to preceding and subsequent periods. In the VAF project, it rapidly became clear that maps could provide a significant tool in negotiating the perspectives of different disciplines, especially when seeking to relate data in the archaeological record, which can be pinpointed in cartographic space on an absolute chronology, with intangible cultural phenomena such as language, folklore and cultural semiotics, of which a chronology and geographical spread can only be reconstructed through later evidence. Suitable maps were lacking and the present work has set out to fill that need, especially for Finnic language areas during the Iron Age. The multidisciplinary impetus and methodological concerns led the authors to combine a wide range of information so that the first map presented here concentrates on cultural areas of speech communities and their networks rather than seeking to outline dialect areas per se. This map concentrates on historically attested groups that can be identified with Finnic speakers as well as neighbouring groups relevant to contextualizing them. The resulting image offers an impression of the language situation in Northeast Europe in roughly the Viking Age in the light of scarce evidence that is often difficult and problematic to interpret. Needless to say, the image developed in this article remains hypothetical and heuristic. Nevertheless, this does not diminish the value of such a model as a frame of reference for 65 even these are not necessarily unproblematic,2 and the term ‘Norse’ is used here as generally inclusive of all Scandinavian dialects (which were mutually intelligible in AD 1000) unless otherwise specified. Both Old Russian and Modern Russian forms are given in Latinized transcription. The model presented here is centrally intended to provide for researchers of different disciplines a general frame of reference that, on the one hand, can be advanced and refined through future discussion and, on the other hand, can provide a platform for modeling language distribution and spread in earlier periods of history that will be developed with maps in future parts in this series. involving a large-scale cultural shift without an actual language shift taking place. In this type of process, changes in a language reflect a cultural change, but the inter-generational transmission of language is unaffected (Thomason & Kaufman 1988; cf. also Saarikivi 2014). Cultural and thus also linguistic change can be considered a prerequisite for the preservation of cultural heritage in the changing world. A lot of cultural and thus linguistic change takes place in horizontal cultural networks where neighbours borrow technologies, raw materials, artifact types, religious concepts, cultural habits, traditions and socio-cultural ideologies from each other, or imitate each other’s culture (on borrowing in a typological perspective with a review of the fields of vocabulary that are most salient for borrowing, see Haspelmath & Tadmor 2009). On the other hand, language also represents a kind of linguistic heritage that is straightforwardly attached to concrete geographical areas, most notably, toponymy, and the vocabulary denoting locally bound features of geography, flora and fauna. These are local phenomena attached to concrete places, for locally bound features, and they typically spread in vertical cultural networks – i.e. they are transmitted from one language to another in a specific area, even if a language shift takes place (cf. Saarikivi 2000). However, the spread of toponymic types (i.e. naming models) has also been associated with migrations and quite remarkable results regarding the movement of people have been reached by studying their spread (cf. Kuzmin 2014a; Vahtola 1980; Kiviniemi 1977). Thus, toponyms may demonstrate both local language substrates (i.e. earlier languages of a particular region) as well as the spread of people within one language area, although the methodologies and materials needed for their analysis in each of these cases are different. It is quite self-evident that the dynamics of spread differ considerably for a) a language as a full-fledged semiotic system, b) only the lexical or structural features of a language, and c) locally bound linguistic heritage. These three categories of linguistic phenomena also have different counterparts in community change and in the archaeological record. To Language and Culture: Preliminary Remarks When discussing language spread, it is necessary to distinguish between different aspects of linguistic heritage that may have different distributions. A language is typically transmitted inter-generationally within a community as a mother tongue, or in between different communities as a lingua franca (an intergroup language). They can spread areally by the migration of a population, or, what is likely to be a more widespread phenomenon, through a language shift caused by a language spreading in culturally dominating social networks. Although a language shift typically involves some migration of socially influential people, this phenomenon is to be kept distinct from mass migrations of language communities which are typically restricted to particular historical frameworks and ecological zones (e.g. they would seem to have been commonplace on the Southern Eurasian steppe; on which, see Mallory 1989; Anthony 2007). In addition to a spread of language there is also a spread of linguistic heritage such as words, structures, sayings, proverbs and other genres of folklore, etc. This kind of spread involves different types of contacts between speech communities that cause language change. It is a widespread phenomenon in the history of languages that a large amount of vocabulary, entire morphosyntactic structures and whole conceptual realms are transmitted from one language to another in a process 66 taxonomy, palaeo-linguistics and language contact studies need to be calibrated to the sources and data from other historically oriented disciplines. In the following, each of these methodologies is briefly described for the purposes of understanding the map presented here. The core methodology underlying the specific methodologies mentioned above is from historical-comparative linguistics (cf. Anttila 1988; Fox 1995). simplify somewhat, the words and elements of speech that spread in horizontal networks are typically labelled as cultural borrowings and those words that spread in vertical networks are labelled as substrate borrowings. The spread of lexical and structural features is typically associated with horizontal human networks of cultural contacts such as trade, taxation, the spread of technologies, artefact types, religious practices, use of raw materials, etc. The spread of areally bound linguistic features tends to be associated with language change in a particular region, often associated with the cultural assimilation of the incoming groups or indigenous population of the area. However, typically the words denoting local features of geography do not spread outside a particular region, and the same is also true of words for flora and fauna unless some cultural innovation makes them, for example, a commodity in trade. In such cases, it is not entirely uncommon that a toponym or a local term turns into a widespread cultural borrowing (cf. the word copper that originates from the toponym Cyprus, or bronze that originates from the denomination of the city of Brindisi, Italy). Those elements employed as a mother tongue in inter-generational transmission, or those elements learned from the prevalent social environment, represent linguistic continuity, i.e. languages as semiotic systems that, however, are a subject of constant evolution. Language, Community and Ethnicity Linguistic areas in the present are the outcome of linguistic areas of the past. These areas have traditionally been pictured using maps, although such mapping is notably problematic. First and foremost, a map often conceals the fact that there may be remarkable social and situational variation within a single linguistic area. For instance, it is not uncommon in Europe that city dwellers have spoken a different language than the surrounding countryside, or that an upper class and ‘the masses’ have belonged to different linguistic groups. It is also commonplace to have communities that use different languages for different purposes. For instance, the literary languages used in many communities differ from their spoken languages, code-switching may occur in multilingual communication, etc. There have been many areas around the globe where the linguistic map has, even traditionally, been extremely complicated. In Europe, such areas have included, among others, Transylvania and most of the Balkans (especially Bosnia and its surroundings), the Volga region, and so forth. Second, there may be notable bi- or multilingualism across large areas. Scholars are just becoming aware of the fact that language communities have been notably more multilingual in the hunter-gatherer past of mankind than in the modern and agricultural present, where language communities are larger and inequality between the members of the community may be much greater (cf. Saarikivi & Lavento 2012). Furthermore, picturing, for instance, the area around the Gulf of Finland as a region of Finnic minority languages (Ingrian Finnish, Izhorian and Votic) conceals the state Historical Language Mapping The orienting principle of the descendanthistorical reconstruction methodology is that each step of the reconstruction of the historical past must be based on the reconstruction of the more recent past. The methodology is thus a step-by-step approach to the past with a priority given to direct linguistic data (rather than e.g. archaeological speculation of the languages related to groups and types). Needless to say, the interpretation of evidence must comply with models for understanding the relationship between language spread and language change For drawing a prehistoric linguistic map of a particular region or language group, methodologies of areal linguistics, linguistic 67 of affairs that, in fact, the whole region is Russian-speaking and the individual speakers of minority languages only speak Finnic languages in very restricted social networks. What is more, the linguistic maps of most of the regions of the western world have been notably blurred by recent urbanization that has caused the relocation of the large swathes of the rural population to cities and the extinction of many traditional rural communities. Regarding new urban communities, sociolinguists sometimes describe them as ‘superdiverse’ (Blommaert & Rampton 2012), meaning that many communities exist simultaneously in one geographical space and employ very complex social networks. seem to have been in many cases linguistically and ethnically fairly neutral – at least as far as this can be assessed on the basis of historical documents. It needs to be noted here that the Russian state emerged especially from diverse Finno-Ugrian and Slavic groups under the rule of a Scandinavian nobility (S1) and that the later state of Sweden appears to have emerged from a multicultural transBaltic maritime network (G1). With such prerequisites, it can be noted that the maps of language areas have not lost their value even in research today. Such maps must, however, be read with care. It is necessary to make an effort not to project linguistic-cultural groups of the present onto the past, nor to think that the groups indicated on the map have been monolingual or lived in a community that resembles a modern language community. The information presented by such maps is best understood in light of the analogies of the particular cultural area found in cultures of similar social structures and livelihoods that have been documented in the ethnographic record. From the point of view of hunter-gatherer cultures, there are many parallels in Northern and Southeast Asia, South America or Papua New Guinea. For purposes of understanding Northern European prehistory, Siberian and Northern American analogies are probably the closest. Iron Age agricultural societies may not be so close to these analogies, but there is more evidence in the historical record for some of these cultures, their ways of life are more observable in the archaeological record, and these types of data enable much more sensitive and sophisticated comparisons with different ethnographically described cultures in order to model relationships of technologies, livelihoods, social structures and so forth reflected in the evidenc available. Map 1. The current distribution of Finnic languages. One more word of caution is needed regarding the relationship between ethnicity and language. In many contexts of presentday Europe, there is a deep-rooted connection between a perceived ethnicity and the mother tongue of an individual. We may be sure that language was also an ethnically meaningful factor in many historical contexts, as is revealed, for instance, in numerous written accounts identifying people according to their language (e.g. in the Russian Primary Chronicle). Nevertheless, it is also clear that the borders between ethnicities were, in many cases, not confined to those of the linguistic groups (as far as these can be seen as distinct). Medieval and later state formations Taxonomy The languages known today that belong to either the Indo-European or the Uralic language family are, in most cases, offspring of at least two levels of proto-languages that can reasonably be differentiated. Such a twolevel structure of the language family is easily perceived in the Finno-Ugric languages, where there is little written record of older phases of the languages. Thus, Finnic 68 languages and other relevant information (discussed below), and then, on the basis of locating and dating several such intermediate proto-languages, to proceed to reconstruct the Proto-Uralic and Proto-Indo-European protolanguages (such a methodology has been followed, among others, by Mallory 1989 and Anthony 2007, but never for Uralic languages). In order to achieve such a goal, one needs to take into consideration that within a group of languages such as Germanic or Finnic there are many clues to be found that can shed light on the question of how the Proto-Finnic or Proto-Germanic unity evolved into the branches and languages of the present. Thus, it is possible to propose the order of sound shifts and vocabulary layers of the individual languages and, on the basis of such information to find out which language forms were the first to separate from the ProtoFinnic or Proto-Germanic unity (cf. Heikkilä 2014a). In the Germanic group, we also possess prerequisites for a much closer inner reconstruction of the process of the disintegration of the proto-language, since there are numerous dialects and a rich amount of written materials that allow us to reconstruct the language development in detail on an absolute chronology of nearly two thousand years. languages, i.e. Finnish, Estonian, Karelian, Vepsian, South Estonian (Võru–Setu), Livonian, Votic, etc. go back to Proto-Finnic, an unattested but certainly real language that was spoken in the Iron Age (see below). The mutual resemblance of the languages of the Finnic group is much greater than the resemblance of any of the Finnic languages to any other language of the Uralic family. Thus, one can safely assume that all the Finnic languages are offspring of Proto-Finnic, which is in turn a descendant of Proto-Uralic. Similarly, the Sámi languages, neighbours of the Finnic groups in the North, are a group of mutually more or less comprehensible languages that form a dialect continuum and that can be traced back to a Proto-Sámi language. Proto-Sámi was in its turn also a daughter language of Proto-Uralic and a ‘sister language’ of Proto-Finnic. ProtoUralic, the predecessor of both Proto-Finnic and Proto-Sámi, as well as several other proto-languages, was a Stone Age protolanguage that is reconstructed on the basis of more recent proto-languages (i.e. a reconstruction based on other reconstructions) and, therefore, much more poorly known about in research than those proto-languages closer to us in time. A similar basic structure of language relatedness can also be found on the IndoEuropean side. Thus, for instance, present Swedish, Norwegian, Icelandic, Danish, German, Dutch, English and several minor languages such as Frisian, Elvdalian, Faroese, etc. are all Germanic languages, and daughter languages of Proto-Germanic that is in turn a daughter language of Proto-Indo-European, from which also Proto-Slavic, the predecessor of Modern Russian, Novgorod Slavic, Pskov Slavic and so on derive. Several other major groups of languages are branches of the IndoEuropean family (Romance, Celtic, Greek, Indo-Iranian languages, etc.) and thus derive from the same proto-language. From the point of view of linguistic taxonomy, it is a reasonable goal to develop a series of linguistic maps representing the reconstructed prehistory of the areas of the intermediate protolanguages (Proto-Finnic, Proto-Sámi, Proto-Germanic, Proto-Slavic, etc.) on the basis of the distribution of modern Areal Linguistics Every language has variation that, in past agricultural, nomadic and hunter-gatherer societies, has been overwhelmingly areal (in modern large speech communities, social variation may be at least as important). In dialectology, it has been found that language areas are typically more conservative at their periphery than in their central areas. This is because many linguistic changes from the centre do not reach the peripheral dialects that are geographically most remote and historically the first to lose contact with the core language area. Of course, there are also linguistic innovations taking place in the peripheral language forms that typically have different contact languages than those dialects of the core language area, but these innovations are likely to be confined to the periphery. Thus, the area with most of those innovations shared with a number of individual branches of a language family will 69 Figure 1. A Finnic language family tree (source Kallio 2014: 163). typically be closest to the historical core area of the proto-language. Therefore, from the point of view of locating the old significant areas of proto-languages and their contacts, it is important to identify those variables of the present languages that point to the oldest splits of the proto-languages. This usually can be done with the help of a methodology that investigates the phonemic contexts of the established sound shifts, especially in borrowed vocabulary (some words of which have undergone a particular sound shift whereas others have not; cf. Heikkilä 2014a). With the same methodology, it is possible to establish a family tree for the closely related languages. This has been done several times for the Finnic languages, most recently by Petri Kallio (2014), who also discusses earlier taxonomies (see Figure 1). As is clear from his investigations, the core area of the Finnic languages has been south of the Gulf of Finland, since the isoglosses that distinguish groups of present-day languages are most numerous and oldest here.3 What is more, several of these isoglosses would seem to split the Estonian language into two historically distinct language forms: Northern Estonian (or the Estonian literary language) and Southern Estonian (Võru-Seto). Sometimes only the historical peripheries of an earlier, much larger language area may remain after the rest of it has disappeared. The Celtic languages, for instance, were spoken in most of Central Europe but now survive only in the British Isles and Brittany, at the northwestern periphery of the European continent, an area nowhere near the presumed old homeland of the Celtic languages. As for the Sámi languages, a similar process would seem to have occurred. There is a firm evidence that Sámi was spoken in most of what is now Southern Finland (cf. Aikio 2007), but the language is only preserved in Northern Fennoscandia (S). In a future article of the present series, an early split into a southern group (Southern and Ume Sámi) and other groups (central and eastern) is proposed with the assumption that the predecessors of the languages of the southern group would have been carried from southern Finland to Central Scandinavia across the Gulf of Bothnia during the Iron Age (Häkkinen 2010). Palaeolinguistics A classical, and often probably over-stressed methodology for locating and dating protolanguages is that of the palaeolinguistics, i.e. reasoning based on the vocabularies of the reconstructed (or otherwise attested) early forms of the language. Thus, from the point of view of dating Proto-Finnic, it is of crucial importance that it has vocabulary related to metal (rauta [‘iron’], kulta [‘gold’], hopea [‘silver’], miekka [‘sword’]), sedentary 70 Map 2. Distribution of Finnic-speaking groups and their neighbours ca. AD 1000. Circles indicate roughly approximated areas of Uralic language groups (dashed lines indicate uncertainty in the identification); squares indicate the same for Indo-European language groups. Stars represent centers that existed in ca. AD 1000; dots indicate corresponding sites that are only probable. F = Finnic tribes; B = Baltic tribes; Sv = Slavic groups and activity areas; G = Scandinavian groups; E = Other (extinct) Finno-Ugrian groups; S = Sámi groups; MG = Mobile groups (the languages of which are unclear). Specific groups are listed by number below. chiefdom society (kylä [‘village’], kuningas [‘king’], linna [‘fortification’]), agriculture (pelto [‘field’], ruis [‘rye’]) and cattle breeding (lehmä [‘cow’], porsas [‘swine’]). From the point of view of locating the protolanguage, it is equally significant that there is maritime vocabulary (laiva [‘ship’], purje [‘sail’], meri [‘sea’]). None of this vocabulary was present in Proto-Uralic, which had a Neolithic or early Palaeolithic character with a vocabulary pointing to nomadic ways of life, a hunter-fisher society with minimal social stratification and no elaborate agriculture or use of metals (cf. Häkkinen 2007). As for Proto-Sámi, a recent discussion on the relevant cultural characteristics has been provided by Ante Aikio (2012). 71 Although Scandinavians seem to have played a fundamental role in the opening of the Eastern Route, polities that developed through those trade networks should be assumed to have been multilingual and they were almost certainly not dominated linguistically by the Scandinavians ca. AD 1000. Groups of Scandinavians settled along the Eastern Route and Scandinavian culture seems to have been dominant in establishing such polities in the 8th and 9th centuries, which then evolved as distinct cultural arenas within the developing multicultural networks (e.g. Callmer 2000; Duczko 2004). Silver trade along the Eastern Route was disrupted in the mid-10th century, and Birka in Central Sweden as well as other trade centres eventually collapsed. A subsequent restructuring of Scandinavian trade networks is reflected in silver beginning to arrive primarily from the west rather than the east. Political as well as trade networks between Scandinavia and especially Novgorod remained vital, but the polities had evolved from satellites of Scandinavian trade into centres of political and economic power. The role of Scandinavian languages in the major polities to the east can be deduced to have declined with the restructuring of trade networks toward the end of the 10th century, leading to the clear impression of the dominance of Slavic dialects in these polities by the time of their written sources from the 13th century. The lively contacts especially connected with nobilities indicated in Old Norse kings’ sagas and elsewhere allow the inference that Scandinavian language was still current in these networks in ca. 1000. Nevertheless, these polities are grouped as predominantly Slavic language areas on Map 2, noting that the actual language situation remains obscure. No attempt will be made to locate and differentiate all Germanic language groups west of the Baltic. As far as is known, Finnic language areas did not extend to the western side of the Baltic in the Viking Age (with the possible exception of the Kven/Kainuu population along the north of the Gulf of Bothnia, on which see F18). It is possible that small Finnic speech communities could have been established in conjunction with transBaltic maritime mobility which was certainly Layers of borrowings In a similar manner to the semantics of the proto-language vocabulary, the origin of the vocabulary is also of interest from the point of view of dating and locating the protolanguage. Much of the vocabulary that points to culturally more developed characteristics of Proto-Finnic and Proto-Sámi speech communities in comparison with Proto-Uralic speech communities derives from IndoEuropean, most notably, Baltic and Germanic. Proto-Finnic has had direct contacts with both of these languages, while Proto-Sámi has likely had direct contacts only with ProtoFinnic. Since the existence of such large layers of borrowings is beyond doubt, one has to reconstruct the areas of Proto-Finnic and Proto-Sámi so that they have been adjacent to Proto-Germanic (cf. Saarikivi 2009; Aikio 2012). Proto-Finnic must in turn have been in touch with the early sources of Slavic vocabulary. Proto-Sámi likely did not have direct contacts with either Baltic or Slavic (cf. Aikio 2012) but the common predecessors of Sámi and Finnic had very early contacts with the satemized Indo-European languages including both Indo-Iranian and Baltic languages (Joki 1973; Holopainen 2014). These contact histories will be addressed more fully in later articles of the present series. Germanic Groups Scandinavians seem to have played a fundamental role in the opening of the austrvegr or ‘Eastern Route’, which passed first through areas settled by Finnic-speaking peoples. Changes in trade activity are apparent already in the 8th century (Ahola & Frog 2014: 40–44; N.B. – language contact between Germanic and Finnic has roots much older than the Viking Age, as later articles of the series will discuss in more detail). This anticipated the opening of trade networks trading especially northern furs for Islamic silver around ca. AD 800.4 The Eastern Route seems to have been maintained centrally by multi-ethnic networks of Scandinavian, Finnic and Slavic language groups, although the relative prominence of those languages seems to have varied. 72 not unidirectional.5 Viking Age contacts with most Scandinavian groups west of the Baltic seem to have been more limited, as were contacts with other more distant Germanic groups engaged in maritime networks on the Baltic. Only three Germanic groups will be introduced here for their significance in contacts with Finnic groups and these groups will be introduced with emphasis on those contacts. of Finland. The route to the Gulf of Riga would cut to the south between the Estonian mainland and the large islands of Hiiumaa and Saaremaa. A legendary history of early kings of the Svear also reports the subjugation of Finnland by the Svear across the Baltic (Ynglinga saga 22; see also Aalto 2014) and the later defeat of one king in a battle in Estonia (Ynglinga saga 36). Archaeological evidence supports the image that the Svear were a significant source of Germanic cultural influences for Finnic groups as well as for other groups addressed in this article (e.g. Salo 2000; Tvauri 2012: 28). The subjugation of other polities in the centralization of power around the king of the Svear gradually linked to the Christian Church and eventually manifested the state that became Sweden. The early extension of Svear power involved compelling other polities to acknowledge a political and ritual centre, taxation and also the hundare system and associated ledung system of conscription for support in military campaigns. They did not impose an administration for local governance and laws on other polities, limiting direct local interference. The development of the hundare and ledung systems and their forms in different periods are difficult to estimate (Salo 2000; Line 2007). The question is relevant because if a Svear system of military conscription was extended to territories across the Baltic (cf. F13), Finnic groups could potentially have been involved in the organization of multi-ethnic fleets, although it has also been hypothesized that conscription would have been compensated by taxation in places like the Åland Islands (Heininen et al. 2014b: 340–341 and works there cited). Svear authority undoubtedly reached polities of Finland, but it is unclear what form this took.6 It has been argued that the geography of Swedish dialects in recent times generally “reflects the political and demographical situation of the Viking Age and Early Middle Ages quite strikingly” (Rendahl 2001: 170). The Sveamål dialect area, which is the eastern-most in Sweden proper, mostly corresponds to those East Norse language areas that became subject to the Svear’s ledung (Rendahl 2001: 141). This suggests a G1: The Svear The Svear are well attested in numerous historical sources. The first probable attestation of the ethnonym is Tacitus’s mention of the Germanic-speaking Suiones, said to be powerful at sea (Germania 44; ca. AD 98). Probable references can be found through later sources and are reliably identified in medieval texts as Old Norse Svíar, Old English Sweonas and Latin Sueones (Valtonen 2008). Old Norse medieval sources in particular represent the Svear as a major sea power in the Baltic during the Late Iron Age, emanating from Central Sweden and associated with the major trade centre of Birka. Their political expansion included the subordination of other polities on the Swedish Peninsula and the island of Gotland, even if the process and precise polities remains unclear (Peel 1999: 6–7; Line 2007: 35ff.). They extended their authority across the Baltic Sea and seem to have subordinated polities in Latvia and Lithuania during or prior to the Viking Age (Valtonen 2008: 137–138; also Tvauri 2012: 28; cf. B1). The Svear were also active in opening the Eastern Route and likely formed the centre of the multi-ethnic network that emanated from the major trade centre commonly called Staraya Ladoga (Sv5). This centre was founded in a region that, by that time, was likely dominated by Eastern Finnic tribes that subsequently evolved into the Vepsians (F9). The main sailing route from the central territory of the Svear across the Baltic was from the area of Roslagen in Central Sweden via the southern parts of the Åland Islands and the archipelago to the south-western tip of Finland. From there, the sailing route to Staraya Ladoga could pass along the northern coast of Estonia, although they could also follow the largely uninhabited southern coast 73 historical connection between the Svear kingdom and the development of this dialect group from East Norse. The Svear also seem to have been central in the later spread of dialects of Östsvenska mål to coastal areas of Finland by the end of the 12th century and to Estonia from the 13th century (Rendahl 2001: 147, 153).7 The etymology of the ethnonym Svea is generally seen as originally an endonym deriving from a reflexive pronoun meaning something like ‘our own (people)’ (e.g. Brink 2008: 60). In practice, the term likely varied its referent by time and by context from referring to members of a specific tribe, polity or local culture to referring to members of a multiethnic confederation. The ethnonym Gutar is likely of the same origin and similar to those of a number of other Germanic groups (Geats, Goths etc.). This makes it difficult to distinguish references to Gotlanders in Classical and early medieval sources. The number of different groups identified with this ethnonym has been interpreted as a sign of its great age, but its etymology remains unclear (e.g. Lehmann 1986: 163–165). The ethnonyms have been considered derivatives of a verb meaning ‘to pour’, reconstructed to *gautaz [‘out-pourer’] (e.g. Andersson 1996; cf. de Vries 1961: 159), but the etymology cannot be considered certain as the implied object of the verb remains unclear.8 G3: Ålanders Viking Age Ålanders and their polity or polities are completely absent from the historical record (on which, see Heininen et al. 2014b: 342–343). Several major toponyms of the islands relevant to navigation may be datable to the Viking Age or perhaps somewhat earlier. These appear to be predominantly of Scandinavian origin.9 The production of such toponymy may be directly connected to increased sea traffic between the trade centres of Birka and Staraya Ladoga in the Viking Age rather than necessarily reflecting place names used by local inhabitants of the islands. (Schalin with Frog 2014.) A riddle of Åland has been the lack of toponyms that are both dependent on continued use by inhabitants of the islands and can also be reliably dated to before the Scandinavian colonization of Åland and coastal Finland beginning from the 12th century.10 This apparent discontinuity in toponymy has been interpreted as a general discontinuity of settlements.11 However, no systematic search for Uralic substrate toponymy has been conducted in Åland. The probability that a Scandinavian dialect was a dominant social language in Åland is inferred based on archaeological evidence. At the beginning of the 6th century, Åland appears to have been culturally linked to cultures of mainland Finland, but probably substantial immigration from mainland Sweden later in that century caused a shift in culture and population expansion in which the indigenous groups were assimilated. The G2: Gotlanders The Gotlanders are attested in diverse historical sources, of which the most central is the locally produced vernacular compendium Guta lag [‘Laws of the Gotlanders’] that is prefaced by Guta saga [‘The Saga of the Gotlanders’], a saga of the origin of their society. These are dated to ca. AD 1220–1275 (Peel 1999: lii–liii; 2009: xxxix). The population of Gotland played a central role in Baltic Sea trade throughout the Iron Age. Gotlanders were especially active on the Eastern Route: more coins from the 8th–12th century have been found in Gotland than in the rest of Fennoscandia combined (Talvio 2014: 134). The polity of Gotland was subjugated by the Svear, apparently in the Viking Age if not before (Peel 1999: 6–7). Guta saga reports an emigration from Gotland to “Dagaiþi”, today’s Hiiumaa / Dagö, one of the major islands off the coast of mainland Estonia (cf. F2), where they established a fortified settlement before travelling the Dvina (Daugava) River opening the Eastern Route (Peel 1999: 4–5). Contacts with Gotlanders were more significant for Finnic areas to the south of the Gulf of Finland (Tvauri 2012; cf. Salo 2000). The language of the Gotlanders likely constituted a distinct branch of Norse by the Viking Age (Palm 2004: 329). In light of the Gotlanders’ central position in Baltic Sea trade, direct lexical borrowing from Gotlandic in to Finnic may well have occurred, but the issue has not yet received a scholarly treatment. 74 through Kiev to the south was thus referred to generally in Old Norse as Garðar, the plural of garðr [‘yard; fenced or walled area; a property; town; realm’; cf. Russian gorod ‘town’ < Proto-Slavic *gradъ < *gardu that is, ultimately, the same word], which was later commonly replaced by the secondary formation Garðaríki [‘Realm of Towns’]. As discussed above, Scandinavians seem to have initially been at the centre of the establishment of many of these polities. As will be discussed below, the same phenomenon seems to have involved the mobility of Finnic-speaking groups and the spread of Finnic dialects especially to the east and north. The chronology and mechanisms of the Slavicization of the region are subject to debate.12 However, the mid-10th century disruption of the silver trade seems to have played a key role in this process. In the Middle Ages, the North Russian region was subject to rule by several emerging Slavic states. These developed around the most influential towns that competed for the domination of the surrounding countryside and especially for control of northern regions where furs, waterways and people could be taxed. Most notably, Staraya Ladoga, Lake Beloye, Novgorod, Pskov, Rostov, Suzdal, Vladimir, Polotsk and Smolensk can be mentioned. The republics and principalities that developed on the basis of such towns had fluctuating borders and multi-ethnic populations. By the end of the 14th century, Novgorod was just about the only one of these that was still independent. By the 15th century, all of the Slavic states were subjugated to Muscovite rule. The gradual Slavicization of the area between the town centres has been investigated by Nikolaj A. Makarov (1997) on the basis of archaeological material from the watersheds between the major Northeastern European rivers. In the material assumed to be associated with the places where people portaged from one water route to another, the local artifact types associated with the FinnoUgrian groups are replaced during the Middle Ages by, for instance, turned ceramics associated with southern Slavic centres. From the perspective of linguistics, a central question regarding the early Slavic immigration process can be viewed in the broader context of settlement expansion and change in regions of the Svear at that time (Line 2007: 39; cf. also B1, F13). The changes in burial practices and settlement patterns led fairly rapidly to a distinct and remarkably uniform cultural area that can be archaeologically classed as ‘Scandinavian’ (e.g. Callmer 1994; Gustavsson et al. 2014). The comprehensiveness of discontinuity from the indigenous culture suggests that changes extended to, for example, systems of laws and social behaviours (Heininen et al. 2014b: 338). A Scandinavian dialect likely became socially central (Ahola et al. 2014a). Later evidence suggests that the islands were viewed from the Swedish mainland as linked to the archipelago or mainland Finland rather than as part of Svealand proper (Sjöstrand 2014: 120–123; also Heininen et al. 2014b: 340–341). Especially north-eastern Åland exhibits intensive contacts with polities of Finland (Gustavsson et al. 2014: 179–182; Heininen et al. 2014b: 334–339). Thus Åland became a kind of a frontier zone for early Finnic–Germanic language contacts. It can be assumed that Åland would become a distinct Scandinavian dialect area in the centuries leading up to AD 1000. The etymology of Swedish Åland, first attested in 13th century sources, and its possible relationships to the corresponding Finnish toponym Ahvenanmaa, remain disputed. The theory that the name goes back to a Proto-Germanic loan into Proto-Finnic later reborrowed into Scandinavian (Schalin 2008) has gained attention. On the other hand, it has recently been pointed out that the name may date to the Viking Age with a naming basis of topographical features relevant to seafaring along the Eastern Route (Schalin with Frog 2014: 289–297). Slavic Groups Around AD 1000, the Slavic areas southwest of the Gulf of Finland were far from uniform, either culturally or linguistically. There were several centres of Slavic habitation. They formed a number of local, multi-ethnic polities along the Eastern Route in order to seek economic control over those networks and over one another. The whole vast region 75 populations has been the character of some Northern Russian language forms, most notably of the Novgorod and Pskov vernaculars, as intermediate language variants between West Slavic (Polish, etc.) and East Slavic proper (Russian literary language). Although the forms of Slavic in this region are presently classified as Russian dialects, they are still substantially different from Russian literary language or East Slavic proper (Zaliznjak 2004). There is firm evidence of early contacts between Finnic and Slavic languages that have taken place already before fundamental Slavic phonological changes such as the disappearance of nasal vowels and reduced vowels, or polnoglasie.13 This can be considered to have taken place during the Iron Age, although dating on an absolute chronology is unclear. From a Finnic perspective, the Slavic borrowings form the youngest layer of the Proto-Finnic vocabulary, although in a few cases the boundary between Slavic and Baltic borrowings may appear slightly blurred due to the fact that they evolved from a common antecedent and in a sense, owing to some common features rooted in the Balto-Slavic period, the Slavic languages represent a “further developed Baltic” (Kortlandt 1977). Significant Slavic language contact with Finnic seems to correlate chronologically with the opening of the Eastern Route or the following period and Finnic speakers borrowed their core vocabulary related to Christianity via these contacts (Kallio 2006b: 156–157). On the other hand, the ethnonym for the Scandinavian group called Rus’ as well as many for Finnic groups appear to have been borrowed into Slavic via Finnic rather than via Scandinavian dialects (Sv1 and Finnic groups below). 2009: 160–161). According to the Primary Chronicle, Scandinavians subjected tribes of the region to pay tribute in 859, and the entry for 860–862 reports that these Scandinavians were driven out but, because the different cultural groups were divided by conflicts and in-fighting, they invited princes of the Scandinavian Rus’ to come and rule them. This was done by three brothers who ruled from Novgorod (or more probably Rjurikovo Gorodišce), Lake Beloye and Izborsk, but two of the brothers died within two years and control of these areas purportedly then fell to Novgorod. The request to be subordinated by a Scandinavian rulership is, without doubt, a retrospective authorization of a more aggressive conquest.14 Historians today generally agree that the Scandinavians had a significant role in early Novgorod as an upper class, although many details remain obscure. The Hypatian codex of the Primary Chronicle presents Staraya Ladoga in the place of Novgorod in the account of the three Scandinavian rulers of the Rus’, and it describes how the king subsequently founded Novgorod [Ru. ‘New Town’]. The logic of this account may be deceptive. Novgorod is known in Old Norse as Hólmgarðr [‘Fortified Island’]. This designation seems most likely to have referred (at least originally) to Rjurikovo Gorodišce, the site of which became an island seasonally (Price 2000: 265; cf. Callmer 2000: 38), although it seems in Scandinavian sources to refer to what is later understood as the political entity of Novgorod. Novgorod became the most prominent political and economic centre of North Russia across the Viking Age, apparently extending authority over the nearby proto-urban centres of Staraya Russa, Staraya Ladoga and Lake Beloye that had been founded already quite early. It seems that the Novgorodians were also oriented to assert control to the south. It is possible that the chronicles could exaggerate the position of Novgorodian princes in what may have begun as some sort of politico-military confederation (cf. Hedenstierna-Jonson 2009: 166), but the outcome was control of the silver trade to the north and from there to Scandinavia. Novgorod developed into a major economic and political power in the east. According to Sv1: Novgorodians What we refer to as Novgorod emerged as a major trading centre in the 9th century. In 13thcentury historical sources, Novgorod seems to be conflated with Gorodišce or Rjurikovo Gorodišce. The latter pre-urban site with a fortification was established on the north of Lake Ilmen and seems to have exerted control over the trade route to the north and developed a pronounced Scandinavian presence with resemblances to Birka (Hedenstierna-Jonson 76 cf. Fi. Ruotsi [‘Sweden’], Ruotsalainen [‘Swedish’]).15 This pattern of borrowing would be consistent with the number of other ethnonyms borrowed from Finnic for groups along the Eastern Route. The term has been considered likely to be borrowed from the first element of a Scandinavian denomination roþs-karlar or roþs-män [‘row-men’]. Although a final consensus has yet to be reached concerning the precise etymology of the Finnic ethnonym (cf. Anderson 2007), it is generally agreed that the Finnic term is of Scandinavian origin for the group with which they first had significant contacts, somehow associated with the toponym Roslagen (cf. Brink 2002: 765–766), and most likely initially referred to the Svear (e.g. Andersson 2007: 7). The Norse ethnonym Ruz (Ryz in eastern dialects), found referring only to the Rus’ and not to people in Scandinavia (cf. DONP, s.v. ‘ruz’), seems in its turn to be a loan derived from the use in Slavic. the Russian Primary Chronicle, Christianity was adopted there in AD 989. The city became a significant population centre during the Middle Ages, with as many as 40,000 inhabitants mentioned in some sources. From the mid-12th century, it functioned as an independent principality that colonized large northern territories, most notably those around Lake Ladoga and Lake Onega as well as the Northern Dvina river basin. According to a historical document written when Novgorod emerged as a sovereign principality (the Ustavnaja gramota knjazja Cvjatoslava Olgoviča), northern parts of the Dvina basin were already under Novgorod’s control in 1137 (cf. Nasonov 1951; Makarov 1997: 18–20). The population of those areas was certainly overwhelmingly non-Slavic. In addition to those exhibiting similarities to the area populated by the Finnic tribes (F15–17), some, such as those in the Kokšen’ga and Sukhona basins, had intensive contacts with groups of the Upper Volga region (Ovsyannikov 1980; Rjabinin 1997; Kolpakov & Ryabtseva 1994). (Saarikivi 2006: art. 2, pp. 7–8.) The language of Novgorod has been preserved well thanks to multiple birch bark documents (cf. Zaliznjak 2004). These also preserve the oldest texts in Finnic languages (Laakso 1999). The dialect spoken in Novgorod and its colonized areas exhibits several archaic or West Slavic characteristics that deviate notably from the Russian literary language of later times. In addition to Slavic, there seems to have been a wide array of Finnic and probably other West Uralic language speakers residing in Novgorod and adjacent areas. This is, among other facts, visible in that a quarter of Novgorod was referred to as Čudskij konec [‘Čud’ Part’] (on Čud’s, see E1 below). In Old Russian sources, the inhabitants of Novgorod are referred to as the Rus’, which is in turn the ethnonym for the ruling Scandinavian group, and which has become the present ethnonym for Russians today. This early ethnonym presents a complex problem. The Slavic form is now widely agreed be a borrowing of a Finnic ethnonym for Scandinavians or, more specifically, for the Svear (later attested in all Finnic languages; Sv2: Central Russian Principalities In the Middle Ages, the core area of presentday Russia was controlled by several independent principalities (Rostov-Suzdal’, Vladimir, Jaroslavl’) that competed with Novgorod in the colonization and taxation of the Northern Finno-Ugrian regions. These principalities collapsed under the emerging Muscovite rule earlier than Novgorod. Nevertheless, they were significant players in the ethnohistorical map of the 11th–13th centuries. In the dialects of the Arkhangelsk region (cf. F15–17, E5), two waves of Slavic influence can be discerned corresponding to two different colonization centres, those of Novgorod and Rostov-Suzdal (Komjagina 1997). The opening of the Eastern Route into Central Russia during the 9th century is indicated through especially Scandinavian presence (Duczko 2004), but even at this stage the development should not be assumed linguistically homogeneous. For example, Timerëvo (superseded by the founding of Jaroslavl’ a few kilometers away in ca. AD 1000) seems to have developed as a multiethnic polity that was likely characterized by a collective identity (Androshchuk 2008: 523). It may be assumed that the area was originally inhabited by the Meryans (E2), but 77 the prominence of Scandinavian presence in the first half of the Viking Age was followed by a shift to Slavic during the Middle Ages. Similar Scandinavian burial customs have also been documented at Lake Beloye (Makarov 1993; cf. Sv4), suggesting they were part of the same contact network (cf. Ahola et al. 2014: 255, 260). A distinctively Ålandic (G3) funerary rite appears along the Eastern Route only in Timerëvo and the surrounding area, where it was adapted into a local form that became remarkably prominent, suggesting a different ethnic and presumably dialectal Scandinavian component.16 The constellations of groups forming these polities were not uniform and they developed on a local basis leaving the relative prominence of languages and dialects uncertain. Among these principalities, names for two areas are attested in Old Norse. Rostov appears as Rostofa, which does not appear to be of Scandinavian etymology, but a Slavic etymology is also not clear. Suzdal’ appears as Súr[s]dalar and Suðrdalaríki [‘Realm of the South Valleys’]. The element dalar (gen. dala) is the plural of dalr [‘valley, dale’] and would be unsurprising for an Old Norse place name. The first element is uncertain (cf. súrr [‘sour’]) and the attested forms could reflect folk-etymologization of a non-Norse element (Valleys of X) or of a fully foreign toponym. The toponym could also be a Slavic formation, in which case it would mean ‘founded by Suzda’ (in which the personal name would not appear to be Slavic). linguistically Slavic in the Middle Ages, as proposed already by Harri Moora (1982). The overwhelmingly Slavic settlement names of the Seto area (Kiristaja 2013) seem to corroborate this hypothesis. Sv4: Lake Beloye (Beloozero or White Lake) Lake Beloye is one of the three centres of political control identified in the foundation myth of the Rus’. Archaeological evidence points to a strong Scandinavian presence (identified through types of clothing and weapons) in this area beginning from the 8th and 9th centuries (Makarov 1993). However, references to it in Old Russian chronicles suggest that it became integrated into the political sphere of Novgorod. The Primary Chronicle mentions that people called the Ves’ resided in this area and it has thus been customary to interpret the Finnic peoples in the vicinity of Lake Beloye as Vepsians (F9). Lake Beloye is by now entirely Russified linguistically, but not far away from it there are Vepsian settlements in the west of the Vologda region even today. Since this is the only area customarily considered as Finnicspeaking in the Volga basin (otherwise the Finnic languages are only spoken along the waterways that flow to the Baltic Sea), it has been argued that it probably functioned as a contact zone between Finnic languages proper and the extinct Finno-Ugrian languages of Central Russia, most notably Meryan (E2). In the toponymy, no Scandinavian elements are discernible. There is, however, a very rich Finnic toponymic substrate and, in addition, there are probably some toponyms from other Uralic languages (Makarova 2012). Sv3: Principality of Pskov Pskov formed an independent republic up to the mid-13th century and Pskov retains its character as a dialect area up to the present day. In a similar manner to Novgorod, Pskov also presents an independent language form and cultural area that, however, represents characteristics closer to the Novgorod dialect group than to Russian literary language or East Slavic proper. The Pskov principality consisted of those lands closest to Estonia that were most likely Slavicized early on, probably motivating the breakup of the ProtoFinnic language area. The exact mechanisms of the Slavicization of this area are not known, but there are grounds to suggest that some parts of Southern Estonia were Sv5: Staraya Ladoga One of the most important centres for trade along the eastern route is now commonly known as Staraya Ladoga, known in Old Norse as Aldeigjuborg [‘Fortified Town of Ladoga’] and described in some sources as ruled by a king. Dendrochronology dates its founding to ca. AD 750 (Kuz’min 2008). This is also not surprisingly the area where the earliest Scandinavian finds along the Eastern Route appear concentrated (Androshchuk 2008: 520). The site appears to have centrally developed as a Scandinavian outpost or settlement in a Finnic cultural area. Scandinavian– 78 The main linguistic division within the Baltic groups has been that between Western and Eastern Baltic. The living Baltic languages belong to the Eastern type, and the now extinct Western languages were substantially different from them. Finnic contacts were thus initially predominant whereas the Slavic contacts became significant here later (Duczko 2004: 64). Staraya Ladoga seems to have developed into a multi-ethnic trading community that was an important site for contacts along the Eastern Route. Its status as a polity and its history of political alignments nevertheless remain obscure. The possibility that Novgorod was founded by a king of Staraya Ladoga (Sv1) is a curious possibility, but it seems more likely that Novgorod developed out of a trade centre that emerged as an independent polity. Staraya Ladoga came into the field of Novgorod’s authority at an early stage, but it is unclear when or how this may have occurred (cf. F8). Vepsian is still spoken in the upper reaches of the rivers flowing into Lake Ladoga and a strong Vepsian substrate prevails in the toponymy of the Pasha and Volkhov Rivers in the vicinity of Staraya Ladoga (cf. Mullonen 1994; 2002). B1: Curonians The Curonions are well attested in medieval sources, called Kúrir living in Kúrland in Old Norse, and Curones living in Curland in Latin, as for example in Henry’s Chronicle. They are associated with the Courland Peninsula on the eastern side of the Gulf of Riga. Although there are no direct sources for Curonian languages, some scholars (for instance, Vytautas Mažiulis) classify them as Western Baltic. Valentin Kiparsky (1939) was of the opinion that the Curonians did not form a unified group but consisted of diverse people residing on the Curonian Peninsula. Multiethnic immigration from Scandinavia established settlements in the area of Grobiņa apparently from the beginning of the 7th century, and these seem to have been or come under Svear authority in the 9th century when rebellion compeled an expedition to regain authority (Callmer 2000: 29; Ferguson 2009: 110–111). However, the fact that the region is currently Latvian-speaking and, in the north, Livonianspeaking (F1) is a strong argument for the Eastern Baltic character or (partly?) Finnic character of Curonian language forms. More arguments are probably to be found in the overwhelmingly Baltic toponymy of the region. It has also been pointed out by some Estonian scholars that in the Chronicle of Henry of Livonia, there are Finnic words associated with the Curonians (cf. kiligunden ~ Estonian kihelkond [‘administrative division’]). Baltic Groups Relevant as Neighbours of Finnic Speakers In comparison with the Germanic and Slavic groups, the Baltic groups are relatively little studied and there is also less direct evidence of their language. Finnic languages were in direct contact with the Baltic languages in an early period, and it has been proposed that the contact was of a substrate character – i.e. that the Finnic languages spread into an area around the Baltic Sea where Baltic languages were spoken at that time (Kallio forthcoming). There is an old assumption that the Baltic contacts are older than the Germanic contacts, and this is corroborated by the fact that the Sámi languages likely did not have direct contacts with Baltic (cf. Saarikivi 2009; Aikio 2012). From the point of view of comparative Indo-European studies, the Baltic languages are very conservative, and it is thus not always possible to distinguish Baltic borrowings from other early satemized IndoEuropean languages. On the other hand, Baltic borrowings are not always easily distinguished from Early Slavic borrowings, which is understandable in view of Slavic characterized as a “further developed Baltic” (Kortlandt 1977). B2: Old Prussians and Other West Baltic Tribes The Old Prussians are the only Western Baltic group whose language has been attested in writing. There are numerous literary documents in Old Prussian including catechisms, word lists, prayers, etc. and these allow a preliminary reconstruction of the Old Prussian language. It was notably different from both modern Lithuanian and Latvian. The nucleus of Old Prussian was in the area of the present-day Kaliningrad enclave of 79 Russia and Northwest Poland. The language has likely been closely related to other Western Baltic languages, most notably Sudovian and Galindian, and was spoken probably until the beginning of the 18th century when it was replaced by German. The region inhabited by the Old Prussians seems to correspond to the region inhabited by the tribes of the Aestii described by Tacitus and the later location of Estland and the Este in the 9th century account of Wolfstan (on which, see Valtonen 2008: 402–447). known, however. The Lithuanian language is traditionally considered the most archaic of the currently-spoken Indo-European languages. It exhibits several archaic features, most notably in declension and accentual paradigms, but also in vocabulary. There are two present main variants of Lithuanian, the lowland dialect (Samogitian) and the highland dialect (Aukštaitian). B5: Selonians The Selonians are a Baltic tribe mentioned in several medieval sources, including the Chronicle of Henry. The area of the Selonians was between the Semigalls, Latgalls and the Lithuanians. However, little is known about the cultural traits that may have differentiated the Selonians from these other groups. They are mentioned by Henry as the allies of the Lithuanians, and their region has likely been subject to the rule of the Slavic principality of Polotsk. The name of the Selonians is preserved in the place name Selija, denoting a region in southern Latvia. B3: Semigallians The Semigallians are a Baltic group involved in conflicts with the Scandinavians, Germans and likely also with Finnic peoples that seem to be mentioned in a number of early sources. This is unambiguous in the Chronicle of Henry, where the Semigalli of Semigallia are noted for their long-lasting resistance against the Crusaders. It is inferred that Semigallia is identical to the place(s) referred to as Samland in Old Norse, Semland in Adam of Bremen’s work and Sembia in Saxo Grammaticus’s Gesta Danorum. The name and identity of the Semigallians live on in the name Zemgale denoting a region in southern Latvia. The sources suggest that this group had close contact with Samogitians, another Baltic tribe in the territory of present-day Lithuania. B6: Latgalians An ancient Baltic tribe that has a modern offspring in the Latgalian ethnic and minority group of Latvia. The Latgalians were subject to influences both from the Teutonic Order as well as the Pskov principality. Of those Baltic tribes mentioned by Henry, the Latgalians were the easternmost. It is widely assumed that a substantial part of the former Balticspeaking area east of present-day Latvia has been Slavicized, and a number of Baltic etymologies have been put forward for Belarus and Pskov hydronyms that would bear witness to earlier Baltic habitation in this area (Trubachev & Toporov 1962). B4: Lithuanians The most influential of the Baltic tribes formed a kingdom (later a Grand Duchy) in the mid-13th century. The Lithuanian kingdom was the last major state in medieval Europe to become Christian and was involved in multiple conflicts with the Teutonic Knights, Kievan Rus’ and Poland. The heyday of its might was the 14th century, when Lithuania was one of the largest countries in Europe, extending from present-day Lithuania to the Black Sea, although only a small part of this region has been Baltic-speaking. There is enough evidence of a continuum of Baltic dialects from the region of present-day Lithuania eastwards, up to the Middle Volga area where, in the medieval period, a group of Galinds (Old Russian Goljad) resided (cf. Trubachev & Toporov 1962). The character of these language forms vis-a-vis present-day Lithuanian and other Baltic languages is not Hunter-Fisher Cultures of the North In ca. AD 1000, the majority of today’s Finland, Karelia and large parts of the Scandinavian Peninsula were inhabited by groups that had mobile ways of life, without fixed settlements. As a consequence, these groups were generally not clearly defined or described in early written sources. They are (weakly) discernible in the archaeological record, which exhibits patterns of continuous change from the Stone Age up through the Metal Period, with far-reaching networks 80 transporting metal artifacts from the Urals and beyond to present-day Finland (Makarov 1997; Saarikivi & Lavento 2012). These groups were significant participants in the economy of the north and had on-going contacts and relations especially with the North Finnic groups who progressively expanded through their areas of habitation. through Germanic groups would have been associated with speakers of an earlier form of Sámi. Ante Aikio (2012: esp. 76) has recently argued that Proto-Sámi’s intensive contacts with Germanic speakers only began in ca. AD 200 at the earliest. There is no reason to assume that Tacitus’s term Fenni was used with reference to speakers of either ProtoSámi or Proto-Finnic; it seems more probable that Germanic speakers would have used the term primarily for other mobile groups with which they had more immediate contacts (Ahola & Frog 2014: 48). In the later sources, Norsemen were clearly involved in reciprocal economic relations with Finnar. However, these sources almost invariably present a constructed image of Finnar that seems largely divorced from ethnographic realities. In Old Russian chronicles, it is impossible to determine whether some mentions of Čud’s (E1) might refer to Sámi language groups (cf. Frog 2014b: 443), and it is possible that other groups referred to in medieval literature could have been Sámi-speaking (F18). The Old Russian ethnonym Lop’ or Lopari seems not to be attested before the 14th century (first referring to ‘Lop’s and Čud’s’ in the Northern Dvina River basin), although the terms become frequent in 16th century tax registers etc. (Korpela 2008: 48). In later Swedish sources, the term Lapp should be understood mainly as a notion denoting the hunter-fisher population that paid tax in fish, whereas the agricultural population paid tax in cereals. These sources reveal multiple cases where people have shifted from Nybonde to Lapp or vice versa (cf. Barth 1969). The primary distinction of these groups by their livelihoods or ways of life makes it possible or even probable that Sámi speakers living in fixed-settlement communities would have been referred to by other ethnonyms. Although there was a cultural differentiation between the agriculturalists and the Lapps of Häme and Satakunta already in the Middle Ages (Salo 2002), the cultural Fennicization of the Sámi population continued up to the 20th century in Finnish Lapland, roughly following the change in settlement patterns. S: Sámi groups The Sámi are well attested through sources and informants of Scandinavia although the sources are highly problematic in many respects. The Sámi appear generally referred to as Finnar, adapted into Old English as Finnas and into medieval Latin as Finni. In eastern dialects of Old Norse, this term seems to shift its referent to inhabitants of Finnland (F11) and mobile groups became designated with the term Lappir (cf. Mundal 2000; Aalto 2014). The late 9th century account of Ohthere of Halogaland is exceptional for his references (in Old English) to Finnas and Terfinnas on the Kola Peninsula. The latter term can be considered to differentiate a subgroup or tribe within the broad category of Finnar, especially as Ter- seems to correspond to the later attested ethnonym for the Ter Sámi of the Kola Peninsula and a toponym of the relevant region (Valtonen 2008: 381–386; cf. SámiKld Ta’rje [‘Eastern part of the Kola Peninsula’] ~ Russian toponym Old Russian Tьrъ > Modern Terskij bereg ~ Finnish Turja, Tyrjä; SSA III: 334). Already at the end of the 1st century AD, Fenni is attested in Tacitus’s Germania (ch. 64), where it is identified with the most primitive culture of the North; the term later became an ethnos mentioned in Classical literature on geography, especially in a compound that appears cognate with later Old English Scridefinnas.17 Although the Fenni mentioned by Tacitus are commonly interpreted as speakers of Proto-Sámi (cf. Valtonen 2008: 75), later attestations of the Scandinavian term reveal it to be a broad category of culture without clear reference to language: the term is applied to mobile groups with livelihoods based on hunting, fishing and gathering without fixed settlements. Current knowledge of the spread of Proto-Sámi suggests that Tacitus is rather early to presume that the Fenni as learned 81 Southern and Ume Sámi, have been argued to have spread to their present area from Southern Finland across the Baltic Sea (Häkkinen 2010).18 In this case, these Sámi languages have backgrounds in different Proto-Sámi dialects (cf. Aikio 2012: 88–92). The etymology of Finnr (pl. Finnar) has been much debated (see e.g. Grünthal 1997). Many of the arguments are somewhat romantic and are problematized by the appearance of Fenni in Tacitus. Fenni (sg. *Fennus) suggests a Germanic form *Fennaz or *Finnaz already in the 1st century or earlier (depending on the source of the information, which could potentially have been a written work). If Fenni is considered related to Old Norse Finnar, many proposed etymologies would require a different form in that period. Consensus on the etymology has yet to be reached. The etymology of Old Norse Lappir similarly remain obscure (Aalto 2014: 206 and works there cited). Although the Old Russian Lop’ could be a Scandinavian loan, it seems more probable that it derives from a Finnic language (Lappi [‘Lapp (language)’], Lappalainen [‘Lapp’]) as with many other ethnonyms for Uralic groups. There are approximately 1,000 toponyms derived from the term Lappi in most parts of Finland and Karelia. Many are without doubt an indication of an earlier Sámi habitation (cf. Saarikivi 2004; Aikio 2007), but it might well be that others are to be derived from lap(p)e(a) [‘side; brink’]. A borrowing from Finnic to Scandinavian would also not be surprising if Scandinavians first designated Finnic speakers in Finnland as Finnar and developed their differentiation of these groups from Lappir through those contacts. The word Sámi, in turn, derives from the Sámi endonyms Sápmelaš and Sápmi [‘Sámi area’]. These words have an ethnonym cognate in Finnish, the tribal and province name Hämäläinen, Häme (F12) that point to a south-western part of inland Finland. Many further etymologies have been proposed for these words, most of them either related to the concept of ‘even land’ (~ Baltic, cf. Latv. zeme) or ‘dark people’ (~ Proto-Germanic *sǣma-) (Kallio 1998; cf. SSA I: 207). It has been argued that there is an oral tradition related to the Fennicization of the Sámi in Notable toponymic and historical evidence regarding the Sámi is found in the Finnish Lakeland Region (F12). The toponymic data has been discussed by, among others, T.I. Itkonen (1948), Alpo Räisänen (2005) and Ante Aikio (2007). However, no up-to-date monographic treatment of the topic has been published so far. In addition to toponymy, there are multiple historical accounts of lappalaiset [‘Lapps’] in the Finnish Lakeland regions and in the Finnish forests. Corresponding Sámi substrates are observable in Karelia from western and northern areas around Lake Ladoga up to the White Sea (Kuzmin 2014). In the Northern Dvina River basin, there is toponymy that has been considered as Sámi (Matveev 2001–2007; Kabinina 2012), but it has also been argued that it rather represents an extinct branch of Uralic that shared features with both the Finnic and Sámi languages (Saarikivi 2004). In the later Swedish evidence, there is every reason to believe that the Lappar or the inland hunter-gatherers consisted mostly of the speakers of Sámi languages and their spread would seem to be similar to the area of Sámi toponyms. The emergence of the Finnic-speaking population in the Finnish Lakeland and especially Savo has traditionally been considered the result of rapidly expanding slash-and-burn agriculture, but recently Korpela (2012) has suggested that a cultural assimilation and language shift of the Sámi speakers played an important role in the process. Although it still remains common to use ‘Sámi’ (or ‘Lapp’) as though it refers to a linguistically and culturally homogeneous group, this image is artificial. Dialectal diversity is already apparent in ProtoScandinavian loans prior to the Viking Age (Aikio 2012: 77–78). When considering Sámi groups in ca. AD 1000, it warrants stressing that there was certainly linguistic diversity among Sámi speech communities at that time, even if cultural diversity may have been greater than diversity of dialects (Aikio 2012: 94; cf. Saarikivi & Lavento 2012: 200–201). In this regard, it should be noted that Sámi dialects on the Scandinavian Peninsula are not necessarily the result of a geographical extension of the language area from the north. The Sámi languages of Central Fennoscandia, 82 Upper Satakunta (the region of present-day Tampere; Salo 2000: 49) and that this would explain the use of the same ethnonym by both Finnic and Sámi people (SSA I: 207), although there is no reason to believe that such local oral history would have a timedepth of roughly a millennium. the beginning of the Viking Age at the latest (2012: 87, 106). The present authors consider the inference of extinction improbable. The ability to correlate the main period of language shift with the period of ProtoScandinavian loans allows it to be situated on an absolute chronology as transpiring in AD 200–700 (Aikio 2012: 87) or perhaps continuing as late as ca. AD 800 (cf. Schalin 2014: 405). This correlates with a period of AD 250–800 during which the lack of evidence in the archaeological record makes it impossible to assess what was happening culturally in the present Sámi area (Aikio 2012: 104). According to this model, the language shift would appear to have been completed before Germanic (via the Barents Sea) and Finnic (via inland waterways) trade routes to the White Sea region became active. It is nevertheless probable that linguistic diversity was maintained across these areas and that later language shifts of additional speech communities had little or no impact on broader dialect networks, especially once Proto-Sámi had been equipped with vocabulary to refer to relevant features of ecology and livelihoods.19 MG: Mobile Groups An ambiguous areal identification for ‘mobile groups’ appears on Map 2 to indicate regions where cultural groups practicing livelihoods based on hunting, fishing and gathering without fixed settlements were active. The movements of these groups should not be assumed free and random but probably at least in most cases followed customary geographical patterns interfaced with the seasonal livelihoods of the individual group or network of groups. The languages of these groups cannot be determined with certainty. By AD 1000, the majority of these groups would probably have spoken forms of Sámi, but the language situation to the north of clearly Germanic and Finnic language areas is highly ambiguous in the Late Iron Age. ProtoSámi seems to have spread in the centuries surrounding the beginning of the present era and across the first millennium AD and it would seem likely that also the spread of the Sámi languages to their present area is related to this process. The Proto-Sámi area has probably been somewhere around Lake Ladoga in inland Karelia and/or Finland. When Sámi began to spread, there was likely a great deal of linguistic diversity in the regions of the north. A substrate or substrates of Palaeo-European language(s) can be identified through non-Uralic (and non-IndoEuropean) features in Sámi languages, as has been extensively explored especially by Ante Aikio (e.g. 2012; see also Aikio 2004 and works there cited). There may be a similar though notably less strong substrate interference also in the Finnic languages (cf. Saarikivi 2004). Aikio (2012: 80–88) has recently offered compelling evidence for the main period in which shifts to Proto-Sámi from PalaeoEuropean language(s) occurred in northern Fennoscandia. He also takes the position that this period would correspond to the extinction of Palaeo-European languages in these regions, dating the process’s completion to Finnic Groups of Today’s Estonia Among the 13th century written sources, groups inhabiting regions of today’s Estonia are addressed in the greatest detail. These sources corroborate archaeological evidence that the Finnic polities south of the Gulf of Finland were more significant to Scandinavian groups than those farther north or east, which easily blur into the fantastic (cf. F15). For example, Óláfs saga Tryggvasonar reports that King Óláfr Tryggvason was captured as a child by Estonian Vikings when fleeing with his mother to Novgorod and thus spent six years as a thrall in Eistland (Aðalbjarnarson 1941: 230). Accounts of the conversion of the different groups of Estonia are presented in especially rich detail in the account of the Baltic Crusades in Henry of Livonia’s Chronicle, written in the early decades of the 13th century (although Henry’s work should not be confused with ethnography; Kivimäe 2011). In his treatment of Finnic peoples of this region, Henry seems to address them according 83 (terms for Livonians and Oeselians will be discussed in the relevant sections). The Old Norse ethnonym Eistr [‘Estonian’] and its Latin equivalent seem to make its first appearance in Classical sources. At the end of the 1st century AD, Tacitus refers to the Aestiorum gentes [‘tribes of Aestii’], apparently situated on the southern part of the Baltic Sea, and whom Tacitus considers culturally similar to Germanic peoples but linguistically different (Germania 46). Although the plural Aestii is normally interpreted as an ethnonym, it could also be a toponym for the place inhabited by these groups (Valtonen 2008: 74). The Aestiorum gentes appear to correspond to the Este of Estland described in the Old English account of Wulfstan from the late 9th century where the location is geographically quite precisely situated in the south-eastern corner of the Baltic Sea, identifying what was most likely a Baltic language area (Valtonen 2008: 414– 420). The toponym and ethnonym therefore seem to have been used for different cultures and places by different groups and at different times (Grünthal 1997: 215–222; cf. also F2 below). The evidence of these terms is too limited to allow a chronology for the use of the toponym and ethnonym in different areas (cf. also F12). A satisfying etymology of the ethnonym has yet to be found (for discussion, see Grünthal 1997: 213–240). However, it would seem quite likely that the ethnonym of the Estonians is the same as the denomination of the presumably Baltic tribes called Aestii or Aestiorum gentes by Tacitus. to three main categories. Most tribes or polities are described with terms related to a particular region or political area which may be discussed as having a clear division into distinct districts (provinciae) but the people are still referred to more generally as Estones [‘Estonians’] (F3–7). Old Norse Eistland, Latin Aestland and their variations are well attested. However, Henry does not refer to the Lyvonenses [‘Livonians’] (F1) as Estones, nor does he normally refer to the Osilienses [‘Oeselians’] (F2) in this way.20 This threefold division between ‘Estonians’, ‘Livonians’ and ‘Oeselians’ seems to be generally present in the western medieval sources, although its precise basis is not entirely clear. It may be preliminarily observed that Estonian toponymy, with no discernible preFinnic substrate of any kind (at least in the light of present research) suggests that all regions of today’s Estonia were predominantly Finnic language areas in ca. 1000. Estonia represents a notable number of pre-Christian anthroponyms present in toponyms and their main bulk must date from Finnic prior to AD 1000 (cf. Kallasmaa 1996– 2000; Saar 2008). The identification of groups in Estonia as Finnic-speaking is therefore fairly secure, although the area has been inhabited by various groups, none of which can be considered as the forefathers of the present Estonian ethnos in any straightforward sense. In the archaeological record, there appears a broad divide into two cultural areas: one spans from Saaremaa and adjacent coastal areas across the northern part of Estonia (F3–4, F10) and the other is a southern, inland cultural area (F5–6) (see Tvauri 2012: 323–325). These seem to roughly correspond to the dialectal boundary between Central and South Estonian. The medieval Swedish language areas of Estonia are generally thought to be the result of immigration beginning in the 13th century following on the political and religious changes imposed in the Northern Crusades (e.g. Rendahl 2001: 153), in which case these language areas emerged slightly later than the Swedish-speaking areas of Finland (but see also the view in Markus 2004). The terms designating ‘Estonia’ and ‘Estonians’ have a complex background F1: Livonians Livonia and Livonians have a central position in the Chronicle of Henry of Livonia. The toponym is known in Old Norse as Lífland (Simek 1990: 343), also found on a commemorative rune stone from Södermandland, Sweden (Sö 39), and the Livonians are referred to as Lib’ or Ljub’ in Old Russian chronicles (Grünthal 1997: 247). Some early possible references to the Livonians or uses of an equivalent ethnonym (Leuonoi/Levoni) might be found already in the Classical period, but these are problematic.21 At the time when Henry was writing, Livonians occupied the region north of the Western Dvina or Daugava River and 84 they formed a significant group along this access to the Eastern Route. The fortified town of Uexküll or Ykskylä [lit. ‘OneVillage’, although the etymology of the toponym is likely something else] formed a major centre especially in the first stages of Christianization until it was superseded by Riga. Livonians were centrally involved in the Baltic Crusades, first as adversaries and then as allies of Christianity. Livonia (and its variations) then came to designate the majority of these regions, but this should be regarded as a political outcome of the crusades which considerably expanded its territory. Livonians survived in the Salaci River basin into the early 19th century, when their language was described to some extent (Winkler & Pajusalu 2009). In Curonia, the Livonian language has been preserved into the 21st century. The ethnic group and their language appears to have a direct continuity from the medieval period. There has been a very notable difference between Livonian spoken in Courland and the Livonian in Northern Livonia. A noteworthy number of Finnic toponyms found throughout Northern Latvia bear witness to an earlier Livonian population residing in the region (Boiko 1993). in the direction of Courland and Livonia (Tvauri 2012: 327). The contacts with Saaremaa are likely also the reason why the Livonian language was preserved on the Courland Peninsula into the 20th century (F1). Sources from the 13th century suggest that the Oeselians were viewed as a formidable political, military and economic power at that time, and that they were known especially as a sea power. This position seems to develop with an unbroken continuity across the Viking Age (cf. Tvauri 2012: 322). Some of the features that set Saaremaa apart also seem to be shared with the coastal territories of Rotalia or today’s Läänemaa province on the mainland. This area is attributed with a centre in Lihula and had strong ties with Saaremaa. The Rotalians are mentioned by Henry as a maritime group politically associated with the Oeselians (HCL XV.2, 5; XIX.1, 3). In Old Norse sources, Eysýsla is paired with Aðalsýsla (Rotalia?)22 as another region, which seems to be the mainland rather than the other major island Hiiumaa. Although Guta saga tells that Gotlanders immigrated to Hiiumaa at some point during the Iron Age (G2), this island seems to be lacking archaeological sites from the Viking Age (Tvauri 2012: 322). The toponym Sýsla also appears in the sagas and has commonly been inferred to be identical to Eysýsla, but might also be a broader area including Ey-sýsla and Aðal-sýsla (and Sýslir varies with Eysýslir [‘Oeselians’] in manuscripts of Óláfs saga helga). In the reconstruction of areas of Finnic groups in ca. AD 1000, this ambiguity made it appear reasonable to tentatively consider Eysýsla (Saaremaa/Ösel) and Aðalsýsla (Rotalia?) as a networked region (Sýsla?), even if there were likely differences between the island and mainland cultures. The island of Hiiumaa is included in this cultural and political sphere because, even if it was not a significant settlement area, it appears that by the Viking Age “Saaremaa had become one of the most densely populated regions in Estonia” (Tvauri 2012: 322) and Hiiumaa was most likely utilized for its natural resources by inhabitants of both Saaremaa and the adjacent mainland coast (cf. Heininen et al. 2014b). F2: Saare / Sýsla / Oeselians The Oeselians are well attested in early sources. The island of Saaremaa was called Eysýsla in Old Norse and Oeselia in Latin with ethnonyms derivative from these. Henry of Livonia clearly and consistently differentiates between the Oselians and Estonians (Kivimäe 2011: 95–96). As an island, the geographical space referred to appears unambiguous. Within the context of his Chronicle, this suggests that he perceived a marked ethnic or cultural divide rather than merely a distinction of polities. Some traces of such cultural difference can also be observed in the material record (cf. Tvauri 2012: 106, 136, 140, 179–180, 196–197, 213, 264, 289, 322–323). Later, the nobility of Saaremaa formed its own house within the Baltic nobility alongside Estonian, Livonian and Courland houses. Although all of the coastal areas exhibit ongoing contacts with Scandinavia, the mainland shows evidence of contacts with south-western Finland while Saaremaa maintained contacts 85 Henry’s chronicle includes quotations ascribed to the Oeselians, such as Laula! Laula, pappi! [‘Sing! Sing, priest!’] (HCL XVIII.8). The expression is unambiguously Finnic and supports the identification of Oeselians as a Finnic language group. This is further corroborated by the Finnic toponymy that does not seem to include substantial earlier substrate layers (at least in the light of research today; cf. Kallasmaa 1996–2000). Being an island culture may have impacted language development much as in the case of Gotland (G2) and Åland (G3) mentioned above (cf. Ahola et al. 2014a: 251–253). The different contact networks in which Oeselians engaged suggests that their dialect was inclined to develop in different directions from those on the mainland. In later periods, dialects spoken in Rotalia on the mainland exhibit a great resemblance to dialects spoken on Saaremaa (Pajusalu et al. 2002). The name Saaremaa is semantically transparent: ‘Island Land’ (-maa ‘land’ referring here to a largish island, in a similar manner to the island names Hiiumaa, Muhumaa, etc.; Saaremaa could thus be interpreted meaning roughly ‘archipelago land’). In Old Norse Eysýsla, the first element ey means ‘island’ while the first element aðal in Aðalsýsla could refer to nobility in some sense but most likely denotes its referent as primary – i.e. ‘Sýsla Proper’ as opposed to ‘Island Sýsla’. In Old Norse, sýsla can either refer to activity or business or refer to an administrative district, although the former is unlikely for structural reasons.23 Viewing sýsla as ‘district’ would potentially make Eysýsla semantically equivalent to Saaremaa with a meaning ‘Island District’ and seems likely to be how the name would be interpreted by 10th century Norsemen. However, the etymology is not satisfying when a) there is no evidence that Eysýsla was seen as a district of a larger polity or even geographically as part of Eistland; b) this interpretation does not produce a satisfying explanation for Aðalsýsla [‘District Proper’?]; c) use of sýsla for an administrative area is found to the west but only infrequently and late in Sweden, which makes it generally odd in the Baltic Sea region; and d) sýslur are relatively small stewardships to which Saaremaa seems disproportionately large and which makes the term rather incongruous as a naming element for a foreign polity.24 In his Old English translation of the Orosius, King Alfred states that Wenda land is called Sysyle and situates it geographically on the southern end of the Baltic (Sweet 1883: 16). If Old English Sysyle is etymologically related to Old Norse Sýsla, the latter would have been established in Scandinavian dialects prior to syncope, and thus presumably by roughly AD 600. In this case, the Sýsla of Eysýsla and Aðalsýsla would appear to be connected to a toponym and culture at the southern end of the Baltic Sea, paralleling what is also observed for Estland (above). This would also correlate well with the fact that the theonym Tharapitha identified as a god of the Oeselians by Henry in several contexts appears to correspond to the theonym Turupið of the Wends according to Knytlinga saga 122 (see also Sutrop 2004: 34–37). In its turn, the ethnonym associated with the Wends in Germanic languages (e.g. Norse Vinðr, Old English Winedas) pops up in the Finnic languages as the denomination of the Russians (Fi. Venäjä, Est. Vene < Proto-Finnic *Venät). F3: Revala / Härjamaa Revala is attested as a district or polity on the northern coast of Estonia around the city known today as Tallinn and also as a name for the city itself. The names is well attested from the 13th century onward. It is Rafali or Rafaland in Old Norse sagas, where it may be referenced in accounts of journeys and events purportedly occurring in the Viking Age or earlier (Simek 1990: 341–343; Tvauri 2012: 31), and where it is better attested in a number of variations in Latin sources (e.g. FMU 100, 155, etc.). In the 13th century Liber census Daniae, there is an account that the area of the Reval province consisted of around 16,000 hectares of arable land, which is considerable for that period. A major sailing route between the trading centres of Birka in the Mälaren region of Sweden and Staraya Ladoga passed along this coast in a contact network that seems to have had a long history (see e.g. Schalin with Frog 2014). It can be inferred that this was a central region of Finnicspeakers already before the Middle Ages. Henry of Livonia also refers to Revala in a 86 number of contexts, addressing it as a tribe or polity of Estonia (e.g. HCL XV.3). The Christianization of Revala appears to have been associated with the Danes and resulted in this region initially falling under Danish control (e.g. described as a Danish fortress: HCL XXIX.7). From the point of view of language, the region represents the Northern Estonian heartland. The language of Tallinn has subsequently become the Estonian written standard. The area seems to have originally consisted of two entities, Härjämaa in the inland (the denomination of which the Estonian province name Harjumaa subsequently derives) and Revala along the coast. Both names likely originally derive from personal names. Härjämaa probably derives from a personal name Härkä [‘ox’] (used in a toponym; cf. surnames such as Finnish Härkönen, Härkänen). Similarly, Revala is likely derivative of a personal name Repa that originally denoted ‘fox’ (cf. Estonian personal names Repane, Rebbase (Rajandi 2005), Finnish Repo, Reponen (SN, 536) from repo or **repä). Other names for the town of Tallinn are also attested: Kesoneemi, Lindanuse (probably *Litnanaluse, i.e. Linnanalusen with a sense like ‘downtown’ as the place below a fortified settlement), Kalõvan and Tallinn, each with its own history of emergence. names of Viru fortified settlements (Tarvanpää, Agelinde; HCL XXIX.7) as well as personal names of Viru inhabitants (Kyriavanus, Tabelin of Pudiviru; HCL XXIII.7). Henry makes repeated reference to Vironia consisting of five districts which seem to form some type of unified polity: Virumaa seems to have formed a notable entity of its own with a complex internal political structure. Virumaa had intensive ties with both the Votic and Ižorian populations of Ingria as well as across the Gulf of Finland with the Karelian, and later Kymenlaakso settlement. Later, the Estonian dialect spoken in this area shows significant resemblance to Finnish dialects. A recent detailed investigation shows that the resemblance is mainly to dialects spoken on the other side of the Gulf of Finland rather than with dialects in Ingria (Björklöf 2012; Söderman 1996). The ethnonym Viru has also given the name of Estonia to Standard Finnish language (Viro). Regarding its further origin, several theories have been put forward, none of them very reliable (cf. SSA, s.v. viro; Grünthal 1997). F5: Sakala Saccalia and the Saccalians are mentioned in medieval sources and especially in the Chronicle of Henry. Henry frequently refers to Saccalians simply as Estonians, but also uses an ethnonym derivative of Saccalia for them (e.g. HCL XII.6). As elsewhere, he refers to personal names (e.g. the elders Lambito and Meme; HCL XV.1) as well as, for example, the province name Aliste (today’s Hallist; HCL XV.7). The identification of provinces suggests a relatively complex political structuring of Saccalia before the introduction of Christianity. According to Henry, when the Ugandians (F6) conceded to the hostile pressures of the Germans and asked for baptism, the Saccalians also asked for baptism in order to avoid a similar fate. By today, the area is (northern) Estonianspeaking but the dialects of the region (Mulgi) bear traits of substrate influences from Southern Estonian hinting at a different character of the local language form in the Middle Ages (Pajusalu 1996). F4: Virumaa Virumaa [Est. ‘Viru-land’] is the eastern major historical province of Estonia clearly attested from the 13th century as a polity also along the eastern route. The Old Norse form Vírland is weakly attested in saga literature but mentioned on three commemorative rune stones in Sweden (Tvauri 2012: 31–32) and corresponding forms are found in other/later Germanic languages. It is referred to as Vironia in Latin. Henry of Livonia mentions in his chronicle that an elder of the Viru was baptized by Gotlanders (HCL XXIII.7). Later, Henry arrives with a companion to spread Christianity through the population, but as representatives from Riga, this produces conflict with the Danes; the conflict is resolved by having the Danes carry out the baptisms (HCL XXIV.1–2), which was tantamount to subjecting the population to Danish political control. Henry preserves the 87 relationship between the terms is nevertheless not entirely clear and it is interpreted in relation to the reconstructed etymology, but the name is probably ultimately of Slavic origin, deriving from the word root denoting ‘south’ (see further Grünthal 1997: 207–213). F6: Ugandi The Ungania and the Ugandians are first referred to in Henry’s Chronicle; corresponding names are not known from either Scandinavian or Old Russian sources (Grünthal 1997: 205–206). Although Henry uses an ethnonym derivative of the toponym for the Ugandians, he also refers to them as Estonians (e.g. HCL XII.6). Within the archaeological record, this region belongs to the relatively homogeneous inland, but it also appears that the waterways of Lake Peipsi and of the Emäjõgi River were the channels through which cultural influence from Novgorod and Pskov spread (Tvauri 2012: 86). Old Russian chronicles also recount the conquest of the region by Jaroslavl’ the Wise in the 11th century. Slavic influences on this region thus appear more significant and more direct than farther north and northeast. Henry claims that, in the early 13th century, forces from Novgorod and Polozk invaded Ungania, at the outcome of which a number of Ugandians were baptized (HCL XIV.2). However, Henry later goes on to describe how they were under duress from Germans and they surrendered and asked for baptism from them (HCL XIX.4). Toponymy of the region mentioned by Henry suggests a Finnic language area. For example the fortress Odenpäh [‘Bear Head’] seems to have been a, if not the major centre for the territory. In more recent times, this is the region of the South Estonian speaking groups, Võru and Seto. The eastern parts of the region are the area inhabited by the Orthodox Seto groups. The rich Slavic substrate nomenclature of South Estonian constitutes significant evidence of the Slavicization of the region (Kiristaja 2013). The historical Ugandi extended up to the city of Tartu, which would still seem to have been largely South Estonian in terms of language even in the 17th century. The great linguistic difference between the dialect spoken here and Estonian proper has been preserved up to the present era. The name Uguani is historically related to Latvian Igaunija as the common term to denote Estonians. The Latvian term appears to be generalized from the polity or tribe with which they had direct and ongoing contact. The F7: Minor Finnic Groups in Estonia Additional Estonian provinces are mentioned in the historical sources, most notably in the Chronicle of Henry. These have been small units but have nevertheless usually been considered independent mini-states that had their own groups of elders as authorities, their own strongholds and to have formed alliances independently. These provinces have been referred to as ‘small provinces’ (väikemaakonnad) and they included the Central Estonian Alempois, Jogentagana, Mõhu, Nurmekund, Soopoolitse and Vaiga (cf. F10). Some of the names of the small provinces have archaic Finnic traits pointing to the fact that many Estonian sound changes had not yet occurred by the 13th century in the relevant dialects. It is also possible that other similar small semi-independent formations have occurred, although there are no historical accounts regarding them. It remains unclear how these minor polities should be viewed in relation to the districts of larger regional polities. Finnic Groups on the Eastern End of the Gulf of Finland and in the Ladoga Region The present Finnish language represents an amalgam of the types of old Finnic vernaculars. It is reasonably clear that whereas one of them entered Finland from the south across the Finnish Gulf, others came via the Karelian isthmus and the Ladoga region. These latter areas have a very notable evidence of the early presence of Finnic groups both in the archaeological record (Uino 1997) as well as in the anthroponymic types from pre-Christian periods (cf. Saarikivi 2007). The division of the Modern Finnish language area to two entities is visible in the division line between the eastern and western dialects through a number of isoglosses of different features. From the point of view of cultural history, the western dialect area has been associated with the networks of the emerging Swedish state and the Western 88 13th century (Korpela 2008a: 46–47).25 This would account for characterization of Karelians through their alliance with the Novgorodians or Rus’ (even if subject to taxation), undertaking joint military action (cf. Grünthal 1997: 78–79), as well as being attributed with the destruction of Sigtuna in 1187 as an independent action.26 Indications of longstanding independence suggest that ‘Karelians’ formed a distinct polity or confederation of polities already by ca. AD 1000. The concept of Karelian language is an ambiguous one. It comprises different types of Eastern Finnic language forms. Some of the current forms are classified as Finnish dialects and others as Karelian language. Today’s Ludic dialects are classified as an independent language form in the Finnish research tradition and as a dialect group of Karelian in the Russian tradition. And those dialects considered as Karelian form a complex dialect continuum between Finnish and Vepsian dialects. The probable historical core area of the Karelian tribe, at the eastern end of the Gulf of Finland, has, in the 20th century, been Finnish-speaking: much of the original population was displaced to Central Russia in the 17th century when the region was seized by Sweden. It is not quite clear what the relationship between this earlier population and the Ingrian ethnos has been to the Karelian community that subsequently evolved. In the 13th century, Ingrians or Ižorians also appear as an ethnic group to the south of the River Neva in Ižora or Ingria. However, the Ižorian language appears to have developed from a dialect of Karelian and there is evidence that at least some Ižorian speakers had earlier used the ethnonym Karielaizet for themselves (Nirvi 1971: 137). Many questions remain open about the cultural history of the Ižorians, but it is in any case not clear that they formed a distinct tribe or polity already in AD 1000 and they are therefore not differentiated on Map 2. The Karelian community seems to have centrally spread from the western and northern shore of Lake Ladoga in many directions, for instance to the Kemi river valley on the northern end of the Gulf of Bothnia (Vahtola 1980), through Northern Church, and the eastern with Novgorod and the Eastern Church. The inner Finland Lakeland area stands between these two areas and is where Savo dialects are today spoken. Linguistically, these belong to the eastern dialects, but from the point of the history of statehood, the Savo region was integrated into Sweden early on. The mechanisms of the Fennicization of this region, where Sámi languages were likely preserved longer than in the western or eastern core areas, still remain to be clarified (cf. Korpela 2012). F8: Karelians (and Ižorians / Ingrians) The Karelians and Karelia are mentioned in Norse, Slavic and Latin sources beginning from the 13th century (Grünthal 1997: 78–80). References in western sources are only occasional, often ambiguous and lack reliable linguistic or ethnographic information. Old Norse sources situate Kirjálaland [‘Land of the Karelians’] as the next ‘land’ after Finnland in a topogeny circumnavigating the Gulf of Bothnia (Egils saga 14; on this topogeny, see the discussion of Finnland below), or otherwise situate it on the Eastern Route before Garðaríki (Fagrskinna K27). The Gulf of Finland in Old Norse was Kirjálabotn [‘Bottom/Gulf of the Karelians’] (Hálfdanar saga Eysteinssonar 15–16, 24– 26), as in Icelandic today. (See also Grünthal 1997: 79.) The extent to which Kirjálaland was understood to extend to the east and north is difficult to assess. The 14th century Erikskrönikan, for example, situates Karelia to the north of Lake Ladoga (Klemming 1865: 51). Slavic Korela [‘Karelia’] is first and foremost an ethnonym but geographically it refers to the north-eastern coastal areas of the Gulf of Finland and areas around Lake Ladoga. At least in the 13th century sources, it is not normally considered to extend to control of the Neva river, beyond which would be Ižora [‘Ingria’] (Korpela 2008a: 46– 47). Western sources generally leave the political status of Karelia ambiguous, although a late 13th-century Old Norse account mentions that it was subject to taxation by Novgorod around that time (Kjær & Holm-Olsen 1910–1986: 623). However, especially Slavic sources suggest that Korela was independent of Novgorod into the mid89 fortress of Korela, the later fortress of Priozersk/Käkisalmi. In this case, the Karelians of the early documents were a group of Eastern Finnic speakers centred around a particular fortified settlement, in a manner somewhat similar to several communities of early Slavs (cf. Sv1). Karelia (Kuzmin 2014) and probably also to the coast of the White Sea. This can be correlated with an archaeological culture that established settlement areas on the Karelian Isthmus and also around accesses to inland water routes to the northwest and north from Lake Ladoga, with the implication of a capacity to exert regulatory control over those water routes (Heininen et al. 2014a: 313). Later evidence of language spread would thus be connected to the association of Karelians with precisely these inland routes to the Gulf of Bothnia and the White Sea. The background of these groups that became Karelian by ca. AD 1000 seems to be complex. Immigrant groups arrived from Southwest Finland in the 8th century and settled in the areas identifiable with Karelians, where they merged with indigenous groups, becoming a distinct culture in the archaeological record across the Viking Age (Uino 1997). This process accounts for the pronounced western substrate in later Karelian linguistic heritage that separates the cultures of Finland and Karelia from those of other Finnic groups (Frog 2013). Viking Age Karelia thus appears to have at least initially been a multi-ethnic environment. It has been proposed that the Finnic toponym Karjala [‘Karelia’] (and ethnonym Karjalainen) derives from karja with the affix -la commonly used in forming toponyms. Thus, it is fairly clear that the name Karjala is originally a toponym but we do not know the original denotation of that toponym, something that makes the attempts to etymologize it fairly suspicious. The likely etymology of karja is from a Proto-Germanic or North Germanic *χarjaz [‘army, host, crowd, mob’], which would originally have been borrowed presumably in the Roman Iron Age.27 There are, however, notable arguments against the etymologization of Karjala from karja. For instance, the commonly spread name form Kariela that, on the basis of the Russian early written forms, could be the original form of the toponym and would point to some other kind of stem word for the etymology. Judging by the ethnonym structure, the most likely explanation is that it is derived from a settlement name. The original settlement could have been that of the F9: Vepsians / Ves’ The Ves’ are mentioned in a number of contexts in Old Russian sources and the ethnonym seems to appear as Wasu in an Arabic travelogue (Grünthal 1997: 103–104, 106–107). Jordanes may nonetheless refer to the same group as Vas already in the 6th century (Getica 23) at the beginning of a series of peoples that appears to represent a trade route to the east:28 Thiudos Inaunxis Vasinabroncas Merens Mordens Imniscaris.... When ethnonyms are distinguished from (probable) prepositional phrases identifying locations in this text, it can be read: ‘Thiudos in Aunus Vepsians in Abronkas(?) Meryas Mordvins in Meštšora...’. Possible but problematic references to Wizzi are found in Adam of Bremen’s Latin History of the Bishops of Hamburg-Bremen: he states that the Albini speak Wizzi, but the two terms look suspiciously like the respective Latin and German words for the colour ‘white’ (Matthews 1951: 40). It is striking that the ethnonym is not found in lists of Finnic groups in annals or ecclesiastical documents (e.g. FMU, 84), and the ethnonym does not seem to appear in medieval Scandinavian sources.29 If Jordanes does in fact present a topogeny, this would suggest that the Vas (in the 6th century, too early to distinguish Vepsian language) were between the Thiudos and the Merens along that route. In the Old Russian chronicles, the Ves’ are mentioned already in the Primary Chronicle as located on Lake Beloye, whereas Čud or more specific ethnic terms are used for closer groups (Grünthal 1997: 104–105; cf. E5). If Jordanes’ Merens indeed is related to Merya (although see note 52), this location (or direction) would seem consistent with his topogeny. Arabic sources also would appear to situate the Wasu considerably to the east (Grünthal 1997: 106– 107). The geographic remoteness of this group would help explain why the ethnonym 90 designated a different group and (whether or not that group underwent a language shift to Vepsian) contacts with that group seem to have somehow led the ethnonym to be transferred to Vepsian speakers in other cultural areas. The etymology of this ethnonym remains obscure (Grünthal 1997: 108–109), and even its language of origin has not been convincingly identified. In the 20th century, only a group of Southern Vepsians used this ethnonym to refer to themselves; other Vepsian-speaking groups employed the ethnonyms Lydiläzet [‘Lude’] or Cudid [‘Čud’’]. is not attested in Scandinavian sources. On the other hand, this location does not accord with the area south and east of Lake Ladoga with which Vepsians are associated in more recent times. In these areas, there appears to be a correlation between toponymy of early Vepsian habitation and landscapes suitable for the variety of agriculture practiced there, suggesting a correlation between language and culture/livelihoods in the region (Kuzmin 2014: 287–291). Archaeology has also considered that the area on the south-eastern shore of Ladoga that is characterized by the type of mound graves called sopkas would have been the core of Vepsian settlement. The later identification of Finnic groups in the Ladoga region with this ethnonym indicates that either ethnonym of a tribe or another language group has been inherited by the Finnic-speaking Vepsian population (Mullonen 2002), or that the current Vepsian population is a direct continuation of the historical Ves’. This picture can be developed by reverseengineering the background of Vepsian language areas on the basis of linguistic data. Terho Itkonen (1983: 216–217) has argued that a substrate of another Finnic language can be detected in Vepsian, a substrate that has resulted from the spread of Gulf of Finland Finnic to the east. Investigation into Vepsian toponymy suggests that the Olonets Karelian and Ludic language areas emerged in an area of earlier Vepsian habitation (Mullonen 1994; 2002). The Vepsian language area was thus much more considerable in these western regions at an earlier time. This would confirm Itkonen’s hypothesis that there is a Vepsian substrate in Olonets Karelian (cf. Mullonen 1994). Detailed investigation into the toponymy of the Lake Beloye region has confirmed some parallels with Vepsian vocabulary, but also other types of substrate, most notably parallels with what has been considered the Meryan core area in Russian toponymic research and also some parallels with Sámi toponyms (Makarova 2012; Saarikivi 2004). When the ethnonym Ves’ is attached to groups in this region and only later to groups in the core region of the Vepsian language, it can be inferred that Ves’ probably earlier F10: Vatja / Vod’ The earliest reliable references to the Votes are from the 13th century in Old Russian chronicles and in a few references in Latin. Perhaps the most prominent use occurs in the 15th century: when Novgorod fell to Moscow and its territory was divided into five districts, one of these was called the Vodskaja pjatina [‘Votic Fifth’]. This is a testament to use of the ethnonym. The region that it describes extends north from Novgorod through Ingria, between Lake Ladoga and the border of the Swedish realm, and continues to the north of Ladoga (thus including Korela/Kirjálaland; F8). The ethnonym had already earlier been used for a volost’ [‘district’] of Novgorod and a similar concept appears to be behind references to Watland beginning from the first half of the 13th century (e.g. FMU 84; see Grünthal 1997: 123–125). The ethnonym is also associated with a different region to the west called Vaiga (etymologically related to Vatja), mentioned already in the Chronicle of Henry of Livonia; Vaiga is on the northwestern side of Lake Peipsi, beyond the scope of the Novgorod Fifths (Grünthal 1997: 125– 127). In more recent times, Votes and the Votic language are situated in the culturally complex region of Ingria alongside Ižorians (F8) and later also groups of the so-called Ingrian-Finns who immigrated from north of the Gulf of Finland (most notably from Savo) probably in the 17th century, following the Stolbova peace agreement and mass emigration of the earlier orthodox inhabitants that were linguistically likely close to the Ižorians.30 The area associated with the Votes in more recent times is only a tiny area within 91 ‘wedge’ (~ Fi vaaja < Proto-Finnic *vakja), which has in turn been linked to the Old Russian expression Čud’ v Klinu [‘Čud’s in Wedge’] as a reference to Votes, where Klin would be a calque. The ethnonym vatja also has a somewhat dubious etymological counterpart in Sámi, a hapax legomena wuåwiåsh: wuåwiåha [‘Lapp person’] that is only mentioned in the (1787) dictionary of Christfrid Ganander (cf. Korhonen 1981: 43); this word has been connected to a Lule Sámi word denoting to a wedge (vuoi`ve). In earlier scholarship, this etymology had been pushed into an interpretation of not ‘wedge’ but ‘club’ with a fanciful idea that it referred to some sort of baton as a mark of identity or authority. This interpretation has been debunked (Koivulehto 1997), but not before being linked to the obscure Old Norse ethnonym Kylfingar.31 Riho Grünthal has argued for an alternative etymology linking it (or the toponym Vaiga, from which the ethnonym could derive) to a Latvian ethnonym for Germanic peoples (Lv. Vācija, Lith. Vókia [‘Germany’]; cf. SSA III: 418), which raises questions about the history of this group and the region. (See Grünthal 1997: 113–149 and works there cited.) the Votic Fifth. It is unclear how prominent the Votes were in the period of the Fifth’s formation as its designation could be based on the naming of the earlier volost’. The earlier volost’ seems unlikely to have originally included Korela/Kirjálaland insofar as this seems to have remained largely or wholly independent at least through much of the 13th century (F8). The Votic volost’ therefore presumably did not originally extend beyond the River Neva on the Karelian Isthmus. The Votic language is a Finnic language that, in many respects, can be considered an intermediate language between the eastern, southern and northern groups of Finnic. It is clearly distinct from Ižorian and so-called Finnish-Ingrian dialects as well as from Estonian. Many questions about the historical background of this linguistic and ethnic group remain open. This is also true of defining its boundaries in the context of many intertwined language forms of Ingria. For instance, the Kukkosi dialect has traditionally been considered Votic, although it lacks many central features of Votic language such as the affrication of the *k and its speakers have not used the corresponding ethnonym. In the archaeological record, certain burial types in the Vaiga region of north-eastern Estonia from the 12th century and later have been interpreted as Votic or Votic and Slavic (Grünthal 1997: 117–118). However, cremation burials predominated in these regions in the Viking Age, which aligns the region with North Finnic cultural areas and distinguishes it from territories of Estonia (Tvauri 2012: 268). Most of Ingria reveals an old agricultural region that is also reflected in the old settlement names derived from pre-Christian personal names (Kepsu 1995). The Vaiga region east of Lake Peipsi has traditionally been viewed in scholarship as the ethnic homeland of the Votes. S.A. Myznikov (2003) points to the peculiar character of the Finnic lexical substrate in the Russian dialects of this area that supports the hypothesis that this region was inhabited by a distinct cultural group. The etymology of the ethnonym Vatja has a long history of being interpreted on the basis of phonetic similarity to a word meaning Western North Finnic Groups For territories north of the Gulf of Finland in today’s Finland, the few early historical sources provide almost no information about the indigenous language and culture. A few general remarks are warranted about the Finnic language groups identified in this section. First, toponymy reveals that the Swedish language areas of Finland were only established from the 12th century onwards, even if a few scattered micronyms might be open to question (e.g. Schalin 2014; Heikkilä 2014b; Pitkänen 1985).32 Toponymic borrowings from Finnic languages into Old Swedish suggest that communities practicing fixed-settlement livelihoods and co-occurring funerary practices in these territories were Finnic speaking at the time Swedish language areas in present-day Finland became established. Sámi language presence in present-day Southern Finland is indicated by Finnic toponyms and dialectal vocabulary of substrate character (Aikio 2007; Aikio 2009). 92 burial practices was directly connected with the Church or even with organized missionary activity.33 The areas under consideration were extremely peripheral both politically and economically at that time, before Christianity had secured its position in the emerging states of either Sweden or Russia, and long before the Baltic Crusades. These changes in practices should therefore be assumed to be linked to internal social, religious and political changes in the networks of these groups (cf. Ahola & Frog 2014: 42–43, 68). The change in practices remains noteworthy, however, because the first inhumation graves appear in cremation cemeteries at the same time through the regions of Finland Proper and Häme (Wessman 2010: 78): whatever languages were spoken by in the local polities at that time (most likely, already predominantly Finnic), they were clearly closely networked in interaction. The rate of spread of cultural practices through these networks suggests some type of shared or convergent identities. The Old Norse term Finnland appears to refer to territories inhabited by Finnic speakers of southwest Finland with continuity through the present day. The name Finland (its eastern form) appears on commemorative rune stones in Uppland, Sweden (U 582, lost, undated), and on Gotland (G 319, 13th century). The ethnonym Finnlendingar [‘people of Finnland’] is found in a skaldic verse dated to the early 11th century (Sigv Víkv 3I.3).34 From the 13th century onwards, the toponym is attested in various written sources in Latin and Old Norse.35 The precise geographical space that it designated in the Viking Age and early Middle Ages nevertheless remains rather vague. Old Norse sources written in distant Iceland generally seem to present Finnland as across the Baltic Sea from the lands of the Svear (e.g. Ynglinga saga 19) and probably adjacent to Kvenland (F18) (Orkneyinga saga 1) or between Kvenland and Kirjálaland [‘Land of the Karelians’] (F8) (Egils saga 14), and perhaps north(?) of Eysýsla [‘Saaremaa/Ösel’] (F2) (Óláfs saga Helga 8–9). In Old Norse topogenies listing a geographical series of ‘lands’, these may refer only to spaces characterized by fixedsettlement habitation without necessarily The substrate becomes increasingly observable to the north of the archipelago area and inland through areas where the fixedsettlement culture of the Häme region was gradually expanding. The relationship of this Sámi habitation to the present Sámi groups and the chronology and mechanisms of the Fennicization of the area of present-day Finland remain poorly investigated. More than a century ago, Alfred Leopold Fredrik Hackman (1905) argued that the Finnic language entered Southwest Finland in the Iron Age due to a migration from Estonia. This ‘migration theory’ (maahanmuuttoteoria) remained the basic paradigm for the prehistory of Finland up to the 1970s and 1980s when it was replaced by a ‘continuity theory’ (jatkuvuusteoria), which attempted to prove that the early contacts between the predecessors of Finnish and the IndoEuropean languages, most notably Germanic, took place in Finland during the Bronze Age or even earlier. During the 2000s, the continuity theory has been increasingly criticized (e.g. Aikio & Aikio 2001), and the paradigm has been gradually shifting again to dating the arrival of Finnic languages in Finland to the Iron Age or even into the present era, as will be discussed in a future article in this series. However, the arrival of Finnic languages is presently not understood as simple migration, but as a more complex phenomenon. On the basis of abundant pre-Christian anthroponyms in the toponymy of Finland Proper (F11), Häme (F12) and Satakunta (F13), it is clear that this region received a bulk of Finnic-speaking settled population well before Christianity. In this respect, these regions are also notably different from other regions of Finland. However, when considering the historical language situation and language contacts, Christianization has potential relevance, not least by incorporating polities and populations into a broader cultural sphere of communication and shared practices. The change to inhumation burial practices across ca. 1000–1150 has been customarily interpreted as the general conversion of Finnic populations to Christianity (Huurre 1979: 224). However, there is nothing to suggest that this change in 93 them gradually motivated distinguishing them from the mobile groups. The authors wish to stress that the four cultural areas of Western Finland considered here for ca. AD 1000 follow from territorial divisions customary to research traditions. These research traditions build centrally from the archaeological record and have evolved in dialogue with (and have sometimes been conflated with) historical provinces with the relevant designations. Patterns in the archaeological record have supported distinguishing these cultural areas (cf. Raninen & Wessman 2014: 331). It must nonetheless be stressed that each contained multiple communities, some of which exhibit settlement continuity from the beginning of the Iron Age (Asplund 2008: 365), others emerged through at least some degree of immigration from Scandinavia in earlier centuries (Wessman 2010: 35) while others emerged and/or expanded during the Viking Age. The communities in each region should not be considered culturally uniform and it cannot be assumed that they were linguistically homogeneous. It does not appear that any of these regions experienced a centralization of power under a single king. The networks formed by different polities remain only poorly understood. mentioning wilderness areas between them. Thus Finnland could be next to Kirjálaland because there seems to have been no settled area of note between them (e.g. Huurre 1979: 158–159).36 This observation is relevant because the location of Kvenland remains uncertain and thus so does the northern extent of Finnland. It is equally unclear whether or to what extent Finnland may have included the archipelago (see also G3). Nevertheless, only two Old Norse terms for large territories of later Finland are attested – Finnland and Tafeistaland (F12) – of which the latter later refers to the inland areas of Finnic habitation whereas Finnland seems to have had a more coastal referent. Later Swedish names for other districts seem to be of medieval origin rather than having continuity from the Viking Age. This implies that the areas were not toponymically distinguished at that time. Thus the Scandinavian term Finnland seems to have been rather broad in scope. The name Finnland is etymologically more or less transparent: it is a compound Finnland of which the latter part indicates an inhabited territory and the former element is Old Norse Finnr (S) with the meaning ‘land of the Finnar’. Where Old Norse sources narrate about the inhabitants, they call them Finnar and present narrative plot-types associated with that ethnonym (cf. Aalto 2014: 216–218) or describe only marshal conflicts rather than cultures (cf. Schalin 2014: 422–425). When this toponym was established and who the Finnar were at that time (cf. S) remain obscure, but the formation suggests a land or geographical space inhabited with fixed settlements by people that the Scandinavians called Finnar (Aalto 2014: 214). In the development of a geopolitical relation with these groups, dialects of (Central) Sweden adopted a new term Lappr to designate the mobile groups for which the term Finnr had been used previously (S). It is not clear how this ethnonym shifted from one group to another, whether groups identified as Finnar were practicing fixed-settlement livelihoods and later underwent a language shift to ProtoFinnic or arriving Finnic groups were initially called Finnar as a broad category of ‘other’ and the political and economic relations with F11: Suomi / Finland Proper The Scandinavian term Fin(n)land and corresponding Finnic Suomi seem to have been established early on. Although not before the 16th century, they became general terms to denote the whole of what in the Middle Ages was called the Swedish Österland, or the Swedish territories east of the Baltic Sea. By or around AD 1200 (Sjöstrand 2014: 95), the bishopric was established in Åbo/Turku in the Aura River valley at the heart of what became the medieval province of (Finnish) VarsinaisSuomi or (Swedish) Egentliga Finland [‘Finland Proper’]. These factors suggest that this region was considered central in the integration of territories north of the Gulf of Finland into the emerging Sweden. In the Viking Age, the cultural area seems also to have extended farther to the west into what is now western Uusimaa / Nyland (Haggrén 2011). This territory consists now of both 94 different views on how long Finnic languages had been spoken in the region prior to this (e.g. Hackman 1905; Lehtosalo-Hilander 1984; Edgren & Törnblom 1993; Salo 2000). The area at the northern end of the archipelago known as Vakka-Suomi appears to be part of this broad territory, and Finland Proper seems to have been closely networked with the inland region of Häme / Tafeistaland, for example in the more or less simultaneous widespread appearance of inhumations in the cremation cemeteries under level ground (Wessman 2010: 28). At the same time, this region does not appear to have been completely homogeneous: Vakka-Suomi maintained particular connections to northeastern areas of the Åland Islands (Heininen et al. 2014b: 332, 334, 339), and whereas Finland Proper and Häme / Tafæistaland both exhibit silver hoards beginning in the 11th century (Talvio 2014: 135), Vakka-Suomi aligns with Satakunta in the lack of evidence of such hoarding (Raninen & Wessman 2014: 331). The origin of the name Suomi has been variously interpreted, mostly as a borrowing (meaning ‘low land’ or some kind of a person) from Baltic or another Indo-European language (cf. Grünthal 1997; Kallio 1998; SPK, 430 for various etymologies). The early attestation of Suomi as a personal name is not counter-evidence to its earlier connection with a geographical entity. On the basis of analogies, it seems likely that such an entity would not have been a territorially large one. Toponyms derived from suomi or its cognates are also found outside the area of modern Finland, most notably, in Estonia (Grünthal 1997: 53–55) and could perhaps shed more light to the issue despite the long history of research on this topic. Finnic-speaking and Swedish-speaking areas, mainly in the archipelago. Toponymic evidence reveals that the Swedish-speaking population did not develop before the 12th century (e.g. Heikkilä 2014b) but archaeological evidence of this process east of Åland remains limited (Asplund 2008: 377– 381). A pronounced Finnic substrate toponymy points to Germanization during but not before the 13th–15th centuries (cf. Pitkänen 1985). It is unclear when Finnland and Suomi converged to refer to the same geographical space. The identification of ‘Finland Proper’ nevertheless identifies a core area from which ‘Finland’ was generalized. Prior to the 12th century, Finnland does not seem to have been complemented by terms for other territories along the coast whereas the Finnic term most likely (at least initially) referred to a culturally defined space as perceived by its inhabitants. It is possible (but speculative) that, near the beginning of the present era, the Greek geographer and historian Strabo offers the earliest attestation of the name Suomi [Fi. ‘Finland’] (Ζούμοι / ‘the Zūm-M.PL’) as an ethnonym for a northern tribe (see Grünthal 1997: 53). This would however only attest to continuity of the term likely connected with place rather than indicating either the culture or language of the tribe in question. A more probable attestation is the appearance of Suomi as a personal name in a list of names in a Frankish annal from AD 811 (Grünthal 1997: 54).37 The term Sum’ appears as an ethnonym in Old Russian chronicles. Although not all uses of this ethnonym seem consistent, the Sum’ are identified as allies of the Swedes in connection with the Battle of Neva (AD 1240), in which context Sum’ could quite possibly be a rendering Suomi (Korpela 2008: 45–46). Forms of Suomi are also found in most Finnic languages as referring to Finns and Finland (Grünthal 1997: 51). Considering the cultural area identifiable with Suomi/Sum’ as Finland Proper concentrates it around the Aura River valley and surrounding region. On the basis of the archaeological record, it is generally agreed that this area has been continuously settled by a Finnic-speaking population since at least the 7th century AD, even if scholars have held F12: Tafeistaland / Tavastia / Häme The core area of the culture of the so-called Lake District of inland Finland had a sedentary settlement centred around the Kokemäki River basin and the southern part of the Lake Päijänne area (cf. also Raninen & Wessman 2014: 331). This area later becomes identified as Häme in Finnish, which is correlated with Old Norse Tafeistaland and Latin Tavastia; it is not clear that the region is referred to in Old Russian sources.38 The 95 Tavastia into Swedish may have developed under influence of the spread of the Church and role of ecclesiastical discourse from the time of the medieval Swedish immigration. cultural area of ca. AD 1000 seems to have extended into northern parts of what is now western Uusimaa / Nyland (Haggrén 2011; cf. F11). The name’s earliest appearance is in an 11th century commemorative runic inscription (Gs 13) and it begins appearing in other written sources in 1237 (Schalin 2014: 416– 418).39 Especially the Kokemäki River basin has a very rich layer of pre-Christian settlement names that point to continuous sedentary habitation from the Iron Age. Review of these names has led to a theory that the distinctive Finno-Karelian mythology developed in this region before spreading to dialect areas in the Ladoga region (Siikala 2012). In later periods, the Häme manor houses (kantatalot) had large wild forest and taxation areas in Keski-Suomi and Ostrobothnia. It has been argued that migrants from Häme had an important role in the settlement of the latter areas. It has also been demonstrated on the basis of the spread of toponymic types that Häme was the core area from which migrants spread to the Tornio River valley and the lower Kemi River valley (Vahtola 1980). Although Finnish Häme and Swedish Tavastia now translate one another, the history that led to their identification remains unclear. The province name Häme derives from the tribe name Hämäläinen that is an etymological cognate of the name of the Sámi people, Sápmelas. It has been argued that Häme was inhabited by Sámi (or Lapp) people who subsequently Fennicized (Salo 2002; Aikio 2012; cf. also Korpela 2012). The Scandinavian name appears to have been Tafeistaland (or Tafæistaland in eastern dialects) which seems to follow a common pattern of an ethnonym *Tafeistr in genitive plural followed by land (Jackson 1993: 43; Schalin 2014: 416–421). This finds support in the Latin ethnonym Tavastus (e.g. FMU 82) of which the Latin toponym Tavastia appears to be derivative.40 The etymology nevertheless presents difficulties, but Johan Schalin (2014: 416–421) has recently argued for the basis of the toponym on an ethnonym (eastern dialectal) *Taf-æistr [‘luggard(?)Estonian’], pointing out that this appears as a name or designation in runic inscriptions (U 467; U 722). The later adoption of the Latin F13: Satakunta The toponym Satakunta is first mentioned in 1331 (Salo 2000: 108). The language of groups in this area in ca. AD 1000 is inferred especially through toponymy correlated with evidence of cultures in the archaeological record. The broad territory exhibits some degree of distinction from the regions of Tafaistaland and Suomi in the absence of evidence of hoarding silver, yet more scales have been found in Satakunta than in the rest of Finland combined (Talvio 2014: 134–135). The Kokemäki River seems to have been a route by which, according to the earliest cemeteries, the culture on the coast advanced into ancient Häme in the 4th century AD (Salo 2000: 85). The burial type of the area is characterized by earth-mixed cairns with cremations and a cremation cemetery under level ground is also found in the river valley (Wessman 2010: 33, 78). The Kokemäki River valley was likely an independent centre of Finnic settlement already in the Iron Age (Salo 2000). This is demonstrated by the rich layer of pre-Christian anthroponyms in the settlement names, pointing to a pre-Christian sedentary settlement. Toponyms have also been used to argue for traces of “early Germanic migration” into the area (Vahtola 1986; Salo 2000: 84–88). However, the evidence is very limited and not unproblematic; the few indisputable etymologies on the basis of anthroponyms do not reveal the language of their bearers (cf. Schalin 2014: 401). Finnic cannot be considered the exclusive language of this region in the Middle Ages. There is notable evidence of Sámi settlement in the area in the toponymic material (Salo 2002).41 It seems likely that there was in some period some kind of an ethnic boundary between the Finnic population in the lower Kokemäki River and the Sámi population in the water basins upstream. Not far to the south, the area around Lake Pyhäjärvi in Lower Satakunta along the Eura River stands out in the archaeological record. When elsewhere throughout Finland burial practices 96 ‘group of men’] has denoted both an entity of people as well as a land area or (later) an administrative district (Heikkilä 2014a: 181– 199). This etymology is phonetically appealing, but it is difficult to account for why it would be generalized to the larger inhabited area. The region was not characterized by harbours (cf. Salo 2000: 92, 100), was more generally remote from major sailing routes (Heininen et al. 2014b: 333), and the major settlement areas were inland, apparently remote from e.g. the harbour on the Kokemäki River rather than characterized by it (cf. Wessman 2010: 32). This presents one possible explanation for the toponym, but the grounds for calling either this region or formalized administrative district ‘Harbour Place’ remain more arbitrary than compelling. were based on cremation, inhumation cemeteries begin appearing here in the at the end of the 6th century, some of which are used continuously into the 13th century (LehtosaloHilander 1984: 297). Although various Scandinavian burial practices seem to have influenced cultural areas in Finland in the Migration and Merovingian periods (e.g. Lehtosalo-Hilander 1984: 272–284), these cemeteries suggest a more significant impact of Germanic funerary customs more likely indicative of at least some degree of immigration (see also Wessman 2010: 35 and works there cited).42 This distinctive cultural area can be viewed in relation to immigration to Åland (G3) and population expansion in Svealand (G1) during the same period. The general absence of Germanic toponymy apart of anthroponymic settlement names from before the medieval immigrations nevertheless suggests that these communities did not speak a Scandinavian dialect in the Viking Age. If they spoke a Finnic dialect, the archaeological record still suggests a marked distinction from the communities in the Kokemäki River basin. The toponym Satakunta does not designate an ethnic group, yet it is of interest because it could mean ‘hundred-company (of men)’ as a calque for a Germanic hundare or ‘hundred’ found widely in Germanic place names. A hundare was a political area subject to the military subscription system requiring the polity to organize a ‘hundred’ (then 120) men on demand for military support in war or raiding. This etymology would suggest that the territories of coastal Finland were under obligation to the kingdom of the Svear. (See Salo 2000: 114–128.) This would not be surprising considering the geopolitical situation of Finland in relation to the activities of the Svear (G1), particularly when relations between the culture of the Lake Pyhäjärvi area and the Svear might have had a longue durée (Heininen et al. 2014a: 298–299, 300– 301). The etymology nevertheless remains uncertain. Satakunta might also have originally simply referred to a ‘harbour place’: Fi. satama likely derives from a Germanic *staþa [‘shore; landing place’] (SSA III: 160, LÄGLW III: 224-225) and kunta, [orig. F14: Kyrö Culture The northern-most cultural area in the Baltic Sea region that can be tentatively inferred as Finnic-speaking in the Mid-to-Late Iron Age is the early centre of habitation in the Kyrö River valley in Southern Ostrobothnia. This area exhibits cremation cemeteries under level ground (Wessman 2010: 20, Figure 4, sites 2 and 4), although exceptional water-burial cemeteries appear in relatively close proximity to each of these (ibid., sites 1 and 3) and other burial types are also found. The cemeteries under level ground are a distinctive cultural type customarily identified with Finnic speakers. They are found in the respective cultural areas of Satakunta, Häme and Finland Proper, and are observable also in Karelia where their appearance is considered a relevant indicator of immigration from these western regions beginning from the 8th century (Uino 1997: 174–179). Similar practices also appear in Estonia from roughly the same period as those in Finland (Jonuks 2009). The spread of these practices to the area of the Kyrö River are therefore suggestive of Finnic language presence, but this is not unproblematic. This type of burial seems to have spread through contact networks. The cemetery type is also found on the Courland Peninsula (cf. B1) and more recently may have been found in Svealand in Central Sweden (Wessman 2010: 21). The cemetery type can be correlated with Finnic-speakers in most areas 97 where it occurs in Finland and it seems associated with Finnic language areas in Estonia and (possibly) on the Courland Peninsula. Nevertheless, it cannot be considered as an uncontroversial criterion for language identification in the fairly remote area of the Kyrö River. More significantly, use of the burial grounds in this northern periphery gradually declines prior to the Viking Age and the culture of this area becomes very unclear in the archaeological record by ca. AD 1000 (e.g. Huurre 1979: 128–136, esp. Maps 15–17; Edgren & Törnblm 1993: 229–233), leaving the language spoken in this area uncertain. No ethnonym for this culture has been identified. There is the possibility that the people could have been called ‘Kvens’, insofar as Kvenland can seem to be a geographical territory of a fixed-settlement polity adjacent to Finnland to the north (Orkneyinga saga 1). However, the referents of Kvenr and Kvenland are so problematic that the possibility remains speculation (see F18). The ethnic and linguistic character of the Kyrö region is disputed, but Vahtola (1980) suggests that some Finnic toponymic types have spread from this area to the Tornio River valley in the north. record of encounters with these groups is that provided by Ohthere of Halogaland recorded in Old English in the late 9th century: Ohthere mentions different groups of Finnar (S) and contrasts them with the Bjarmians (see Valtonen 2008). References to the Bjarmians and Bjarmaland [‘Land of the Bjarmians’] are subsequently found in the histories and literature of Scandinavia in both Old Norse and Latin. The Bjarmians appear as a historical group or groups in the White Sea region, but they also became popular as a semi-fantastic ‘other’ dwelling at the periphery of the inhabited world, in which case they are depicted with a society and culture paralleling those of the Norsemen (see e.g. Jackson 1993; Koskela Vasaru 2008). Many accounts of the Bjarmians nonetheless appear ultimately rooted in historical encounters with fixed-settlement groups on the White Sea. The precise location of the Bjarmians is not consistent. The term Bjarmar is contrasted with Finnar, apparently to distinguish groups in the region according to fixed-settlement and mobile ways of life. As noted above, Finnar was used as a broad category that could include a variety of linguistic-cultural groups. The Norsemen only refer to two categories of culture in the White Sea region. Use of the term Bjarmar in this way makes it probable that it referred to just such a broad category of culture type without necessary reference to language. In other words, Norsemen would most likely refer to any fixed-settlement group on the White Sea as ‘Bjarmians’, whether they spoke a Finnic, Sámi or some other language. The primary grounds for identifying the Bjarmians as Finnic speaking are problematic. This identification has developed on suppositions that: a) the extent of the Finnic language area in the Viking Age was more or less the same as it is today; b) there were only two languages in the region, ‘Finnish’ and ‘Sámi’; and c) language can be directly correlated with culture-type. When these are premises, it is quite natural to conclude that the mobile Finnar were all Sámi-speaking and the fixed-settlement agriculturalist Bjarmar were ‘Finnish’-speaking. Ohthere reports that the Bjarmians spoke almost the same language as the Finnar Poorly Attested Finnic Groups In addition to those groups well attested in the historical sources that have offspring associated with documented Finnic languages, there is notable evidence on extinct and Russified Finnic groups both in the Scandinavian sagas as well as in the Russian chronicles. The toponymic investigation into Northern Russian nomenclature has corroborated that the spread of Finnic languages in the medieval period was substantially different from the present (cf. Matveev 2001–2007; Saarikivi 2006). F15: Bjarmians With the Viking Age, a sailing route opened to circumnavigate the northern coast of Norway and the Kola Peninsula to the White Sea. In this area, the Norsemen encountered groups with fixed settlements and practicing some form of agriculture there were referred to as Bjarmar, and have customarily been identified as Finnic-speaking. The earliest 98 settlement lifeways, then Finnic-speaking groups practicing the appropriate culture would be called Bjarmar – even if not all Bjarmar were necessarily Finnic-speaking. Several Old Norse sources situate the Bjarmians on the river Vína, which is generally identified with the Northern Dvina (see e.g. Jackson 2002: 7–8). An early and noteworthy Finnic-speaking population has been demonstrated in the Dvina basin on the basis of toponymic evidence (cf. Matveev 2001–2007; Saarikivi 2006). There are thousands of relatively well-studied substrate toponyms denoting both macro- as well as micro-objects in this region. They derive from at least two different types of substrate languages, namely Eastern Finnic languages and other, phonematically archaic West Uralic languages that exhibit resemblance to South Estonian and also to Sámi (Saarikivi 2007). Thus, irrespective of the exact cultural character of the Bjarmians, there should be no doubt that many parts of the Dvina basin, such as the Pinega basin (Saarikivi 2007), the Lower Dvina (Kabinina 2012) and the Vaga basins had substantial Finnic settlements. Presently, a rich oral tradition everywhere in the Dvina basin tells about the Čud’s – whiteeyed original inhabitants of the region who fought with the Novgorodians and, in some cases, escaped to more remote lands or were buried under ground (Saarikivi 2007: 11). Some parts of the Finnic settlement of this region may be of relatively recent origin, for instance, those ethnotoponyms pointing to Karelian immigration. However, the stratified character of the toponymy, as well as early Finnic–Permian contacts both point to a relatively early presence of a West Uralic language form bearing resemblance to Finnic and Sámi in this region (cf. Saarikivi forthcoming). In addition to the Scandinavian sources, Slavic literary sources are also suggestive of Finnic speakers in the Dvina basin in several connections, including chronicles, administrative documents and hagiographies. In the Primary Chronicle, the inhabitants of the area are called Zavolockaja Čud’ [‘Čud’ beyond the portage’, i.e. in the next water basin (presumably beyond the Baltic Sea basin)]. The denomination is derived in connection to (Finnas in OE; Sweet 1883: 17). This has been interpreted as a contrast between Sámi and Finnic as uniform languages (e.g. Jackson 1993). However, it could equally represent a contrast between different Proto-Sámi dialects or between different Uralic languages of the region. Two Old Norse texts present raids on temples of the god Jómali (infl. Jómala) among the Bjarmians.43 This theonym has long been considered a loan from a Finnic jumala [‘god’] and thus treated as evidence of the language spoken by the Bjarmians, which would be consistent with Ohthere’s comment on the Bjramian language (e.g. Koskela Vasaru 2012: 42).44 Outside of a few macrotoponyms of the region (e.g. the Vína River; cf. Ter- in the discussion of Sámi above), Jómali seems to be the only non-Norse name or term associated with the Bjarmians in these or other sources. The earlier of these sources was written more than two centuries after the events described, while the second is set in mytho-heroic time and quite probably dependent on the first (Koskela Vasaru 2008: 303–304). Both sources present the name in a popular mytho-heroic legend plot about a raid on a pagan temple (Power 1985). It is improbable that a Bjarmian theonym would circulate for centuries in the oral tradition when corresponding anthroponyms and toponyms did not. The theonym was most likely simply added to a popular plot in a historical saga to give it a more authentic touch (which is common for this literature), and then borrowed directly into the fantastic saga. The name Jómali may be a Finnic loan, but cannot be taken to indicate that the Bjarmians were Finnic-speaking. Instead, it suggests that the Bjarmians had a culture-type for which a fixed-structure temple and (exotic) Finnic god could be appropriate. The borrowing of Jómali itself most likely reflects contacts in the Baltic Sea region. (Frog 2014b: 443; cf. also Schalin with Frog 2014: 286–289.) Neither the comparison to the language of the Finnar by Ohthere nor the theonym Jómali in sagas provide reliable evidence that Bjarmians spoke a Finnic dialect. If the term Bjarmar is accepted as referring to cultures in the White Sea region according to fixed99 the ethnonym Čud’ that is also mentioned in the Primary Chronicle (cf. E1) as one of the constituent tribes of Russia. Dvinskaja letopis also mentions the Zavoločkaja Čud’ as a population residing in the Dvina area: the Zavoločkaja Čud’ were baptized and were then called the Dvinjane [‘Inhabitants of the Dvina’] (cf. Saarikivi 2006: 29 and works there cited). The Dvinjane would seem to correspond to the culture associated with the Finnic language toponymy of this area and can also be inferred to be the Bjarmar of the Vína River in Old Norse sources.45 At least some parts of the Finnic population in this region (those associated with Karelians and other clearly Western tribes) would be the result of migration. These can be contextualized in the movements of Finnic populations in the Viking Age associated with the rise of the fur trade as well as changes in climate that enabled agriculture in new areas (Ahola & Frog 2014: 47–48, 66– 67). However, the relatively large number of Western Finnnic borrowings into KomiZyryan and other Permian language forms suggests that such an explanation cannot account for the toponymic phenomenon in its totality (cf. Saarikivi forthcoming). If the identification of the Vína River with the Northern Dvina is correct, then the Finnic groups encountered here would most likely have been regarded as Bjarmar by the Norsemen. However, it should be stressed that the economic draw of these northern regions was not exclusive to any one language or culture. The toponymy shows a social prominence of Finnic dialects, but it should not be underestimated that Iron Age immigration to the region was most probably multiethnic from the outset. It also seems that, within a few centuries, Slavic languages became common in the region (Saarikivi 2006: 295). On the basis of phonetic similarity and proximity to the White Sea, Bjarmaland was identified with Old Russian Perm’ or Perem’ by scholars already in the 17th century (see further Koskela Vasaru 2008: 31–59). The Old Russian term is attested already from the 13th century and could be connected with a large geographical area extending from the White Sea to the Ural Mountains (Koskela Vasaru 2008: 394). An etymological relationship between the terms is not improbable, but the etymology itself remains opaque. Comparison has also been made with the Finnish term permi used to refer to traders coming from Viena Karelia (cf. Koskela Vasaru 2008: 43–44) – i.e. from the Russian side of the border of the region extending to the White Sea. All of these terms may be historically related through the multiethnic trade networks that connected the cultures of this region, but the etymology seems irresolvable. F16: Sura poganaja Old Russian chronicles also refer to the Pinežane and Sura poganaja [‘pagans of the Sura River’] inhabiting the river basin of a tributary of the Northern Dvina. The area has a rich layer of Finnic toponyms, and there is a notable oral tradition regarding the Čud’s of the area. There is no doubt of the Finnic character of the Iron Age settlement in the area even if the details are fairly scarce (on the historical sources, see Saarikivi 2006). F17: Važane and Additional Possible Finnic Settlement Areas The Važane are mentioned in Old Russian chronicles as inhabiting the Vaga River basin. The spread of Finnic language groups to river basins on the Baltic Sea and to the east suggest that additional settlements were also established in intermediate regions, but the toponymy of these areas has not yet been sufficiently explored. For this reason, the Važane and possible additional groups of this type are not identified with specific locations on the map. However, the Vaga region is especially rich in substrate toponyms of Finnic origin. Old maps also indicate that the river valley has been densely inhabited. A birch bark letter from 1315 mentions three persons with Finnic names from the Vaga region (Saarikivi 2007). F18: Kvens The Kvenir [‘Kvens’] in Old Norse46 (> Cwenas in Old English) are identified in various sources as travelling overland from the northern end of the Gulf of Bothnia to raid in northern Norway; in other sources, they appear to be on the eastern side of the Gulf.47 Since the 19th century, there has been a 100 saga, Orkneyinga saga). Just as Bjarmar were contrasted with Finnar on the White Sea, it may be that Kvenir were contrasted with Finnar on the Gulf of Bothnia, in which case the culture of the Kyrö River valley (F14), at a geographical remove from other groups in Finnland, could also have been called Kvenir living in Kvenland. During the early Middle Ages, Finnic groups gradually spread to the north both in the east and the west. This can be seen as a continuation of Finnic mobility and spread that characterized the Viking Age, but the precise mechanisms and chronologies as well as the degree of these processes are not clear (Ahola & Frog 2014: 56). The overland trade routes from Karelia (F8) may have resulted in more permanent settlement or the establishment of Finnic dialects on the Gulf of Bothnia.50 The identification of the Kvens as one or more Finnic language groups in ca. AD 1000 remains problematic, and a clear identification of Kvens in the archaeological record is lacking (cf. Kuusela 2014). However, the contact networks of trade with which this region was linked suggest that there was at least some penetration of Finnic language into the region in much the same way that Slavic and Scandinavian languages were penetrating into Uralic language areas elsewhere. The Kvens have therefore been listed as tentatively associated with the Finnic language here, although it is quite possible that the Cwenas mentioned as coming overland into Norway in the 9th century may have spoken a dialect of Sámi. The etymology of Old Norse Kvenr is obscure and problematic. Its similarity to Old Norse kven [‘woman’] is almost certainly coincidental. The phonetic similarity to Finnic Kainu- does not offer a good reconstruction of a loan from either language into the other (Koivulehto 1995). The Finnic term is generally argued to be a loan. The proposal that it derives from a Proto-Germanic (*χwainō) or later derivative word referring to ‘lowland’ has found appeal but is phonetically problematic; Jorma Koivulehto (1995) has made an interesting argument that the Finnic word derives from a Proto-Scandinavian *gain- and was applied in the sense of ‘gap’ to the northern part of the Gulf of Bothnia, presumption that these groups were Finnicspeaking. This interpretation seems to have been arrived at on the basis of the Kvens being distinguished from Finnar (interpreted as Sámi) and also from Norsemen, in which case the only other linguistic group inhabiting the region acknowledged at that time was Finnic (cf. F15). This evolved an etymological identification of Kvenr with the Finnish toponym Kainuu and ethnonym Kainulainen [‘Kainu-person, person of Kainuu’]. Although Kainuu now designates a region inland and to the east, it is argued to have originally designated a coastal territory of the northern part of the Gulf of Bothnia (see below). These groups can also be compared with the Kajan’ in Old Russian sources: the Kajan’ seem to have been in eastern areas at the northern end of the Gulf of Bothnia, at the outlet of river routes from Karelia. (Valtonen 2008: 386– 389.) Possible references to the same culture in Latin sources (e.g. as Adam of Bremen’s terra feminarum [‘land of women’]) have also been suggested, but these are problematic (Valtonen 2008: 76–77, 390n). The location of Kvenland [‘Land of the Kvens’] also seems inconsistent in the sources.48 The Norse toponym element -land suggests fixed settlement livelihoods. The difficulty may be that the Kvens were not strictly identified with any particular polity but were rather seen in contrast to the mobile groups referred to as Finnar. Use may thus have varied, but it is noteworthy that information about the Kvens does not seem to be recorded among groups who encountered them on the Baltic Sea. Whether use of the term varied in different dialects is less relevant here than whether the term, which is never distinguished through language in any source, would be identified with Finnic language groups.49 A crucial problem for identifying the Kvens with Finnic language groups is that there is little reason to believe that Finnicspeakers had spread to the northern end of the Gulf of Bothnia already in the 9th century in order to undertake the overland raids into Norway described by Ohthere and in Egils saga. On the other hand, some of these sources also situate Kvenland adjacent to Finnland along the Gulf of Bothnia (Egils 101 offering an account that is phonetically viable but semantically problematic (see Kylstra et al. 1997–2012: II: 11–12). appears in North Russian folklore in regions where Čud’s were known in the chronicles, rendering fantastic accounts of their disappearance from the region (Davydov 2009: 20–21). It has been used by parts of the Vepsian population as an endonym (Grünthal 1997: 152, cf. 150) and groups in the Northern Dvina River basin continue to distinguish themselves as Čud’s as opposed to Russians although they speak the Russian language (Saarikivi 2007: 11). A cognate in Sámi is often translated as e.g. ‘enemy’ and appears in folklore to refer to a people who subjected the Sámi to taxation and engaged with them in hostile conflicts and warfare (Saarikivi 2007: 11; cf. Itkonen 1948: 537–545). These tales may refer more specifically to ‘Russian Čud’s’ or ‘Karelian Čud’s’, and in northern Norway these are groups coming ‘from the east’ (see e.g. Drannikova & Larsen 2008). The term equally appears in Komi folklore to describe primal beings at the beginning of the world (Konokov et al. 2003: 57–58) and it has also been used of the old inhabitants of the Komi-speaking area that partly assimilated to Komi, or escaped to the other regions, in a similar manner to Čud’ in Russian folklore. The area east of Lake Peipsi (Russian Čudskoe ozero [‘Čud’ Lake’]) has been identified with the Čud’ tribe already in the Russian Primary Chronicle. Substrate place names of the area were first investigated in detail by Pauli Rahkonen (2011), who argues that this toponymy is attributable to a nonFinnic group speaking a West Uralic language. Rahkonen stresses the prominence of Čud’ as an element in the toponymy of the region. The toponymic material treated by Rahkonen is not of the best quality and he is somewhat inclined to construct Čud’ as a clear historical linguistic group and ethnos rather than simply applying the term as a practical tool to refer to a broad linguistic group as done here. Taking into account that the assimilation of the Čud’s east of Peipsi must have occurred relatively early, many of the substrate toponyms must have disappeared and it is probably not possible to study them without archive materials deriving from fieldwork. Rahkonen’s findings must therefore be considered to remain conditional on the degree to which the data used is indeed representative. (Extinct) Non-Finnic West Uralic Groups In addition to groups that can be identified with the Finnic language, several other groups can be identified in the historical sources with West Uralic languages that were probably not Finnic-speaking, although the precise relationship of their languages to surviving Uralic languages is not necessarily clear. E1: Čud’ For the purposes of the present discussion, Čud’ is used as a practical term to designate the cultural group or groups between clearly Finnic language areas and Novgorod and that were described as Čud’ in the chronicles without other ethnonymic identification. However, the ethnonym Čud’ and its cognates present a number of difficulties. The earliest potential attestation is from the topogeny in Jordanes’ Getica mentioned above (F8): Thiudos inaunxis. The term Thiudos is often viewed as referring to Čud’s (cf. Grünthal 1997: 153–154). If in here is a preposition, this would seem to read ‘Thiudos in Aunus’. If this is both correct and Aunus was in roughly the region it is today (which cannot be assumed), Jordanes would seem to situate the Thiudos somewhere in the vicinity of Lake Ladoga. However, the prepositional phrase could also be read with the following ethnonym Vas, as pointed out above (F9). An equivalent ethnonym does not seem to be found in Norse sources and later Latin materials. In Old Russian sources, Čud’ appears as broadly applied to a category of cultures whereby it could be used interchangeably with more specific ethnonyms for West Uralic groups. Different ethnic groups could also be distinguished as Čud’s of a particular type or location – e.g. Čud’ v Klinu (F10) or the Zavoločkaja Čud’ (F15). In this respect, the term functioned similarly to Old Norse terms like Finnar and probably Bjarmar and Kvenir. In later evidence, Čud’ and its cognates appear widely in folklore as an ethnonym for various groups of ‘others’ in many regions, a number of which can be identified as Finnicspeaking, as has been discussed with regard to the Zavolockaja Čud’ above (F15). The term 102 having an earlier sense of ‘strange, foreign’. Different forms of the term are found in all Sámi languages (Grünthal 1997: 164: Kulonen et al. 2005: 57). Evidence of the term in Sámi and also in Komi make it probable that the ethnonym was borrowed into Slavic dialects. Thiudos in Jordanes’ Latin exhibits a striking correspondence to Gothic þiuda [‘people’],51 which is derivative of Proto-Germanic þeudō [‘people’]. This would be quite natural as an endonym (cf. de Castro 1998). Insofar as such an ethnonym would not be expected east of the Baltic, this should be viewed with caution: it could also reflect a Germanic interpretation of a phonetically similar term in another language. The etymology of Čud’ should be left open. Nevertheless, Rahkonen (2011: 227) observes a geographical boundary in the distribution of hydronyms that can be assigned a clearly Finnic etymology, which suggests some type of marked linguistic distinction. This boundary roughly extends east from Lake Peipsi. This boundary in the toponymy also roughly corresponds to the boundary of the so-called Long Barrow Culture in the region (Rahkonen 2011: 209–210). The latter cultural phenomenon in the archaeological record penetrates only into the periphery of later Finnic language areas of Southwest Estonia or the edge of the Ugandi territory (cf. Tvauri 2012: 270, Fig. 199). It otherwise generally appears to extend to the south and east of unambiguously Finnic language areas. It has been a subject of debate whether the long barrows should be considered as culturally Slavic (as suggested by Sedov, e.g. 1997) or Finnic (as suggested by, for instance, Selirand or Tvauri, e.g. 2008). As the area is in the vicinity of many notable centres of Novgorod Slavs, it has likely been a subject of a relatively early Slavicization. V.L. Vasil’ev (2005) has convincingly argued that the Slavic habitation around Novgorod is fairly old on the basis of many old Slavic anthroponymic types. However, the spread of this type of burial practice into territories of south-eastern and eastern Estonia in the second half of the 6th century would antedate significant Slavic– Finnic language contacts by more than two centuries (cf. Kallio 2006b). The linguistic evidence does not support Slavic cultures extending into Finnic language areas and becoming their south-eastern neighbours significantly prior to the Viking Age. The etymology of the term Čud’ remains obscure. Outside of the possible hapax reference in Jordanes’ Getica, it has been proposed that the ethnonym has only had an established place in Sámi and Russian languages, of which cognates in other languages are derivative; however, the term cannot be reduced to a clear loan from either one of these two languages into the other (Grünthal 1997: 164–171). In Russian, the ethnonym becomes associated with the partial homonym čuda [‘wonder, miracle’] which has given rise to a folk-etymology of the term E2: Meryans The Merya is a tribe mentioned in the Old Russian literature on several occasions. This ethnonym has also been interpreted as the same as the Merens referred to by Jordanes among different Finno-Ugric peoples (F10).52 Already the Russian Primary Chronicle notes that the Meryans reside in the vicinity of Lake Plesceevo and Lake Nero. They are thereafter mentioned in the sources up to the 15th century and there is even a layer of archaeological findings attributed to Iron Age and medieval Meryans (Leont’ev 1997). Their relatively late survival appears interesting in light of the fact that the Meryans resided in the core area of the emerging Moscow principality, which subsequently developed into the core of modern Russia. Toponymic evidence shows that the Meryan language area extended to the Lake Beloye region (Makarova 2012). In the scholarly sources on Russian ethnohistory, it has usually been taken for granted that the Meryans were a Finno-Ugric tribe. Several attempts to reconstruct the Meryan language on the basis of toponyms, dialectal vocabulary and jargon lexicon have been made, and practically all of them consider Meryan as an independent extinct branch of Finno-Ugric (cf. Tkachenko 1985). A.K. Matveev (1996) stresses the many parallels between Meryan and Mari toponymy, whereas Arja Ahlqvist (2001) has considered Meryan to be more closely related to the Finnic languages. However, Ahlqvist also 103 points out some notable differences, such as the river names with a word-final element – gda. Tkachenko (1985) and others consider Meryan to be a more independent branch of Uralic, and Rahkonen (2012) points to a probable sound shift (Uralic *a > vo in anlaut) that connects Meryan to the Sámi languages. Most of the linguists who have considered the etymology of Merya, have connected it with the Indo-European borrowings in the Uralic languages indicating ‘people’, as in Mari ‘man’, Komi mort, Ud-murt ~ Fi. marras (: marta-) [‘person; mortal’]. On the basis of careful analysis of the early context of the Meryan people in the Central Russian lake district, Ahlqvist (1998) has argued for an alternative etymology from a compound meaning ‘people of the big lake(s)’ (the hydronym *Enä-järvi > Nero > MerV). inhabited is difficult to define. Rahkonen (2013) has argued on the basis of substrate toponymy that the Meščera were a Permian tribe. E5: Tojmicy poganaja The heathens of the Tojma River have been mentioned in the preserved fragments of the Slovo o pogibeli Russkoj zemli [‘Story of the Fall of the Russian Land’], where they are mentioned as people residing near the northeastern border of the Rus’ on both the Upper and Lower Tojma, tributaries of Northern Dvina in the vicinity of Komi. The Tojmicy have been considered a Permian tribe by A.I. Turkin (1989). There is, however, almost nothing known of this group. Conclusion Many of the groups above that have been placed on Map 2 are mentioned in the earliest historical sources related to the area of the present Finnic-speaking region. Some of these sources (some oral poetry of the Scandinavian sagas, Old English literature) antedate the estimated time period of the map, AD 1000, by some hundreds of years. Others (Russian and Latin chronicles, saga literature) are slightly later and mainly date from about AD 1200 and later. In addition, some groups have been included on the maps that are not to be identified on the basis of such early texts, but that have certainly existed as observable in concomitant archaeological material. The authors have made an effort to linguistically identify the groups under consideration on the basis of the research tradition, ethnonyms, toponymic data and, above all, historical descendant reconstruction of the later linguistic map. As far as is known, none of the literary sources referred to above have been written by the Finnic-speakers themselves. Thus, the groups portrayed above are, first and foremost, identified by representatives of the other, culturally dominant groups. The ethnonyms are thus exonyms used by merchants, priests, warriors, chronicle writers, etc. Although the explicit aim of the authors of this article was to provide a map of the linguistic groups, the languages of many of the groups treated above cannot be identified with great certainty. Moreover, there are many grounds to believe that, in fact, many of the groups mentioned in E3: Muroma Another tribe known from Old Russian sources to inhabit areas between modern branches of the Finno-Ugric languages are the Muroma. Less is known about the Muroma than about the Meryans. They are first mentioned in the Russian Primary Chronicle, where they are said to be the first inhabitants of the centre Murom (mentioned as Moramar in Old Norse) and to be more generally on the Oka River. They are said to pay tribute to the Rus’, but the geographical area they were associated with as well as their ethnic and linguistic character and possible cultural attributes all remain extremely hypothetical. Pauli Rahkonen (2013) has recently argued on the basis of toponymic evidence that Muroma was linguistically quite close to Meryan and that these could have been two dialects of the same language. However, such positions are highly problematic, not least because of the limitations of the sources concerning the historical groups. The question requires further investigation and the detailed analysis of archive materials deriving from fieldwork. It also may prove that definitive answers to this question are impossible. E4: Meščera The Meščera are another tribe known from Old Russian chronicles that inhabited territories that later underwent Slavicization. Little is known of this group and the area they 104 millennium and that the present (Map 1) and past (Map 2) linguistic maps do not resemble each other very closely. This observation is natural on the basis of analogies from other regions. One must still remember that the changes in the linguistic map between the approximately AD 1000 and the present occurred not only because of migrations and the displacement of earlier linguistic groups but, more notably, because groups changed their languages. This also reminds the reader that the groups presented on the map are far from uniform and stable: they have been subject to a continuous flux of people, multilingualism, and multi-language networking that have connected them with their neighbours. Second, the map reveals the wide Finnic and Uralic periphery to the east of Finland and Estonia that has been relatively little discussed in scholarly literature. Most of the discussion regarding the ethnogenesis of the Finns and Estonians has taken place in their respective countries, but this has been done by researchers more familiar with Germanic than non-Finnic Uralic languages. A notable amount of information regarding the Finnic languages east of the Baltic Sea area has thus been left without thorough scrutiny. If the Finnic language area today stretches some 1,200 kilometers from south to north in a stretch some 500 kilometers wide, in the Iron Age and early Middle Ages, the opposite has been the case: the language area has probably been more than 1,000 kilometers from east to west but substantially less so from south to north, while north of the Finnic area, Sámi languages and unidentifiable substrate languages of Northern Fennoscandia were spoken. Third, the map points to the fact that the historical centre of the Finnic languages, and the centre of Finnic linguistic diversity, is south of the Gulf of Finland. There is a long tradition of considering the whole region surrounding the Gulf of Finland as the ProtoFinnic homeland (cf. Itkonen 1983; Koivulehto 1983). Notable arguments against this hypothesis have been put forward since the turn of the new millennium (Aikio & Aikio 2001; Saarikivi 2006; Häkkinen 2009), as will be discussed in the next article in this series. The model presented here has concentrated on groups or populations that can be the medieval written sources were not linguistically uniform nor did they form monolingual speech communities or exclusive and clearly definable language areas. It is equallylikely that the languages spoken by a number of the groups in question have not been direct predecessors of presentday language forms. It is now often assumed in scholarly literature that the linguistic diversity in the past (especially among huntergatherers) has been greater than at present. This is an inevitable conclusion related to the absence of forces such as communications technologies, literary languages, formalized education and so forth that create and maintain linguistic similarity in the present. Many lines of development of Finnic, Sámi, Germanic, Baltic, Slavic and other languages have certainly also died out. Today, such extinct language forms are discernible mainly either as some peculiarities in dialects that have intruded as borrowings into the present language forms or as (for the most part littlestudied) linguistic substrates (cf. F9). Especially in the case of substrate languages of Northern Russia, it has been proposed that there have also been languages that represent characteristics of both Finnic and Sámi, which should be considered an independent course of development of Uralic languages that have gone extinct (cf. Saarikivi 2004b). It is likely that Meryan, Muroma, Meščera and other similar tribes mentioned in early Russian chronicles may have also represented similar extinct language groups. There are certainly many problems related to the identification of the groups on the Map 2, yet this cartographic visualization still highlights many aspects of linguistic history that are of importance for understanding of the present linguistic map of the eastern Baltic Sea area. The working model remains a heuristic sketch that has been developed to provide a broad framework for multidisciplinary discussion of culture formations in the past, a stetch to be refined and developed through future discussion. Although general and still hypothetical in many respects, this model nonetheless brings some significant points into the light. First, it underscores the degree to which language areas have changed during the last 105 reasonably identified as speakers of particular languages with varying degrees of probability. This has been possible in large part because the target period of the map is relatively close to numerous written sources referring to the groups and even to their languages. The farther removed from those sources reconstruction becomes, the more abstract, hypothetical and problematic it is to identify language with specific groups or cultures attested through other types of materials. The next articles in this series will address the longer-term perspective according to two interreated aims. The Urheimats of different branches of Uralic and how those can be situated on a map will be explored, locating Finnic and Sámi language families in relation to other Uralic and non-Uralic language families. Reverse-engineering the dispersal of Finnic languages and antecedent Proto-Finnic dialects will also be done with the goal of targeteting a particular period in that history for cartographic representation: ca. AD 1. The model developed for rendering Map 2 above provides essential information for considering parts of that reconstruction. east of the Baltic Sea and especially the later Christianization (see Ahola & Frog 2014). 2. The standardizations are in essence reconstructions that have become established in the scholarship. Especially in cases of infrequent terms, the reconstructed forms have been shaped by interpretations or by prioritizing certain sources: they require a thorough review. Here, uncertainty mainly concerns vowel quantities, which were often not indicated in early manuscripts (e.g. in Jómali), and some diphthongs (e.g. Kven- often appears with the vowel -æ- and Kirjál- with the diphthong transcribed -ia-). Variant transcriptions are not listed: without manuscript contexts and their relations, the forms and their representativeness may be misleading (e.g. “Kyrialar”, “Karelar” and “Kereliar” for Kirjálar). Examples of transcribed forms can be found under the entries of the DONP (N.B. – some variant forms appear under separate entries). 3. It may be noted that the earlier dominant research paradigm considered Proto-Finnic to be a phenomenon extending around the Gulf of Finland (cf. Itkonen 1983). In this connection, it was also assumed that the development of Finnic from ProtoUralic took place in the this area. This had the consequense that the areal linguistics of the Finnic language area was not considered relevant for the study of the early history of Finnic languages. 4. The opening of the Eastern Route was not simply a function of changes in trade activity in the North. It is also associated with political changes that produced a demand for furs in the south (Kovalev 2001) and changes in the geopolitical situation of intermediate regions that made trade routes through them viable (Talvio 2014: 132–133). 5. Evidence of Finnic material culture found in areas to the west suggests contacts and probably some degree of mobility from east of the Baltic (e.g. Raninen & Wessman 2014: 336). Place names with the elements Finn- [‘Finn-’] and Est- [‘Estonian’] in Roslagen, eastern Central Sweden, may indicate evidence of Finnic speakers in the region connected with maritime networks (Sjöstrand 2014: 110n), depending on the interpretation of the toponymic type. Even if the individual toponyms are only from the medieval period and later, there is no reason to believe that the mobility of Finnic groups from the east only began with immigration from Sweden to Finnic language areas rather than during Finnic language spread in the Iron Age. 6. Insofar as it is likely that the Svear sought, at a minimum, to extort taxes from groups in Finland, the fixed-settlement polities would be more comparable to Norse groups than to mobile Finnar or Sámi, and thus it becomes a question of whether and to what degree the systems applied to such Scandinavian communities might be transposed on non-Scandinavian communities. 7. For the present discussion, it is not necessary to resolve whether immigration to coastal Finland happened under motivation or encouragement of the king of the Svear (e.g. Hellberg 1987: 274; Haggrén 2011: 160–163) or was undertaken by Acknowledgements: The present article series developed in relation to the Viking Age in Finland (VAF) project. The work provided maps for the introduction to the volume Fibula, Fabula, Fact – The Viking Age in Finland (Ahola et al. 2014b) and other 2014 publications associated with the VAF and with the Retrospective Methods Network. These appeared prior to the present article, which had been initially prepared under the title “Reconstruction of Historical Language Distribution and Change East of the Baltic Sea (with Consideration of the Problem of Ethnic Identities)”. The present form of the work has developed considerably in relation to comments and constructive criticism from readers and peerreviewers, whose questions, discussion and valuable suggestions have greatly helped to strengthen it. Frog (mr.frog[at]helsinki.fi) Folklore Studies / FHKjT Department, P.O. Box 59 (Unioninkatu 38 A), 00014 University of Helsinki, Finland. Janne Saarikivi (janne.saarikivi[at]helsinki.fi) Helsinki Collegium for Advanced Studies, P.O. Box 4 (Fabianinkatu 24), 00014 University of Helsinki, Finland. Notes 1. The Viking Age in Scandinavia is customarily dated to ca. AD 800–1050. The VAF project led to the proposal of a calibrated dating for Finland to ca. AD 750–1250 on the basis of historical processes 106 peasants independently (Lindqvist 2002: 47; cf. also Sjöstrand 2014: 112–115). Emphasis is here placed on the implication that immigration, which spread through areas that had been depopulated during the Viking Age, was associated with geopolitical changes in the region associated with the Svear and can be assumed to have been initially within domains of Svear authority (and protection) in 12th-century Åland and Finland. The Swedishspeaking settlements of Estonia have been generally believed to have been established slightly later, following on the 13th-century Northern Crusades, which established Christian political authority in these territories. However, the development of these populations remains poorly researched and little understood. Felicia Markus (2004) has argued that the Scandinavian language population in Estonia may have been established already in the Viking Age (cf. B1). 8. Interpretations like ‘out-pourer of semen’ as a selfreference emphasizing virility remain speculation. Interpretation is further problematized by the fact that the ethnonym corresponds to a theonym identified as the mythic progenitor of the race (identified with Odin in later evidence: see e.g. de Vries 1956–1957 II: 41–42). The link to a theonym has led, for example, to interpreting the ethnonym as having religious significance (North 1997: 139– 143). A mythological frame of reference also multiplies the potential significations: effusion of fluid was symbolic of the expression and transfer of mythic knowledge, supernatural power and poetry, symbolized as an intoxi-cating drink (also associated with spittle/semen). 9. The island-name Jomala (in Swedish) has been interpreted as deriving from a Finnic jumala [‘god’], but this is not unproblematic. A similar toponymic type is attested in Finnic-speaking regions and in regions with Finnic substrate. On this etymology and the relevant place names in Swedish, see Sjöstrand 2014: 97–100; Schalin with Frog 2014: 284–289. 10. For a possible localized (Scandinavian) name in Åland that may have been established in the Viking Age, see Heikkilä 2014b: 310–311. 11. For discussion and a review of the literature, see Sjöstrand 2014; see also Ahola et al. 2014a: 229–238. 12. For two fundamentally different perspectives, especially in respect to the ethnic affiliation of the so called long-barrow graves, see e.g. Sedov 1994; Tvauri 2007. 13. Such loans may be fairly limited, but there are problems is assessing their precise number (Kallio 2006b: 163 and note). 14. This sort of mythologization of history to validate conquest can be compared to the description of Etruscan rulership in Roman histories or the Gotlanders’ account of their relation to the Svear mentioned above. 15. On the complexity of the issue of ethnicity behind the term Rus’, see Lind 2007; on the borrowing from a Finnic dialect into Slavic, see e.g. Heide 2006: 76; Schalin 2014: 428–429 and works there cited. 16. Duczko 2004: 193; cf. also Callmer 2000; Ahola et al. 2014a: 255, 257; on the Ålandic rite’s development in this region, see Callmer 1994; Tarsala 1998; Frog 2014a: 379–398. 17. The latter term is thought to render an Old Norse ‘Skiing-Finnar’, but such a term is not attested in Old Norse among the many uses of Finnar, although it is found in a number of texts in Latin and Greek from earlier centuries (see Valtonen 2008: 106–108, 246–248). 18. Populations and cultures had been moving between Finland and the Scandinavian Peninsula already for thousands of years (cf. Ahola et al. 2014: 230) and thus the spread of Sámi language in this way is not at all surprising. 19. Aikio (2012: 105–106) tentatively suggests that the rise of the Scandinavian fur trade was the probable socio-economic motivator for the spread of Sámi and the language shift of Palaeo-European speech communities. Although this may have been a relevant factor in some parts of the Scandinavian Peninsula, it should be observed that expansion of the fur trade seems to have been especially around the beginning of the Viking Age, which also seems to be the period a) when Scandinavian sailing routes began to reach as the White Sea region (Valtonen 2008), b) when Finnic groups were drawn to the same region, presumably for trade (Ahola & Frog 2014), c) when evidence in the archaeological record suggests a change in orientation of cultural activities from the coast on the northern Gulf of Bothnia to inland or to overland networks probably linked to trade (Kuusela 2014), and d) when the so-called ‘fur road’ of trade along the Eastern Route opened with transformative effects on the fur economy in the North (Kovalev 2001). The period identified as primary for the language shift on the basis of linguistic evidence would thus seem to conclude with the period when significant change in longdistance fur trade began reaching especially the more remote regions where Sámi became dominant. Aikio’s proposal that the fur trade may have been a socio-economic factor in the processes of language change is not unwarranted, but viewed in this light, the Viking Age may have been a catalyst in the gradual eclipse of other languages by Proto-Sámi dialects. 20. See Kivimäe (2011: 95–96), noting that there are places where Estones is used as a broader blanket term with reference to Oeselians in addition to other groups (e.g. the conflict described in HCL XV.2– 5), but this seems to be practically motivated by the context. Henry also makes reference to Vinlandia [‘Finland’] as the place from which another priest has come (e.g. HCL XIX.4). 21. Ptolemy refers to the Leuonoi as the people inhabiting the centre of Scania, but the ethnonym appears to be based on a misreading or misspelling of Tacitus’s Suiones (G1), whose work seems to have been a source (Valtonen 2008: 80–81). Pliny 107 the Elder’s description of the Hilleviones inhabiting 500 villages or districts of Scantinavia has been interpreted as a corruption of *Levoni, but this interpretation is based on comparison with Ptolomy’s Leuoni (Valtonen 2008: 69 and works there cited). 22. In Ynglinga saga 32, Aðalsýsla seems to be presented as a part of Eistland. This is in the elaboration of a rather ambiguous verse in which the king attacks Sýslu kind [‘kin of Sýsla’] and is defeated by an eistneskr herr an [‘Estonian army’] (not necessarily Sýslu kind). The identification of Aðalsýsla with Eistland here may be the outcome of trying to interpret the verse and an etymologization of Old Norse sýsla as ‘district’ (see below). 23. Eysýsla could thus mean ‘Island-Business’, which would conform to the economic activities of the Oeselians, but normally ey would be the second element in such a compound and this interpretation would not account for another location as ‘Business Proper’ (Aðalsýsla). 24. There are, for example, seven sýslur in the Faeroe Islands (half the size of Saaremaa); on sýslur, see further Andersen et al. 1972. 25. In some historical contexts, the name Karjala would also seem to point to the old fortress of Karela (drevnjaja Korela) by what is now the town of Priozersk in Priladozh’e (cf. below). 26. This is attributed to Karelians in the 14th century Rhymed Chronicle of Erik, which also appears to reference their alliance with Novgorod in that context (Klemming 1865: 17). It is not improbable that the Papal request for trade embargoes against the ‘un-Christian Russians’ until they stopped causing trouble for the ‘Christian Finns’ refers to activities also involving Karelians (FMU, 74–76). 27. This would produce χ > k, after which χ > h (i.e. **harja). 28. Napolskikh 2006; 2014; on early lists of ethnic groups linked to trade routes, see also Valtonen 2008. This part of the text was interpreted as a list of ethnonyms and locations already in 1882, although identifying Thiudos with Čud’s and without viewing the list as an ordered topogeny (see the review of interpretations in Christiansen 2002: esp. 164–168, 176–177). 29. Comparison with the Old Norse personal name Vísinn and Latin version Wisinnus found in Saxo Grammaticus’ Gesta Danorum (e.g. Grünthal 1997: 105–106; Valtonen 2008: 367) is based on a reading of the ethnonym in Jordanes as Vasina. The Old Norse personal name (and its feminine counterpart Vísna) can be better explained on the basis of other possible Old Norse lexemes than on the basis of an otherwise unattested ethnonym. 30. For a brief historical overview of the culturally dynamic region of Ingria, see Nenola 2002: 54–58. 31. The ethnonym is found in a single chapter of Egils saga, where the Kylfingar (like the Kirjálar in another chapter of the saga) have arrived on the Scandinavian Peninsula from the east at the nothern part of the Gulf of Bothnia and are defeated by the hero’s men. It also appears as an element in a toponym in a geographical treatise in which Kylfingaland [‘Land of the Kylfingar’] is identified with Garðaríki – i.e. Novgorod (Einarsson 2003: 15). The ethnonym Kolbjagi in Old Russian sources and Greek Κούλπιγγοι (Kūlpingoi) mentioned in Byzantine sources in association with the Varangians seem to reflect adaptations of the Scandinavian term, whether it is an ethnonym per se or designates some other form of league or society (see Koivulehto 1997: 152–155). 32. Although this is generally accepted today, the topic was surrounded by political controversy in the past owing to the desire of the Swedish-speaking population of Finland to construct a heritage with linguistic and cultural continuity in Finland going back to the Viking Age. 33. Traditions of a so-called First Finnish Crusade appear to be retrospective constructs, but some degree of organized missionary activity could potentially have been initiated already in the first quarter of the 12th century, and certainly in the second half of that century (when place names borrowed into Swedish are established), with the bishopric of Finland being founded by or around AD 1200 (Line 2007: ch. 12; Sjöstrand 2014: 95). 34. The ethnonym Finnlendingar is not attested in prose (it receives no entry in the DONP) and could have been produced for metrical reasons (cf. Schalin 2014: 422–425). It is nevertheless relevant as evidence of the toponym Finnland from which it is formed, which in this case is found in verses referring to contemporary events there. 35. Some references that might be identified with Finnland are problematized by ambiguous use of the ethnonym Finnar/Finnas, as in the reference to Finna land [‘land of the Finnas’] in Beowulf (Valtonen 2008: 207–210). 36. Recent research suggests some degree of probably Finnic-speaking habitation during the Viking Age, on which see Alenius et al. 2014. Johan Schalin has also been working on the question of the geographical scope of Old Swedish Finland at least back to the mid-12th century. 37. Suomi is also later found as a personal name in Finnish dialectal sayings and proverbs. 38. Jukka Korpela (2008b: 20) has suggested a possible reference with the obscure ethnonym Jem’, but the Jem’ and their location remain mysterious and problematic (see Korpela 2008a: 44–45). 39. The second vowel is unwritten in the runic inscription (reconstructed for the dialect asTaf[æi]staland) and is later attested in the Hauksbók redaction of Ǫrvar-Odds saga in the (West Norse) form Tafeistaland among a list of toponyms for non-Scandinavian places in the Baltic Sea region (Simek 1990: 343). Part of the runic inscription was earlier interpreted as referring to Tafeistaland being subject to the military conscription or taxation system of Sweden, but this reading is problematic and the inscription may instead indicate martial conflicts (see Williams 2005). 108 locations. This reading would be more compelling if it were corroborated by evidence that Gothic words with immediate Latin equivalents were used more widely in Getica. 52. However, the ethnonym Merya is similar to terms for other groups such as Muroma (E3) and the attested groups Mari and Mordvin slightly farther to the east. Although Merens appears to correspond most closely to Merya, the specific identification remains speculative. The term Merens nevertheless resembles ethnonyms of Finno-Ugric groups concentrated in this region. 40. The Latin toponym is not a straight loan from Old Norse per se, which would presumably yield **Tavastalandus (cf. Finlandus). Instead, it can be seen as formation for the realm of an ethnos with the ending -ia in the place of Scandinavian -land (cf. Old Norse Bjarmaland appearing as Biarmia). If the ethnonym were derivative of the Latin toponym, the people would be **Tavastiani rather than Tavasti (cf. Oselia and Oseliani). 41. Unto Salo (2000: 49) has also argued on the basis of more recent oral history that the earliest groups in the region to practice slash-and-burn agriculture in Satakunta were Sámi speaking, but the period in question is so remote from the oral sources that use of this evidence is problematic in the extreme. 42. Cf. also the question of genetic evidence in Salama 2014: 357. 43. Óláfs saga helga [‘The Saga of St. Óláfr’] (Aðalbjarnarson 1941–1951 II: 294) and Bósa saga ok Herrauðs [‘The Saga of Bósi and Herrauðr’] (Jiriczek 1893: 25, 29); we would like to thank Bent Chr. Jacobsen (e-mail 28.12.2007) at the Dictionary of Old Norse Prose for confirming that this term has not been identified in any additional Old Norse texts according to their archive index. 44. The inference has seemed reasonable insofar as the development *juma [‘sky, god’] > *juma-la [‘god’] is only attested in Finnic languages. 45. This theory was first put forward in the 19th century; see Koskela Vasaru 2008: 43 and works there cited. 46. As observed in note 2 above, the form Kvænir is commonly found in West Norse manuscripts. 47. They also appear in exceptional contexts, such when it is stated in Norna-Gests þáttr that the king of the Svear had to oppose threats from the Kvenir and Kúrir [‘Curonians’ (B1)]. Although the specific account is certainly fictionalized, it raises the question of whether the Svear were thought to have subordinated the Kvens at times as they did the Curonians. 48. On different theories of the location of Kvenland, see e.g. Vilkuna 1957; Julku 1986. 49. For an investigation of the routes across the Scandinavian Peninsula from the point of view of Sámi toponyms pointing to kainu ~ kvens, see Korhonen 1982; but see however Koivulehto 1995). 50. The 13th century annals reporting Kirjálir harrying in northern Norway (F8) suggests that, at least by that time, Karelians had been taking up activities earlier associated with Kvens. It is interesting that in the annals Karelians are mentioned with Kvens or only Karelians are mentioned, which may indicate a change in the significance and/or recognisability of Kvens in this role. 51. Most recently, Napolskikh (2006; 2014) has revived this position and the interpretation that it represents an inflected common noun, reading it as introducing the list of ‘peoples’ rather than as an ethnonym. The latter interpretation has the benefit of reading the subsequent three ethnonyms as each paired with a location whereas reading Thiudos as an ethnonym presents four ethnonyms to three Works Cited Sources Bósa saga ok Herrauðs = Jiriczek 1893. DONP = Dictionary of Old Norse Prose. University of Copenhagen. Available at: http://onp.ku.dk/english/. Dvinskaja letopis. See Old Russian Chronicles. Egils saga = Einarsson 2003. Fagrskinna = Jónsson 1902–1903. FMU = Hausen 1910–1935. G 319 = Runic inscription in Rute, Gotland, Sweden, ca. AD 1200–1250. Available at: https://abdn.ac.uk/ skaldic/db.php?table=mss&id=18759&if=runic. Germania = Schweizer-Sidler & Schwyzer 1902. Getica = Mommsen 1882. Gs 13 = Runic inscription, Söderby, Gästrikland, Sweden, v.m. 11th century. Available at: https://abdn.ac.uk/ skaldic/db.php?id=18358&if=runic&table=mss&va l=&view=. HCL = Heinrici chronicon Lyvoniae: Arndt 1874. Henry of Livonia, see HCL. Liber census Daniae = Gallén 1993: 50–53. Norna-Gests þáttr = Bugge 1864. Óláfs saga helga = Aðalbjarnarson 1941–1951 II. Óláfs saga Tryggvasonar = Aðalbjarnarson 1941–1951 I: 225–372. Old Russian chronicles, see PSRL. Orkneyinga saga = Nordal 1913–1916. Orosius = Sweet 1883. Primary Chronicle = Ostrowski 2003. Saxo Grammaticus, Gesta Danorum = Olrik & Ræder 1931. Sigv Víkv = skaldic poem Víkingarvísur by Sigvatr Þorðarson, 11th century. PSRL = Polnoe sobranie russkix letopisej. Leningrad: Akademija Nauk SSSR, 1926–2000. SN = Mikkonen & Paikkala 2000. Sö 39 = Runic inscription, Åda, Södermanland, Sweden. Available at: https://abdn.ac.uk/skaldic/db. php?id=15981&if=runic&table=mss&val=&view=. SPK = Suomalainen paikannimikirja. Ed. Sirkka Paikkala et al. Kotimaisten kielten tutkimuskeskuksen julkaisuja 146. Helsinki: Karttakeskus, Kotimaisten Kielten Tutkimuskeskus, 2007. SSA = Suomen sanojen alkuperä: Etymologinen sanakirja I–III. 1992–2000. Helsinki: Suomalainen Kisjallisuuden Seura. U 467 = Runic inscription, Tibble, Uppland, Sweden. Available at: https://abdn.ac.uk/skaldic/db.php? id=17291&if=runic&table=mss&val=&view=. 109 Etymological Evidence Erom Lohjansaari Island, Western Uusimaa, Finland”. Journal of Archaeological Science 47: 99–112. Andersen, Per Sveaas, Troels Dahlerup, Jarl Gallén, Jan Liedgren & Björn Þorsteinsson. 1972. “Syssel”. In Kulterhistoriskt Lexikon fär Nordisk Medeltid XVII. Malmö: Allhelms Färlag. Cols. 645–651. Andersson, Thorsten. 1996. “Göter, goter, gutar”. Namn och bygd 84: 5–21. Andersson, Thorsten. 2007. “Rus’ und Wikinger”. Arkiv för Nordisk Filologi 122: 5–13. Androshchuk, Fjodor. 2008. “The Vikings in the East”. In Brink with Price 2008: 517–542. Anthony, David. 2007. The Horse, the Wheel and Language: How Bronze Age Riders from the Eurasian Steppes Changed the Modern World. Princeton, N.J.: Princeton University Press. Anttila, Raimo. 1998. Historical and Comparative Linguistics. Amsterdam / Philadelphia: John Benjamins. Arndt, Wilhelm. 1874. Heinrici chronicon Lyvoniae. Monumenta Germaniae Historica, Scriptores 23. Hannover: Hahnia. Barth, Fredrik. 1998 [1969]. “Ethnic Groups and Boundaries”. In Ethnic Groups and Boundaries: The Social Organization of Culture Difference. Ed. Fredrik Barth. Long Grove: Waveland Press. Pp. 9–38. Björklöf, Sofia. 2012. “Viron rantamurteen sanaston alkuperä suomalaislainojen valossa”. Masters thesis. Suomalais-Ugrilainen Kielentutkimus, University of Helsinki. Blommaert, Jan, & Ben Rampton. 2012. “Language and Superdiversity. MMG Working Paper 12-09. Göttingen: Max Planck Institute for the Study of Religious and Ethnic Diversity. Available at: http://pubman.mpdl.mpg.de/pubman/item/escidoc:1 615144/component/escidoc:1615143/WP_1209_Concept-Paper_SLD.pdf. Boiko, Kersti. 1993. Baltijas jūras somu eogrāfiskie apelat vi un to relikti Latvijas vietvārdos – BaltoFinnic Geographic Words and Their Relicts in Toponymy of Latvia – Ostseefinnische georgaphische Appellativa und ihre Relikte unter Ortsnamen Lettlands. R ga: Latvijas Zin tņu Akad mija. Brink, Stefan. 2008. “Naming the Land”. In Brink with Price 2008: 57–66. Brink, Stefan, with Neil Price (eds.). 2008. The Viking World. New York: Routledge. Pp. 57–66. Bugge, Sophus (ed.). 1864. “Söguþáttr af NornaGesti”. In Det norske oldskriftselskabs samlinger VI: Norrøne skrifter af sagnhistorisk inhold. Christiania: Brøgger & Christie. Pp. 47–80. Callmer, Johan. 1994. “The Clay Paw Burial Rite of the Åland Islands and Central Russia: A Symbol in Action”. Current Swedish Archaeology 2: 13–46. Callmer, Johan. 2000. “The Archaeology of the Early Rus’ c. A.D. 500–900”. Medieaval Scandinavia 13: 7–63. Christensen, Arne Søby. 2002. Cassiodorus, Jordanes and the History of the Goths: Studies in a Migration Myth. Copenhagen: Museum Tusculanum Press. U 582 = Runic inscription, Soderby-Karls k:a, Uppland, Sweden. Available at: https://abdn.ac.uk/ skaldic/db.php?id=17406&if=runic&table=mss&va l=&view=. U 722 = Runic inscription, Lots k:a, Uppland, Sweden. Available at: https://abdn.ac.uk/skaldic/db.php? id=17539&if=runic&table=mss&val=&view=. Ynglinga saga = Aðalbjarnarson 1941–1951 I: 9–83. Literature Aalto, Sirpa. 2014. “The Finnar in Old Norse Sources”. In Ahola et al. 2014c: 199–226. Aðalbjarnarson, Bjarni (ed.). 1941–1951. Heimskringla I–III. Íslenzk Fornrit 26–28. Reykjavik: Hið Íslenzka Fornritafélag. Ahlqvist, Arja. 1998. “Merjalaiset: Suurten järvien kansaa”. Virittäjä 102(1): 24–55. Ahlqvist, Arja. 2001. “Субстратная топонимия Ярославского Повольжя”. In Очерки исторической географии. Северо- Запад России: Славяне и финны. Ed. А.С. Герд & Г.С. Лебедеева. Санкт-Петербург: Издательство СанктПетербургского университета. Pp. 436–467 Ahola, Joonas, & Frog. 2014. “Approaching the Viking Age in Finland: An Introduction”. In Ahola et al. 2014b: 21–84. Ahola et al. 2014a = Joonas Ahola, Frog & Johan Schalin. “Language(s) of Viking Age Åland: An Irresolvable Riddle?”. Ahola et al. 2014c: 227–265. Ahola et al. 2014b = Joonas Ahola & Frog with Clive Tolley (eds.). 2014. Fibula, Fabula, Fact – The Viking Age in Finland. Studia Fennica Historica 18. Helsinki: Finnish Literature Society. Ahola et al. 2014c = Joonas Ahola, Frog & Jenni Lucenius (eds.). 2014. In The Viking Age in Åland: Insights into Identity and Remnants of Culture. Annales Academiae Scientiarum Fennicae Humaniora 372. Helsinki: Academia Scientiarum Fennica. Aikio, Ante. 2004. “An Essay on Substrate Studies and the Origin of Sámi”. In Etymologie, Entlehnungen und Entwicklungen: Festschrift für Jorma Koivulehto zum 70. Geburtstag. Mémoires de la Société Neophilologique de Helsinki 63. Helsinki : Société néophilologique. Pp. 5–34. Aikio, Ante. 2007. “The Study of Saami Substrate Toponyms in Finland”. In Borrowing of Place Names in the Uralian Languages. Ed. Janne Saarikivi & Ritva Liisa Pitkänen. Onomastica Uralica 4. Debrecen / Helsinki. Aikio, Ante. 2009. The Saami Loanwords in Finnish and Karelian. Oulu: University of Oulu. Available at: http://cc.oulu.fi/~anaikio/slw.pdf (last accessed 12.01.2013). Aikio, Ante. 2012. “An Essay on Saami Ethnolinguistic Prehistory”. In Grünthal & Kallio 2012: 63–117. Aikio, Ante, & Aslak Aikio. 2001. “Heimovaelluksista jatkuvuuteen: Suomalaisen väestöhistorian tutkimuksen pirstoutuminen”. Muinaistutkija 2001(4): 2–21. Alenius, Teija, Georg Haggrén, Markku Oinonen, Antti Ojala & Ritva-Liisa Pitkänen. 2014. “The History of Settlement on the Coastal Mainland in Southern Finland: Palaeoecological, Archaeological, and 110 Häkkinen, Jaakko. 2009. “Kantauralin ajoitus ja paikannus: Perustelut puntarissa”. SuomalaisUgrilaisen Seuran Aikakauskirja 92: 9–56. Häkkinen, Jaakko. 2010. “Jatkuvuusperustelut ja saamelaisen kielen leviäminen”. Muinaistutkija 2010(1): 19–36; 2010(2): 51–64. Haspelmath, Martin, & Uri Tadmor (eds.). 2009. The Loanwords in the Worlds Languages: A Comparative Handbook. New York: de Gruyter. Hausen, Reinh. (ed.). 1910–1935. Finlands medeltidsurkunder I–VIII. Helsingfors: Kejserliga Senatens Tryckeri. Available at: http://extranet. narc.fi/DF/index.htm. Hedenstierna-Jonson, Charlotte. 2009. “Rus’, Varangians and Birka Warriors”. In The Martial Society: Aspects of Warriors, Fortifications and Social Change in Scandinavia. Ed. L. Holmquist Olausson & M. Olausson. Stockholm: Archaeological Research Laboratory, Stockholm University. Pp. 159–178. Heide, Eldar. 2006. “Rus ‘Eastern Viking’ and the víking ‘Rower Shifting’ Etymology”. Arkiv för Nordisk Filologi 121: 75–77. Heide, Eldar. 2008. “Viking, Week, and Widsith: A Reply to Harald Bjorvand”. Arkiv för Nordisk Filologi 123: 23–28. Heikkilä, Mikko. 2014a. Bidrag till Fennoskandiens språkliga förhistoria i tid och rum. Helsinki: Unigrafia. Heikkilä, Mikko. 2014b. “The Changing Language Situation in the Åland Islands and Southwest Finland during the Late Iron Age”. In Ahola et al. 2014c: 303–321. Heininen et al. 2014a = Lassi Heininen, Joonas Ahola & Frog. 2014. “Geopolitics in the Viking Age? – Actors, Factors and Space”. In Ahola et al. 2014b: 296–320. Heininen et al. 2014b = Lassi Heininen, Jan Storå, Frog & Joonas Ahola. 2014. “Geopolitical Perspectives on Åland in the Viking Age”. In Ahola et al. 2014c: 323–348. Hellberg, Lars. 1987. Ortnamnen och den svenska bosättningen på Åland. 2nd edn. Studier i Nordisk Filologi 68. Helsingfors: Svenska litteratursällskapet i Finland. Holopainen, Sampsa. 2014. “Indoiranilais-uralilaisten kontaktien tarkastelua hantin, mansin, saamen ja samojeidn sanaston pohjalta”. Master’s thesis. Suomalais-Ugrilainen Kielentutkimus, University of Helsinki. Huurre, Matti. 1979. 9000 vuotta Suomen esihistoriaa. Helsinki: Otava. Itkonen, T.I. 1948. Suomen lappalaiset vuoteen 1945 I– II. Porvoo: WSOY. Itkonen, Terho. 1983. “Välikatsaus suomen kielen juuriin”. Virittäjä 87: 190–229, 349–386. Jackson, Tatjana N. 1993. “The North of Eastern Europe in Early Nordic Texts: The Study of PlaceNames”. Arkiv för Nordisk Filologi 108: 38–45. Jackson, Tatjana N. 2002. “Bjarmaland Revisited”. Acta Borealia 19: 165–179. Jiriczek, Otto Luitpold (ed.). 1893. Die Bósa-Saga in zwei Fassungen. Strassburg: Karl J. Trübner. Davydov, Alexander N. 2009. “Forest as a Phenomenon of Spiritual Culture”. In The Last Large Intact Forests in Northwest Russia: Projection and Sustainable Use. Ed. Tor Kristian Spidsø & Ole Jakob Sørensen. Copenhagen: Nordic Council of Ministers. Pp. 17–25. de Castro, Eduardo Viveiros. 1998. “Cosmological Deixis and Amerindian Perspectivism”. Journal of the Royal Anthropological Institute 4(3): 469–488. Drannikova, Natalia, & Roald Larsen. 2008. “Representations of the Chudes in Norwegian and Russian Folklore”. Acta borealia 25(1): 58–72 Duczko, Wladyslaw. 2004. Viking Rus: Studies on the Presence of Scandinavians in Eastern Europe. Leiden: Brill. Edgren, Torsten, & Lena Törnblom. 1993. Finlands historia I. Esbo: Schildts. Einarsson, Bjarni (ed.). 2003. Egils saga. London: Viking Society for Northern Research. Ferguson, Robert. 2009. The Hammer and the Cross: A New History of the Vikings. Harmondsworth: Penguin. Fox, Anthony. 1995. Linguistic Reconstruction: An Introduction to the Theory and Method. Oxford Textbooks in Linguistics. Oxford: Oxford University Press. Frog. 2013. “Shamans, Christians, and Things in between: From Finnic–Germanic Contacts to the Conversion of Karelia”. In Conversions: Looking for Ideological Change in the Early Middle Ages. Ed. Leszek Słupecki & Rudolf Simek. Studia Mediaevalia Septentrionalia 23. Vienna: Fassbaender. Pp. 53–98. Frog. 2014a. “From Mythology to Identity and Imaginal Experience: An Exploratory Approach to the Symbolic Matrix in Viking Age Åland”. In Ahola et al. 2014c: 349–414. Frog. 2014b. “Myth, Mythological Thinking and the Viking Age in Finland”. In Ahola et al. 2014b: 437–482. Gallén, Jarl. 1993. Franciskansk expansionsstrategi i Östersjön. Ed. Johan Lind. Helsingfors: Svenska litteratursällskapet i Finland. Grünthal, Riho 1997. Livvista liiviin: Itamerensuomalaiset etnonyymit. Castreanumin Toimitteita 51. Helsinki: University of Helsinki. Grünthal, Riho, & Petri Kallio (eds.). 2012. A Linguistic Map of Prehistoric Northern Europe. Mémoires de la Société Finno-Ougrienne 266. Helsinki: Société Finno-Ougrienne. Gustavsson, Rudolf, Jan-Erik Tomtlund, Josefina Kennebjörk & Jan Storå. 2014. “Identities in Transition in Viking Age Åland?”. In Ahola et al. 2014c: 159–186. Hackman, Alfred Leopold Fredrik. 1905. Die ältere Eisenzeit in Finnland I: Die Funde aus den fünf ersten Jahrhunderten n. Chr. in Finnland. Helsingfors: [Hackman]. Haggrén, Georg. 2011. “Colonization, Desertion and Entrenchment of Settlements in Western Nyland ca. 1300–1635 AD”. Iskos 19: 152–179. 111 Ostseefinnischen. Suomalais-Ugrilaisen Seuran Soimitteita 185. Helsinki: Suomalais-ugrilainen Seura. Koivulehto, Jorma. 1995. “Ala-Satakunnan Kainu ja pohjoisen Kainuu”. In Kielen ja kulttuurin Satakunta: Juhlakirja Aimo Hakasen 60vuotispäiväksi 1.11.1995. Vammala: Turun Yliopisto. Pp. 71–104.” Koivulehto, Jorma. 1997. “Were the Baltic Finns ‘Clubmen’? – On the Etymology of Some Ancient Ethnonyms”. In You Name It: Perspectives on Onomastic Research. Ed. Ritva Liisa Pitkänen & Kaija Mallat. Studia Fennica Linguistica 7. Helsinki: Finnish Literature Society. Pp. 151–169. Kolpakov, E.M., & E.N. Ryabtseva. 1994. A New Type of Chud Burial Construction. Fennoscandia Archaelogica 11: 77–86. Komjagina, L.P. 1994. Leksicheskij atlas Arkhangelskpoj oblasti. Arkhangelsk: Izdatelstvo Pomorskogo universiteta im. Lomonosova. Konakov, N., et al. 2003. Komi Mythology. Encyclopedia of Uralic Mythologies 1. Budapest: Académiai Kiádo. Korhonen, Olavi. 1982. Samisk-finska båttermer och ortnamnselement och deras slaviska bakgrund: En studie i mellanspråklig ordgeografi och mellanfolklig kulturhistoria. Umeå: Dialekt-, ortnamns- och folkminnesarkivet i Umeå. Korpela, Jukka. 2008a. “North-Western ‘Others’ in Medieval Russian Chronicles”. Proceedings of Petrozavodsk State University N2(93): 42–55. Korpela, Jukka. 2008b. The World of Ladoga: Society, Trade, Transformation and State Building in the Eastern Fennoscandian Boreal Forest Zone c. 1000–1555. Berlin: Lit. Korpela, Jukka. 2012. “Migratory Lapps and the Population History of the Eastern Finns: The Early Modern Colonization of Eastern Finland Reconsidered”. In Networks, Interaction and Emerging Identities in Northern Fennoscandia and beyond: Papers from the Conference Held in Tromsö, October 13–16, 2009. Ed. Chatlotte Damm & Janne Saarikivi. Mémoires de la Société FinnoOugrienne 265. Helsinki: Société Finno-Ougrienne. Pp. 241–261. Kortlandt, Frederik. 1977. “Historic Laws of Baltic Accentuation”. Baltistica 13(2): 319–330. Koskela Vasaru, Mervi. 2008. “Bjarmaland”. Unpublished PhD dissertation, University of Oulu. Koskela Vasaru, Mervi. 2012. “Bjarmaland and Interaction in the North of Europe from the Viking Age until the Early Middle Ages”. Journal of Northern Studies 6(2): 37–58. Kulonen, Ulla-Maija, Irja Seurujärvi-Kari & Risto Pulkkinen (eds.). 2005. The Saami: A Cultural Encyclopaedia. Helsinki: Finnish Literature Society. Kuusela, Jari-Matti. 2014. “From Coast to Inland: Activity Zones in North Finland during the Iron Age”. In Ahola et al. 2014b: 219–241. Kuz’min 2008 = Кузьмин С.Л. 2008. Ладога в эпоху раннего средневековья (середина VIII — начало XII в.) In Исследования археологических памятников эпохи средневековья. СПб. Pp. 69–94. Joki, Aulis. 1973. Uralier und Indogermanen: Die älteren Berührungen zwischen den uralischen und indogermanischen Sprachen. Helsinki: SuomalaisUgilainan Seura. Julku, Kyösti. 1986. Kvenland – Kainuunmaa. Studia Historica Septentrionalia. Jyväskylä: Gummerus. Jónsson, Finnur (ed.). 1902–1903. Fagrskinna – Nóregs kononga tal. København: S.L. Møllers. Kabinina, Nadežda. 2012. Substratnaia toponimiia Arkhangel'skogo Pomor'ia. DrHab thesis. Ekaterinburg: Izdatelstvo Ural’skogo gosudarstvennogo universiteta. Kallasmaa, Maarja. 1996–2000. Saaremaa kohanimed I–II. Tallinn: Eesti Keele Instituut. Kallio, Petri. 1998. “Suomi(ttavia etymologioita)”. Virittäjä 1998(4): 613–620. Kallio, Petri. 2006a. “On the Earliest Slavic Loanwords in Finnic”. In The Slavicization of the Russian North: Mechanisms and Chronology. Ed. Juhani Nuorluoto. Slavica Helsingiensia 27. Helsinki: Helsinki University Press. Pp. 154–166. Kallio, Petri. 2006b. “Suomen kantakielten absoluuttista kronologiaa”. Virittäjä 110(1): 2–25. Kallio, Petri. 2012. “The Prehistoric Germanic Loanword Strata in Finnic”. In Grünthal & Kallio 2012: 225–238. Kallio, Petri. 2014. “The Diversification of ProtoFinnic”. In Ahola et al. 2014b: 155–168. Kallio, Petri. Forthcoming. “Substrates in Finnic”. In Substrate Languages in Northern Europe: Case Studies and Methodological Perspectives. Ed. Ante Aikio & Santeri Palviainen. Studies in Language Change. Berlin / New York. Kepsu, Saulo. 1995. Pietari ennen Pietaria: Nevansuun kaupungin vaiheita ennen Pietarin perustamista. Helsinki: Suomalaisen Kirjallisuuden Seura. Kiparsky, Valentin. 1939. Die Kurenfrage. Annales Academiae Scientiarum Fennicae B42. Helsinki: Academia Scientiarum Fennica. Kiristaja, Arvis. 2013. Setomaa kohanimed. Seto Instituudi Toimetised 1. Värska: Seto Instituut. Kivimäe, Jüri. 2011. “Henricus the Ethnographer: Reflections on Ethnicity in the Chronicle of Livonia”. In Crusading and Chronicle Writing on the Medieval Baltic Frontier: A Companion to the Chronicle of Henry of Livonia. Ed. Tamm, Marek, Linda Kaljundi & Carsten Selch Jensen. Farnham: Ashgate. Pp. 77–106. Kiviniemi, Eero. 1977. Väärät vedet: Tutkimus mallien osuudesta nimenmuodostuksessa. Suomalaisen kirjallisuuden seuran toimituksia 337. Helsinki: Suomalaisen Kirjallisuuden Seura. Kjær, Albert, & Ludvig Holm-Olsen (eds.). 1910– 1986. Det Arnamagnæanske Håndskrift 81a Fol. (Skálholtsbók yngsta): Inneholdende Sverris saga, Bǫglunga sǫgur, Hákonar saga Hákonarsonar. Oslo. Klemming, G.E. 1865. Gamla eller Eriks-krönican. Svenska Medeltidens Rim-Krönikor 1. Stockholm: Norstedt & Söner. Koivulehto, Jorma. 1983. Seit wann leben die Urfinnen im Ostseeraum? – Zur relativen und absoluten Chronologie der alten idg. Lehnwort-schichten im 112 Matveev, A.K. 2001. Substratnaja toponimija Russkogo Severa I–II. Ekaterinburg: Izdatelstvo Uralskogo Universiteta. Mikkonen, Pirjo, & Sirkka Paikkala. 2000. Sukunimikirja. New edn. Helsinki: Otava. Mommsen, Theodorus (ed.). 1882. Iordanis Romana et Getica. Berolini: Weidmannos. Mullonen, I.I. 1994. Ocherki vepsskoj toponimii. Sankt-Peterburg: Nauka. RAN. Karelskij nauchnyj centr. IJALI. Mullonen, I.I. 2002. Toponimija Prisvir’ja: Problemy etnojazykovogo kontaktirovanija. Petrozavodsk: RAN, Karelskij nauchnyj centr. IJALI. Mundal, Else. 2000. “Coexistence of Saami and Norse Culture: Reflected in and Interpreted by Old Norse Myths”. In Old Norse Myths, Literature and Society: Proceedings of the 11th International Saga Conference. Ed. Geraldine Barnes & Margaret Clunies Ross. Sydney: Center for Medieval Studies, University of Sydney. Pp. 346–355. Myznikov, S.A. 2003. Atlas substratnoj i zaimstvovannoj leksiki govorov Russkogo Severo-Zapada. SanktPeterburg: Nauka. Napolskih, V.V. 2006. “Булгарская эпоха в истории финно-угорских народов Поволжья и Предуралья”. In История татар с древнейших времён в семи томах II: Волжская Булгария и Великая Степь. Казань. Pp. 100–115. Napolskikh, Vladimir. 2014. “Onoma in Iordanes’ “List of Hermanaric’s Peoples’”. Unpublished paper presented 30.5.2014 at the seminar of the Department of Finnish, Finno-Ugric and Scandinavian Languages, University of Helsinki. Nasonov, A.N. 1951. “Русская земля” и образование территории Древ- нерусского государства. Москва: Наука. Nenola, Aili. 2002. Inkerin itkuvirret – Ingrian Laments. Helsinki: Suomalaisen Kirjallisuuden Seura. Nirvi, Ruben Erik (ed.). 1971. Inkeroismurteiden sanakirja. Lexica Societatis Fenno-Ugricae 18. Helsinki: Suomalais-Ugrilainen Seura. Nordal, Sigurður (ed.). 1913–1916. Orkneyinga saga. København. North, Richard. 1997. Heathen Gods in Old English Literature. Cambridge: Cambridge University Press. Olrik, J., & H. Ræder (eds.). 1931. Saxonis Gesta Danorum. Hauniæ: Levin & Munksgaard. Ostrowski. Donald. 2003. The Povest’ vremennykh let: An Interlinear Collation and Paradosis I–III. Associate ed. David J. Birnbaum. Cambridge: Harvard Ukrainian Research Institute. Electronic edition: http://hudce7.harvard.edu/~ostrowski/pvl/ index.html. Ovsyannikov, O.V. 1980. “First-Discovered Burialfield of ‘Zavolochye Tschud’”. In Fenno-ugri et slavi 1978: Papers Presented by the Participants in the Soviet–Finnish Symposium “The Cultural Relations between the Peoples and Countries of the Baltic Area during the Iron Age and the Early Middle Ages” in Helsinki May 20–23, 1978. Helsinki: University of Helsinki. Department of Archaeology. Pp. 228–236. Kuzmin, Denis 2014a. Vienan Karjalan asutushistoria nimistön valossa. Helsinki: [Kuzmin]. Kuzmin, Denis. 2014b. “The Inhabitation of Karelia in the First Millennium AD in the Light of Linguistics”. In Ahola et al. 2014b: 269–295. Kylstra, A. D., Sirkka-Liisa Hahmo, Tette Hofstra & Osmo Nikkilä (eds.). 1991–2012. Lexikon der älteren germanischen Lehnwörter in den ostseefinnischen Sprachen I–III. Amsterdam: Radopi. Laakso, Johanna. 1999. ”Vielä kerran itämerensuomen vanhimmista kielenmuistomerkeistä”. Virittäjä 103: 532–555. Lehmann, Winfred P. 1986. A Gothic Etymological Dictionary. Leiden: Brill. Lehtosalo-Hilander, Pirkko-Liisa. 1984. “Keski- ja myöhäisrautakausi”. In Suomen historia I. Ed. Eero Laaksonen, Erkki Pärssinen & Kari J. Sillanpää. 2nd edn. Espoo: Weilin & Göös. Pp. 251–405. Leont’ev, A.E. 1996. Arkheologija Meri. K predystorij severno-vostochnoj Rusi: Arkheologija epoxy velikoj pereselenija narodov I rannego srednevekov’ja. Vyp. 4. Moskva. Lind, John. 2007. “Problems of Ethnicity in the Interpretation of Written Sources on Early Rus’”. In The Slavicization of the Russian North: Mechanisms and Chronology. Ed. Juhani Nuorluoto. Slavica Helsingiensia 27. Helsinki: University of Helsinki. Pp. 286–301. Lindkvist, Thomas. 2002. “Sverige och Finland under tidig medeltid”. In När kom svenskarna till Finland? Ed. Ann-Marie Ivars & Lena Huldén. Helsingfors: Svenska litteratursällskapet i Finland. Pp. 39–50. Line, Philip. 2007. Kingship and State Formation in Sweden 1130–1290. Leiden: Brill. Makarov, N.A. 1993. Russkij Sever: Tainstvennoe srednevekov’e. Moskva: Institut Arxeologii RAN. / Rossijskij fond fundamentalnyx issledovanii. Makarov N.A. 1997. Колонизация северных окраин древней Руси в XI-XIII вв.: По материлам археологических памятников на волоках Белозерья и Поонежья = Colonization of the Northern Periphery of Ancient Russia, 1000–1300 A.D.: In the Light of Archeological Data from the Portages of Beloozero and Onega River Regions. Москва: Скрипторий. Makarova, Anna Andreevna. 2012. Russkaja ozernaja gidronimija Belozer’ja: Sistemno-funkcionalnyj analiz. Dissertacija na soiskanie učenoj stepeni kandidata nauk. Uralskij federalnyj universitet imeni pervogo prezidenta Rossii B.N. El’cyna. Ekaterinburg. Markus, Felicia. 2004. Living on Another Shore: Early Scandinavian Settlement on the North-Western Estonian Coast. Occasional Papers in Archaeology 36. Uppsala: University of Uppsala. Mallory, J.P. 1989. In Search of the Indo-Europeans: Language, Archaeology and Myth. London: Thaems & Hudson. Matthews, W.K. 1951. Languages of the U.S.S.R. Cambridge: Cambridge University Press. Matveev, A.K. 1996. “Substratnaja toponimija Russkogo Severa i merjanskaja problema”. Voprosy jazykoznanija 1996(1): 3–24. 113 Saar, Evar. 2008. Võrumaa kohanimede analüüs enamlevinud nimeosade põhjal ja traditsionaalise kogukonna nimesüsteem. Dissertationes philologicae Universitatis Tartuensis 22. Tartu: Tartu Ülikool. Saarikivi, Janne. 2000. “Kontaktilähtöinen kielenmuutos, substraatti ja substraattinimistö”. Virittäjä 104: 393–415. Saarikivi, Janne. 2004. “Über die saamischen Substratennamen des Nordrusslands und Finnlands”. FinnischUgrische Forschungen 58: 162–234. Saarikivi, Janne. 2006. Substrata Uralica: Studies on Finno-Ugric Substratein Northern Russian Dialects. Tartu: Tartu University Press. Saarikivi, Janne. 2007. “Finnic Personal Names on Novgorod Birch Bark Documents”. In Topics on Ethnic, Linguistic and Cultural Making of North Russia. Ed. Juhani Nuorluoto. Slavica Helsingiensia 32. Helsinki: University of Helsinki. Pp. 196–246. Saarikivi, Janne. 2014. “Reconstruction of Culture and Ethnicity: Remarks on the Methodology of the Historical-Comparative Linguistics (on the Basis of Western Uralic Languages)”. In New Focus on Retrospective Methods. Ed. Eldar Heide & Karen Bek-Pedersen. FF Communications 307. Helsinki: Academia Scientiarum Fennicae. Saarikivi, Janne. Forthcoming (2015). “Finnic Borrowings in Permian”. In Uusiutuva etymologia. Ed. Sampsa Holopainen & Janne Saarikivi. Uralica Helsingiensia. Helsinki. Saarikivi, Janne, & Mika Lavento. 2012. “Linguistics and Archaeology: A Critical View of an Interdisciplinary Approach with Reference to the Prehistory of Northern Scandinavia”. In Networks, Interaction and Emerging Identities in Fennoscandia and Beyond: Papers from the Conference Held in Tromsø, Norway, October 13– 16 2009. Ed. C. Damm & J. Saarikivi. Helsinki: Suomalais-Ugrilainen Seura. Pp. 177–239. Salama, Elina. 2014. “The (Im)Possibilities of Genetics for Studies of Population History”. In Ahola et al. 2014b: 347–360. Salo, Unto. 2000. “Suomi ja Häme, Häme ja Satakunta”. In Hämeen käräjät I. Ed. Jukka Peltovirta. Hämeenlinna: Hämeen heimoliitto. Pp. 18–231. Schalin, Johan. 2008. “Ahuen maa ja Alandh”. Sananjalka 50: 24‒37. Schalin, Johan. 2014. “Scandinavian–Finnish Language Contact in the Viking Age in the Light of Borrowed Names”. In Ahola et al. 2014b: 399–436. Schalin, Johan, with Frog. 2014. “Toponymy and Seafaring: Indications and Implications of Navigation along the Åland Islands”. In Ahola et al. 2014c: 273–302. Schweizer-Sidler, H., & Eduard Schwyzer (eds.). 1902. Tacitus’ Germania. 6th edn. Halle: Waisenhaus. Sedov, V.V. 1994. Slavjane v drevnosti. RAN. Institut arheologii. Rossijskij fond fundamentalnyx issledovanii. Siikala, Anna-Leena. 2012. Itämerensuomalisten mytologia. Helsinki: Suomalaisen Kirjallisuuden Seura. Pajusalu, Karl. 1996. Multiple Linguistic Contacts in South Estonian Variation of Verb Inflection in Karksi. Turun Yliopiston Suomalaisen ja Yleisen Kielitieteen Laitoksen Julkaisuja 54. Turku: Turun Yliopisto Pajusalu, Karl, et al. 2002. Eesti murded ja kohanimed. Ed. Tiit Hennoste. Tallinn : Eesti Keele Sihtasutus. Palm, Rune 2004. Vikingarnas språk 750–1100. Stockholm: Norstedts. Partanen, Niko, & Janne Saarikivi. Forthcoming (2015). “Fragmentation of the Karelian Language and Its Community: Growing Variation at the Threshold of Language Shift. In New and Old Language Diversities. Ed. Janne Saarikivi & Reetta Toivanen. Clevedon: Multilingual Matters. Peel, Christine. 1999. Guta saga: The History of the Gotlanders. London: Viking Society for Northern Research. Peel, Christine. 2009. Guta lag: The Law of the Gotlanders. London: Viking Society for Northern Research. Pitkänen, Ritva Liisa. 1985. Turunmaan saariston suomalainen lainanimistö. Helsinki: Suomalainen Kirjallisuuden Seura. Power, Rosemary. 1985. “Journeys to the Otherworld in the Icelandic Fornaldarsögur”. Folklore 96(2): 156–175. Price, Neil. 2000. “Novgorod, Kiev and Their Satellites: The City-State Model and the Viking Age Polities of European Russia”. In A Comparative Study of Thirty City-State Cultures. Copenhagen: Royal Danish Academy. Pp. 263–275. Rahkonen, Pauli. 2011. “Finno-Ugrian Hydronyms of the River Volkhov and Luga Catchment Areas”. Journal de la Société Finno-Ougrienne 93: 205– 266. Rahkonen, Pauli. 2012. “Границы распространения меряно- муромских и древнемордовских гидронимов в верховьях Волги и бассейне реки Оки”. Voprosy Onomastiki 2012(1): 5–42. Rahkonen, Pauli. 2013. South-Eastern Contact Area of Finnic Languages in the Light of Onomastics. Jyväskylä: Bookwell. Räisänen, Alpo. 2003. Nimet mieltä kiehtovat: Etymologista nimistöntutkimusta. Helsinki: Suomalainen Kirjallisuuden Seura. Rajandi, Edgar. 2005, Raamat nimedest. Tammerraamat. Raninen, Sami, & Anna Wessman. 2014. “Finland as Part of the ‘Viking World’?”. In Ahola et al. 2014b: 327–346. Rendahl, Anne-Charlotte. 2001. “Swedish Dialects around the Baltic Sea”. In The Circum-Baltic Languages: Typology and Contact I. Ed. Östen Dahl & Maria Koptjevskaja-Tamm. Studies in Language Companion Series: 54. Amsterdam: John Benjamins. Pp. 137–177. Rjabinin, E.A. 1997. Финно-угорские племена в составе Древней Руси: К истории славянофинских этнокультурных связей. Историкоархеологические очерки. Санкт-Петербургский государственный университет – Институт истории материальной культуры РАН. СанктПетербург: Издательство Санкт- петербургского университета. 114 Archaeology 4. Tartu: Institute of History and Archaeology, University of Tartu. Uino, Pirjo. 1997. Ancient Karelia: Archaeological Studies – Muinais-Karjala: Arkeologisia tutkimuksia. Suomen Muinaismuistoyhdistyksen Aikakauskirja 104. Helsinki: Suomen Muinaismuistoyhdistys. Vahtola, Jouko. 1980. Tornionjoki- ja Kemijokilaakson asutuksen synty: Nimistötieteellinen ja historiallinen tutkimus. Studia Historica Septentrionalia 3. Rovaniemi: Pohjois-Suomen historiallinen yhdistys. Vahtola, Jouko. 1983. “En gammal germansk invandring till västra Finland i bynamnens belysning”. Historisk tidskrift i Finland 3: 253–279. Valtonen, Irmeli. 2008. The North in the Old English Orosius: A Geographical Narrative in Context. Mémoires de la Société Néophilologique de Helsinki 73. Helsinki: Société Néophilologique. Vasil’ev, V.L. 2005. Arhaičeskaja toponimija Novgorodskoj zemli: Drevneslavjanskie deantroponimnye obrazovanija. Novgorodskij mežregionalyj institut obščestvennyx nauk. Velikij Novgorod 2005. Vilkuna, Kustaa. 1957. Kainuu – Kvenland: Missä ja mikä? Helsinki: Suomalaisen Kirjallisuuden Seura. de Vries, Jan. 1956–1957. Altgermanische Religionsgeschichte I–II. 2nd edn. Berlin: de Gruyter. de Vries, Jan. 1961. Altnordisches etymologisches Wörterbuch. Leiden: Brill. Wessman, Anna. 2010. Death, Destruction and Commemoration: Tracing Ritual Activities in Finnish Late Iron Age Cemeteries (AD 550–1150). Iskos 18. Helsinki: Suomen Muinaismuistoyhdistys. Williams, Henrik. 2005. “Vittnar runstenen från Söderby (Gs 13) om Sveriges första ledungståg? – Runfilologi och konsten att läsa som det star”. Annales Societatis Litterarum Humaniorum Regiae Upsaliensis 2004: 39–53. Winkler, E., & K. Pajusalu. 2009. Salis-livisches Wörterbuch. Lingua Uralica, Supplementary Series 3. Tallinn: Estonian Academy Publishers. Zaliznjak, A.A. 2004. Drevnenovgorodskij dialect: Vtoroe izdanie: Pererabotannoe s učetom materiala naxodok 1995–2003 gg. Moskva: Jazyki Slavjanskoj Kultury. Simek, Rudolf. 1990. Altnordische Kosmographie: Studien und Quellen zu Weltbild und Weltbeschreibung in Norwegen und Island vom 12. bis zum 14. Jahrhundert. Berlin: de Gruyter. Sjöstrand, Per Olof. 2014. “History Gone Wrong: Interpretations of the Transition from the Viking Age to the Medieval Period in Åland”. In Ahola et al. 2014c: 83–152. Söderman, Tiina. 1996. Lexical Characteristics of the Estonian North Eastern Coastal Dialect. Studia Uralica Uppsaliensia 24. Uppsala: University of Uppsala. Storm Gustav (ed.). 1888. Islandske annaler indtil 1578. Christiania: Den Norske Historiske Kildeskriftsfond Sutrop, Urmas. 2003. “Taarapita: The Great God of the Oeselians”. Folklore (Tartu): 27–64. Sweet, Henry. 1883. King Alfred’s Orosius I: OldEnglish Text and Latin Original. London: Early English Text Society. Tadmor, Uri. 2009. “Loanwords and the World’s Languages: Findings and Results”. In Haspelmath & Tadmor 2009: 55–75. Talvio, Tuukka. 2014. “The Viking Age in Finland: Numismatic Aspects”. In Ahola et al. 2014b: 131– 138. Tarsala, Ilse. 1998. “Öar i strömmen: Den yngre järnåldern på Åland”. Aktuell Arkeologi 6: 107–123. Thomason, Sarah Grey, & Terrence Kaufman. 1988. Language Contact, Creolization and Genetic Linguistics. Berkeley: University of California Press. Tkachenko, O.V. 1985. Merjanskij jazyk. AN Ukrainskoj SSR. Kiev: Naukova dumka. Turkin A.I. “Топонимия нижней Вычегды. Канд. дисс”. АН СССР, Институт языкознания. Москва. Manuscript. Tvauri, Andres. 2008. “Migrants or Natives: The Research History of Long Barrows in Russia and Estonia in the 5th–10th Centuries”. In Topics on Ethnic, Linguistic and Cultural Making of the Russian North. Ed. Juhani Nuorluoto. Slavica Helsingiensia 32. Helsinki: University of Helsinki. Pp. 286–301. Tvauri, Andres. 2012. The Migration Period, PreViking Age, and Viking Age in Estonia. Estonian 115