Sei sulla pagina 1di 45

J Archaeol Res (2008) 16:37–81

DOI 10.1007/s10814-007-9017-8

ORIGINAL PAPER

The Archaeology of Regions: From Discrete Analytical


Toolkit to Ubiquitous Spatial Perspective

John Kantner

Published online: 26 September 2007


 Springer Science+Business Media, LLC 2007

Abstract In the 1970s and 1980s, regional analysis was an influential part of
archaeological research, providing a discrete set of geographical tools inspired by a
processual epistemological and interpretive perspective. With the advent of new
technologies, new methods, and new paradigms, archaeological research on regional
space has undergone significant changes. This article reviews the state of regional
archaeology, beginning with a consideration of its history and a discussion of the
fundamental issues facing regional investigations before focusing on developments
over the last several years. On one hand, the diversification of archaeological theory
has created new paradigms for thinking about human relationships with one another
and with the physical environment across regional space; in this regard, historical
ecology, landscape archaeology, and evolutionary theory have been particularly
influential in recent years. This has led to a corresponding diversification of the
traditional methods of regional analysis. Most notably, the advent of powerful
digital technologies has introduced new tools, especially those from the geographic
information sciences, that build on the quantitative methods of past approaches. The
investigation of regional data is no longer based on a discrete toolkit of simple
mathematical and graphical procedures for representing spatial relationships.
Instead, regional archaeology has matured into a diversity of multiscalar spatial and
geostatistical techniques that inform many areas of archaeological inquiry.

Keywords Regional analysis  Settlement patterns  Spatial analysis 


Archaeological method and theory  Geographic information systems

J. Kantner (&)
School for Advanced Research on the Human Experience, P.O. Box 2188, Santa Fe, NM 87504,
USA
e-mail: kantner@sarsf.org

123
38 J Archaeol Res (2008) 16:37–81

Introduction

What is the state of regional analysis in archaeology today? Although a review of


the current literature in archaeology reveals that scholarship can still be divided
between site-focused and region-focused studies, the line between the two is no
longer clearly demarcated. In fact, perhaps the majority of studies in archaeology
are more accurately described as ‘‘region-sensitive,’’ for considerations of human-
human and human-environment relationships across regional space are usually
important, even if the claim is not made that a ‘‘regional analysis’’ is being
conducted. One could argue that the region as a scale of archaeological inquiry is
rapidly becoming ubiquitous (e.g., Alcock and Cherry 2004, pp. 1–3). On the other
hand, the classic tools of regional analysis, such as fall-off analysis, central-place
modeling, and catchment analysis, are being replaced by more sophisticated
geostatistical and graphical techniques made possible with modern computing
technologies. The result is that regional archaeology no longer focuses on a distinct
suite of tools of archaeological inquiry. Space on various scales is part of almost any
form of archaeological research today, whether explicit or not, whether emerging
from social theory or processual archaeology, and whether the classic techniques of
regional analysis are formally applied or not.
This article first considers the history of regional analysis in archaeology before
reviewing issues that continue to challenge investigations at large spatial scales,
including problems in identifying what regions are, determining meaningful scales
and units for spatial analysis, and thereby defining what regional archaeology
includes. Accepting that a regional perspective is a valid way of thinking about and
studying the past, the article next reviews current methodological directions in
regional archaeology before considering important theoretical paradigms that are
influencing how the regional scale of analysis currently is manifested in the
discipline.

History of regional archaeology

Regional archaeology in the Americas is said to have originated with Julian


Steward’s research in the 1930s (Parsons 1972, p. 128). In his investigations of
societies in the Southwest and Basin areas of the United States, Steward (1937,
1938) advocated an ecological approach that explicitly considered the relationship
between environmental characteristics, human populations, and patterns of regional
settlement. His emphasis on entire landscapes rather than just individual sites
influenced a number of studies in the 1940s and 1950s. Gordon R. Willey’s ground-
breaking project in the Virú Valley of Peru (Willey 1949, 1999), for example,
combined aerial photography, architectural observations, and regional maps of site
distributions to reconstruct sociopolitical organization across an entire region and to
compare these patterns with the physical environment. No longer was archaeolog-
ical survey done exclusively to identify sites for excavation; now, the pattern of sites
across a region was itself an important source of information about past societies
(Anschuetz et al. 2001, pp. 168–170; Stanish 2001, p. 217).

123
J Archaeol Res (2008) 16:37–81 39

Although Willey’s analyses consisted of basic inferences derived from obser-


vations of the settlement data, his work shaped the future of settlement pattern
studies. In 1956, Willey published a collection of papers written by several
anthropologists interested in regional patterns of human behavior. One contributor,
William T. Sanders (1956), provided additional precision in the terminology used
for regional research, with a focus on defining the units of interaction at a regional
scale and on how these could be identified archaeologically. By the early 1960s,
Sanders was applying his ideas on a large scale in the Teotihuacan Valley, Mexico,
where he explored the symbiotic relationships that formed between agricultural
communities (Sanders 1965, 1999). Meanwhile, K. C. Chang’s work in the Arctic
considered ecological relationships between hunter-gatherers and the landscape
(Chang 1958).
Archaeology in many other parts of the world during the first several decades of
the 20th century was not explicitly interested in regional patterns of human
settlement (Galaty 2005, pp. 292–297; Parsons 1972, p. 136). Much of the reason
for the greater emphasis on regional archaeology in the United States was the
presence of expansive landscapes with comparatively little contemporary occupa-
tion, which facilitated large-scale surveys. In contrast, European archaeologists had
to aggregate the results of numerous small surveys and excavations to identify any
spatial patterning. The British landscape school derived from ‘‘topographic
approaches’’ that explored the rich complexity of local relief, elevation, soils, and
the very high density of archaeological remains, often on a very small spatial scale.
The classic study by Cyril Fox (1923) on the Cambridge region, however, was the
blueprint for a larger-scale exploration that examined the ‘‘personalities’’ of
different landscapes, from the higher, wetter, cooler, and less fertile ‘‘highland
zone’’ of west and north Britain to the lower, drier, warmer, and more fertile
‘‘lowland zone’’ of the east and south (Fox 1932). The use of aerial photography for
enhancing large-scale archaeological research was pioneered by British scholars
such as O. G. S. Crawford (1929).
The emergence of the ‘‘New Archaeology’’ in the mid-1960s profoundly
influenced the direction of regional archaeology (Anschuetz et al. 2001, pp. 170–
171). Its concern with scientific methodologies and new analytical tools, many
borrowed from other disciplines, had a strong impact on all kinds of formal
approaches. Kent Flannery’s ecosystem approach (Flannery 1968) emphasized the
interrelationships between human behavior and the surrounding environment,
thereby advocating a holistic regional perspective. An accompanying trend was
toward greater precision in the application of archaeological concepts. By the late
1960s, for example, several archaeologists were clearly distinguishing a ‘‘settlement
pattern’’ from a ‘‘settlement system,’’ the former reflecting empirical observations
made by the archaeologist and the latter representing the theory-bound interpre-
tations of the patterning (Flannery 1976; Parsons 1972, p. 132).
Another influence of the New Archaeology was an increasing use of quantitative
methods for investigating spatial data, many borrowed from geography, biology,
and economics, in an effort to make archaeology more scientific. Lewis and Sally
Binford’s (1966) reconstruction of Paleolithic settlement systems through the
application of multivariate statistics was one of the first such studies. Statistical

123
40 J Archaeol Res (2008) 16:37–81

techniques were commonly employed for identifying spatial patterns in the


distribution of artifacts, which could then be associated with particular social units
(e.g., Hill 1970; Whallon 1968). The New Archaeology also encouraged the
development of new techniques for producing detailed environmental reconstruc-
tions and comparing these with patterns of human settlement using methods such as
catchment analysis (e.g., Roper 1979; Viti-Finzi and Higgs 1970). The result was a
growing suite of discrete quantitative and graphical tools that comprise the classic
techniques that many scholars regard as ‘‘regional analysis.’’
Through the 1970s and into the 1980s, New Archaeology matured into processual
archaeology, which continued an interest in multidisciplinary perspectives on
human behavior and the use of formal quantitative methods to identify spatial
patterns. Although archaeologists were already employing techniques such as
central-place and gravity models (Hodder and Orton 1976), the 1976 publication of
Regional Analysis by Carol A. Smith further exposed the discipline to geographical
methods. Spatial location and allocation models, distance decay models, and
network analysis joined the suite of quantitative spatial tools in regional analysis
(Kantner 2005). Throughout these two decades, however, archaeology sometimes
struggled to apply these borrowed methods adequately enough to serve the
circumstances of its own discipline. Some of the more challenging issues included
how to sample and bound regions appropriately, given the vagaries of archaeolog-
ical data and no a priori knowledge of the spatial scale of the original sociocultural
landscape (e.g., Fish and Kowalewski 1990; Johnson 1980; Parsons 1972). A related
debate revolved around the proper units of analysis given the difficulty in
empirically defining entities such as ‘‘sites’’ or ‘‘communities’’ (e.g., Dunnell and
Dancey 1983; Lewarch and O’Brien 1981). Such issues were not as challenging to
the geographers and ecologists who contributed the tools of regional analysis—and
who could directly observe behavior—as they were to the archaeologists attempting
to apply them for reconstructing behaviors across past landscapes.
After the early 1990s, a number of influences in archaeology promoted new
approaches for reconstructing past regional behavior. Because the sometimes
uncritical application of techniques taken from other disciplines often was
challenged (e.g., Bell and Church 1985, p. 355; Paynter 1983; Rhoades 1978;
Ruggles and Church 1996, pp. 161–164), archaeologists became more cautious in
assessing new approaches for analyzing regional data. One result was the
development of unique approaches for examining complex regional data that built
on the strengths of classic analytic techniques but were designed specifically for the
challenges of archaeology (e.g., de Montmollin 1989; Gaffney et al. 1996; Lake
et al. 1998; Ruggles and Medyckyj-Scott 1996; Zubrow 1994). Theoretical
perspectives on regional space also were changing. Archaeologists interested in
‘‘historical ecology’’ explicitly considered the long-term trajectory of human-
environment interactions and their complex interplay in regional change (Crumley
1994a; Crumley and Marquardt 1990). From a somewhat different angle, the
postprocessual critique challenged the discipline’s focus on a strict positivist
epistemology and opened the way to qualitative techniques for evaluating past
landscapes. Influence from more humanistic geographers, with a lengthy intellectual
heritage tracing back to Carl Sauer (1925), also provided inspiration for this new

123
J Archaeol Res (2008) 16:37–81 41

direction (Anschuetz et al. 2001), especially in Europe, where ‘‘landscape


archaeology’’ continues to reflect an interest not just in spatial or ecological
relationships but also in how people perceive and construct the unique and often
ideologically charged regions in which they live (e.g., Kelso and Most 1990; Knapp
and Ashmore 1999).
The various influences that shaped regional archaeology in the late 1990s
occurred in the context of increasingly powerful computer technology for modeling,
analyzing, and visualizing complex multidimensional landscapes. The ease with
which sophisticated spatial analyses can be accomplished using geographic
information system (GIS) software not only makes the analysis of regional spaces
widely accessible but enables researchers to better apply and build on the traditional
tools of regional analysis (Aldenderfer and Maschner 1996; Allen et al. 1990;
Conolly and Lake 2006; Kvamme 1989; Maschner 1996). For example, trend-
surface analysis, introduced to archaeology in the 1970s for interpolating artifact
distributions to identify potentially meaningful patterns across regional space (e.g.,
Hodder and Orton 1976), can now be conducted quickly on a computer using a
variety of sophisticated geostatistical algorithms (Gillings and Wheatley 2005;
Johansen et al. 2004; Lloyd and Atkinson 2004). The power of computer hardware
and software also has reintroduced the possibility of simulation modeling across
digital ‘‘regions’’; archaeologists, after flirting with simulation in the 1970s (e.g.,
Aldenderfer 1981; Hodder 1979; Wobst 1974), had abandoned it in the face of
theoretical and technological challenges. The approach is now reentering the field at
the beginning of the 21st century, influenced especially by complexity theory,
agent-based modeling, and powerful new technologies. The result is that archae-
ologists today have at their disposal an extensive suite of tools for applying a
regional perspective to their research.

Fundamental issues in regional archaeology

A consideration of regional archaeology necessarily requires a working definition of


‘‘region.’’ In general, archaeological regions are spaces for which meaningful
relationships can be defined between past human behavior, the material signatures
people left behind, and/or the varied and dynamic physical and social contexts in
which human activity occurred. Unfortunately, archaeologists have too frequently
defined regions of interest only indirectly, often relying upon current constraints on
their research, such as modern political boundaries or the budget available for
fieldwork (e.g., chapters in Billman and Feinman 1999; Fish and Kowalewski 1990;
Parsons 2004, pp. 6–8). With so much archaeological investigation currently
conducted in the realm of cultural resource management (CRM), regional
approaches also are limited by immediate and practical spatial concerns such as
the areal extent of the proposed impact by modern development and land use.
When made explicit, regions are conceptualized at many different scales, from
continental regions incorporating enormous areas (e.g., Lekson 1999; McGuire
et al. 1994) to localized regions centered on small drainage systems (e.g., Madry
and Rakos 1996). Demarcation of regions is often determined by the specific

123
42 J Archaeol Res (2008) 16:37–81

questions and theoretical perspective guiding the research. In a review of what is


‘‘local’’ vs. ‘‘distant’’ in the greater U.S. Southwest, for example, Whalen and
Minnis (2003, p. 314) note that many scholars start with topographic boundaries but
also consider geographic patterning in material culture to bound analytical regions
(e.g., Duff 2000; Hegmon et al. 2000), often equating a spatially contiguous
distribution of distinctive artifacts and features with a sociopolitical or sociocultural
group. For many archaeologists, the region is conceptualized as equivalent to a
human ‘‘population,’’ which then requires that the material expression of such a
sociodemographic unit be identified (Parsons 2004).
While many archaeologists confine their regional studies to landscapes bounded
by prominent geographic features, this assumes that people in the past would have
had similar criteria for defining their landscapes. Similarly, the use of a model of
human settlement to identify a sociocultural or political region relies on the
assumption that the model correctly represents the criteria that actually shaped that
particular human landscape. Archaeologists are increasingly acknowledging that
past boundaries between people, as well as the relationships between humans and
the environment, were constantly changing (Ashmore 2002; Bender 2002; Dewar
1991; Dewar and McBride 1992; Lightfoot and Martinez 1995; Wandsnider 1992).
The challenges and assumptions that underlie the bounding of archaeological
regions are not insurmountable, but they do impact the kinds of questions and
analytical techniques that can be applied to the resulting regional data (Green and
Perlman 1985; Johnson 1977, p. 499; Neitzel 1994; Paynter 1983, pp. 253–260;
Wandsnider 1998). Accordingly, regional archaeology requires the researcher to
identify clearly how and why boundaries are determined.

Defining regional analysis and regional archaeology

What, then, is regional analysis? The term is borrowed from geography—its use in
archaeology has historically revolved around quantitative and graphical methods
taken from economic and cultural geography and to a lesser degree from spatial
modeling in ecology (Ebdon 1985; Hodder and Orton 1976; Smith 1976).
Archaeologists interested in region and landscape continue to benefit from the
methods of geography, which are evolving at a rate often unappreciated by
archaeologists, particularly new techniques for analyzing social and human-
environment relationships, not the least of which are developments in GIS
applications (e.g., Conolly and Lake 2006; Goodchild and Janelle 2004; Kvamme
1999). With the diversification of archaeological paradigms, however, regional
analysis has expanded in terms of applications and techniques, from a discrete suite
of tools borrowed from geography and ecology to diverse spatial analytic methods
available from a wide variety of disciplines as well as developed by archaeologists
themselves.
Modern regional analysis—or what we might call ‘‘regional archaeology’’ to
distinguish it from the traditional tools of regional analysis—is more than just a
limited set of tools for identifying and describing information from a large area. The
fundamental criterion is that regional archaeology is concerned with spatial

123
J Archaeol Res (2008) 16:37–81 43

relationships among human entities and between them and the nonhuman physical
world. Regional approaches accordingly contrast most obviously with the smaller
spatial scale of site-focused approaches, although of course the results of the latter
provide critical data for the former. Stanish (2001, p. 217) further notes, ‘‘a key
point is that, by controlling for context, the whole of the information collected from
a region is greater than the sum of the individual sites’’; ‘‘context’’ in this sense is
the spatiotemporal location of sites in a dynamic physical world. Coulam and
Schroedl (2004), for example, assess the meaning of Late Archaic split-twig figurine
styles in the U.S. Southwest from a regional perspective. They not only examine the
archaeology of individual sites to aid their interpretation, but especially consider
spatiotemporal context to conclude that figurine styles reflect ritual and domestic
functions, with the ritual style found in the ethnographically important Grand
Canyon. Coulam and Schroedl are not interested just in collecting and describing a
lot of data from a large area; they are interested in how and why the figurines are
distributed across that space in relation to one another and to the sociocultural and
physical environment.
Regional archaeology is in many ways synonymous with settlement pattern
analysis, to the point that the two approaches are often conflated (e.g., see
discussions in Billman and Feinman 1999; Fish and Kowalewski 1990). Regional
archaeology tends to be more interested in spatial relationships among a diversity of
human and environmental phenomena, whereas settlement pattern analysis tends to
concentrate more narrowly on quantifiable spatial relationships among material
remains. Regional archaeology, with its genesis in classic regional analysis and
processual archaeology, is more obviously informed by middle range theory—and
thus archaeological theory in general—compared to the data-generation focus of
settlement pattern archaeology. In practice, however, archaeologists use the terms
‘‘regional analysis,’’ ‘‘settlement pattern analysis,’’ ‘‘regional archaeology,’’ and
‘‘settlement archaeology’’ interchangeably, and archaeological theory is differen-
tially implicated in specific applications. The bottom line is that contemporary
regional archaeology is a widespread, method-oriented perspective for answering a
variety of anthropological problems through the use of spatiotemporal and
contextual data from a sizable, contiguous area.
One important caveat is that multiscalar approaches using regions as a point of
departure rather than the exclusive level of analysis are increasingly important. In
their zooarchaeological study of the northern San Juan region of the U.S. Southwest,
for example, Muir and Driver (2002) consider multiple scales of analysis, from the
household up to the regional level, to identify varying patterns of faunal use and
disposal. These differing spatial scales together provide a unique perspective on
questions of changing hunting strategies, community economies, regional aban-
donment, and sociopolitical organization. Similarly, supraregional scales of analysis
are enjoying a resurgence of interest, including those informed by world systems
theory (e.g., Stein 1999; but see Jennings 2006). In one such study, Smith and
Montiel (2001) consider economic exchange between a capital and distant
provincial areas to identify domination in hegemonic empires that do not otherwise
formalize their boundaries in the manner of territorial empires. A similar
supraregional analysis was conducted by Parker (2002), who attempts to identify

123
44 J Archaeol Res (2008) 16:37–81

and characterize the northern, Anatolian boundary of the Assyrian empire.


Exceptionally large scales of analysis—well beyond what might be considered a
‘‘region’’—also are employed for investigations of identity and migration. Geib
(2000) uses the supraregional distribution of Early Archaic sandal types to identify a
stylistic boundary dividing the northern and southern Colorado Plateau that
potentially reflects ethnic divisions (see also Wonderley 2005). Long-distance
migrations also are assessed at such supraregional scales (e.g., Kulischeck 2003;
Lekson and Cameron 1995; Lekson et al. 2002). Regional archaeology is clearly
just one scale in increasingly spatially sensitive and scale-inclusive archaeological
research (e.g., see chapters in Alcock and Cherry 2004).

Units of regional archaeology

A critical challenge for a regional approach is how to identify appropriate units of


analysis (e.g., Parsons 1972, pp. 137–144, 2004, pp. 9–12; Stafford and Hajic 1992,
pp. 138–143; Wandsnider 1998). Most techniques in regional archaeology require
the identification of comparable entities that can be quantified. Traditionally, the
unit of interest at this scale has been the ‘‘site,’’ an entity that is the subject of
considerable controversy; identifying boundaries of an archaeological site presents
problems similar to the definition of a region (Dunnell and Dancey 1983, p. 271;
Kooyman 2006). In some circumstances, the edges of a site seem quite obvious
because the archaeologist can easily distinguish the bounded area containing
material culture from surrounding space with no artifacts at all. A spatial clustering
of artifacts around the remains of a structure, for example, is often interpreted as
analytically distinct, perhaps representing a ‘‘household.’’ Even in these situations,
however, the investigator cannot be certain about how the bounded material culture
articulates with past human behavior; ethnographic cases, to continue the example,
illustrate how multiple domestic structures can often comprise a single sociocultural
‘‘household.’’ Of course, material culture recovered from any region inevitably
represents numerous episodes of human activity, and even seemingly well-bounded
material culture may represent a palimpsest of very different behaviors that occurred
over long periods of time (e.g., Dewar and McBride 1992, pp. 231–237).
Most regions exhibit a seemingly continuous distribution of material culture
deposited during many different episodes, and defining spatial units may seem more
an art than an empirical process. Archaeologists have developed many methods for
identifying unit boundaries, most of which can be classified as ‘‘distributional
approaches’’ that rely on the spatial patterning of the material remains. Nearest-
neighbor analysis, for example, is a venerable technique in which each item is
associated with its closest neighbor, producing clusters of material culture that can
be used to identify boundaries (e.g., Earle 1976; Washburn 1974); revisions to this
technique are enabling greater sophistication and flexibility (e.g., Bailey 1994, pp.
25–27; Hill 2004). ‘‘Fried-egg’’ techniques distill spatial distributions of material
culture into isolines and identify boundaries, with various smoothing and scaling
techniques used to identify patterns of interest and highlight site edges (e.g.,
Henderson and Ostler 2005; Kintigh 1988; Peterson and Drennan 2005). Clustering

123
J Archaeol Res (2008) 16:37–81 45

algorithms such as unconstrained cluster analysis similarly group artifacts or other


items together to identify patterning that is not readily visible to the eye and that can
be used to identify sites (Aldenderfer 1982; Whallon 1984).
Distributional approaches provide explicit methods for identifying boundaries,
but the underlying assumptions are often questioned (Cowgill 1989) Nearest-
neighbor analysis and most clustering algorithms will force material culture into
bounded spatial clusters without acknowledging that humans deposit items both in
and between meaningful points on the landscape. ‘‘Fried-egg’’ techniques are
extremely sensitive to the specific criteria used for defining isolines; just as
bounding a ‘‘hill’’ by the lines on a contour map is not always so obvious, defining
boundaries using isolines of material culture distributions is a decidedly subjective
task, even using methods such as progressive smoothing to identify clusters
(Peterson and Drennan 2005, pp. 16–17). Spatial statistics, however, can be helpful
for validating clusters of material culture and establishing them as meaningful units
for analysis (e.g., Bailey 1994; Kvamme 1996, pp. 46–49). Bevan and Conolly
(2005), for example, advocate the use of Ripley’s K function as a statistical
procedure for identifying the aggregation and segregation of point data at different
spatial scales; they illustrate the approach on the Aegean island of Kythera. In his
study of Holocene settlement patterns in Jordan, Hill (2004) supplements nearest-
neighbor statistics with Hodder and Okell’s A statistic and local density analysis,
complementary approaches that measure segregation and association between
points. Despite these successes, Bevan and Conolly (2005) do note that the
assumptions of most geostatistical procedures are problematic for typical archae-
ological distributions.
Problems associated with defining discrete units of analysis have promoted a
fully distributional approach known as ‘‘siteless’’ archaeology in which the goal is
to analyze and interpret artifact distributions directly, instead of the creation and
analysis of interpretive units such as sites (Ebert 1992; Ebert et al. 1996; Galaty
2005, pp. 300–302). Proponents contend that any attempts to distinguish sites from
nonsite areas are misguided; they argue that units of space rather than arbitrary
clusters of artifacts should provide the basic analytical elements for regional
archaeology (Dunnell 1992; Lewarch and O’Brien 1981, p. 322). They contend that
the notion of the ‘‘site’’ is confounded by both the methodological issues described
above and, more importantly, theoretical inadequacies. Because human behavior is
the subject of interest, and because all material culture represents this behavior,
proponents of siteless archaeology believe that investigators should be more
interested in the distribution of the actual cultural material across a region rather
than in arbitrary behavioral units that disguise variability or impose interpretation
(Dunnell 1992, pp. 32–33; Dunnell and Dancey 1983, pp. 271–274). Methods used
in siteless archaeology include surface analyses that display spatial patterning of
artifacts or other items of interest. Increasingly accessible computer technologies
such as GIS can accommodate the large volumes of data that siteless analyses at
regional scales require (e.g., Ebert et al. 1996; Given 2004; Thompson 2004).
Most archaeologists appreciate the motivations of siteless archaeology (e.g.,
Parsons 2004), but the approach has not been widely applied for two reasons (Ebert
1992; Lewarch and O’Brien 1981, pp. 322–325). First, archaeologists can make

123
46 J Archaeol Res (2008) 16:37–81

interpretations of past behavior only by building associations between groups of


artifacts and other entities in the archaeological record; most individual items are
meaningless outside of these associations. Accordingly, archaeologists use these
associations as units of behavioral analysis, and entities such as activity areas and
sites help represent these associations. Although siteless approaches make this
process explicit and encourage the recognition of off-site distributions, the
additional effort may seem excessive when the results are comparable to traditional
site-focused archaeology. Second, our discipline’s traditions have permeated
modern procedures, and, for better or for worse, the ‘‘site’’ has become the
archaeological entity recognized by government agencies, preservation law, and
CRM procedures. Continuous distributions of material culture are much more
difficult to work with for making management or policy decisions about the
recording and preservation of cultural heritage.
Although siteless/distributional archaeology has seen little application in
practice, legitimate concerns linger about whether the site is an appropriate unit
for understanding human behavior, on a regional scale or otherwise (Dunnell 1992;
Peterson and Drennan 2005). What does the site represent? Is it a ‘‘household,’’ a
‘‘community,’’ or an ‘‘activity area’’? And the reverse also is a significant problem:
If we want to identify households and communities, what should they look like on
the regional landscape (e.g., Henderson and Ostler 2005; Kolb and Snead 1997;
Mahoney 2000; Mahoney et al. 2000)? While many researchers advocate spatial
analyses at varying scales for differentiating and interpreting various sociocultural
units and for reconstructing the physical landscapes of which they are a part (e.g.,
Bevan and Conolly 2005; Fisher and Feinman 2005, pp. 65–66; Holdaway and
Fanning 2004), ultimately the meaning given to spatial units is an interpretive
problem. Wells and his colleagues (2004), for example, note that nearly three-
quarters of the sites recorded in surveys of the middle Gila River landscape cannot
be assigned behavioral meaning any more precise than ‘‘artifact scatters.’’ They
therefore employ density and diversity measures to categorize these between-village
scatters into interpretable activity areas needed to reconstruct meaningful regional
histories.
The definition of appropriate spatial units of analysis is clearly an important
issue, but of equal concern is the placement of units on a temporal scale (Coffey
2006; Wandsnider 2004). This, of course, is not just a problem of regional
archaeology, but insofar as all human landscapes are a palimpsest of millennia of
behavior, analysis from a regional perspective cannot proceed without accurate
methods to identify chronological change. In the absence of absolute dates, the
traditional approach is to use cross-dated stylistic complexes of diagnostic material
culture to divide archaeological units into different time periods. Contemporary
methods for developing precise regional chronologies include microseriation, which
improves traditional frequency seriation by focusing on attributes rather than styles
(e.g., Duff 1996). Correspondence analysis, a multivariate scaling procedure, is
particularly popular for creating chronological seriations for regional archaeology
(e.g., Baxter 1994; Steponaitis and Kintigh 1993). For example, in their analysis of
regional demographic changes in the Zuni region of the U.S. Southwest, Kintigh and
his colleagues (2004, pp. 436–439) compare the use of traditional stylistic

123
J Archaeol Res (2008) 16:37–81 47

complexes with correspondence analysis. They conclude that while procedures such
as correspondence analysis provide some level of statistical rigor, they are not
necessarily superior to traditional approaches. The authors accordingly advocate the
use of both statistical seriation and traditional stylistic complexes for refining
regional chronologies. For defining regions, the units of analysis within them, and
their chronological relationships, multiple analytical approaches like that advocated
by Kintigh and his colleagues provide the most secure results.

New methods in regional archaeology

Arguably the most notable changes in regional archaeology are the advances in
analytical methods. The classic tools of regional analysis, such as fall-off analysis,
catchment studies, location-allocation modeling, and rank-size distributions, are
used less often today (Kantner 1996, 2005), although some important exceptions do
exist (e.g., Brown and Witschey 2003; Drennan and Peterson 2004; Hare, 2004; Lee
2004). In contrast, the phenomenal growth of inexpensive and accessible computing
power combined with the methodological maturity resulting from several decades of
regional archaeology have made a variety of new techniques available that in many
cases have replaced or substantially built on the analytical tools of the past. Parallel
to this is the growing utility of spatially sensitive analytical approaches emerging
from allied fields such as geophysics and geochemistry, techniques that have
experienced remarkable technological advances and increasing accessibility due to
lower costs and greater ease of use. While the older methods still have a place in
regional archaeology, in practice their use has declined as—for better or for
worse—the attention of archaeologists has turned to new technologies for
conducting large-scale spatial and geostatistical analyses.

Survey techniques

Regional archaeology cannot proceed without the application of successful survey


techniques. Most archaeologists continue to advocate and practice full-coverage
survey as opposed to sampling approaches, particularly as the importance of
complex forms of regional interaction is demonstrated again and again in diverse
archaeological cases (e.g., Bauer and Covey 2002; Covey 2003; Feinman and
Nicholas 1999; Finsten and Kowalewski 1999; Fish and Kowalewski 1990; Spencer
and Redmond 2001; Whalen and Minnis 2003). While full-coverage survey is
costly, several decades of practice (especially with the ubiquity of pedestrian survey
in CRM) have refined survey methods (Banning et al. 2006; Collins and Molyneaux
2003). Archaeologists continue to develop new techniques and approaches as they
confront new challenges and integrate new technologies to reconstruct past
settlement patterns and physical landscapes. Cannon (2000), for example, demon-
strates the utility of a sophisticated soil probe that extracts subsurface stratigraphy
intact and thereby characterizes deposits quickly. Use of this tool to test shell-
midden sites on the central coast of British Columbia made it possible for Cannon

123
48 J Archaeol Res (2008) 16:37–81

and his colleagues to record sites twice as fast as traditional testing techniques,
allowing them to rapidly reconstruct regional settlement history.
Increasing accessibility to remote-sensing technologies enables archaeologists to
augment their survey data at varying spatial scales. Aerial photography and space-
based remote sensing have been in use for some time (e.g., Eddy et al. 1996; Madry
and Crumley 1990; Obenauf 1980), and their utility is greater now than ever before,
with a wide variety of multispectral data products available at varying levels of
precision from both government and private industry sources (Fowler 2002; Harmon
et al. 2006; Lock 2003; Sever and Irwin 2003; Wilkinson et al. 2004). Ground-
based geophysical survey has received the most attention in the last several years,
with many universities and CRM firms purchasing the increasingly less-expensive
resistivity, conductivity, magnetometry, gradiometry, and ground-penetrating radar
equipment (Hargrave et al. 2002; Kvamme 2003; Silliman et al. 2000). While
geophysical surveys are not yet directly possible at the regional scale, the techniques
speed the acquisition of the data needed for regional assessments (e.g., Conyers
et al. 2002).
Other new approaches for identifying archaeological remains on a regional scale
concentrate on past landscapes that are inaccessible through standard procedures.
Pleistocene occupations that are currently below sea level due to Holocene warming
and coastal flooding are of great interest to archaeologists but notoriously difficult to
identify. In his investigations of the Florida coast, Faught (2004) uses a combination
of bathymetric enhancement, subbottom profiling, and side-scan sonar remote
sensing to find the courses of paleo-river channel, thereby identifying likely
locations for Paleoindian remains that can be targeted during diver surveys. Lewis
(2000) explores Holocene-era sea-level rise from another angle, examining state-
maintained site records in Mississippi to identify coastal regions impacted by
eustatic sea-level rise. In this case, archaeological remains older than 2500 BP are
proportionally underrepresented along the coast compared to nearby lands that were
protected from the effects of inundation, leading to a biased pattern of current site
distribution. Water is not the only deterrent to identifying regional patterns; glaciers
and ice patches hide evidence important for understanding human cultural behaviors
at both high latitudes and high altitudes, as evidenced by the frequency of remains
currently being uncovered as a result of global warming. Dixon and his colleagues
(2005) present a GIS-based modeling technique for identifying locations where
archaeological remains are likely to be revealed as glaciers and ice patches melt.
The model combines a variety of biological, geological, and cultural datasets with
satellite imagery to evaluate archaeological potential and guide aerial and pedestrian
regional survey.

Geographic information science

The attraction of GIS is its ability to organize spatially referenced data of varying
types into a single database (Conolly and Lake 2006; Kvamme 1999; Lock 2003;
Wheatley and Gillings 2002). Because the archaeological record and environmental
characteristics are represented by a wide variety of point, linear, polygonal, and

123
J Archaeol Res (2008) 16:37–81 49

continuous spatial data, the data-management strengths of GIS have largely fueled
its growth in archaeology. Hill and his colleagues (2004), for example, have
assembled an enormous GIS database of the entire prehistoric U.S. Southwest that
includes over 3,000 sites with nearly 6,000 temporal components. They are using it
to track trends in depopulation, migration, and the resulting coalescence in
multiethnic towns. The true power of GIS, however, is its analytical capabilities
(Church et al. 2000), which go far beyond the simple comparison of different
‘‘layers’’ of data that marked earlier applications. Today, the variety of techniques
for capturing, storing, displaying, and analyzing geospatial data is known as
‘‘geographic information science’’ (GISci) and includes digital technologies from
remote sensing to total station mapping to viewshed and watershed analyses. While,
historically, archaeologists have been several steps behind other disciplines in
appreciating and harnessing GISci capabilities, this is rapidly changing as GIS and
GISci become integrated into graduate education and applied training in
archaeology.
Some GISci applications in regional archaeology are building on the classic
techniques of regional analysis, such as network analysis and distance decay
models, which can be done much more quickly with tremendous amounts of data
using GIS software (Conolly and Lake 2006). In one study, Johansen and his
colleagues (2004) use network analysis to investigate the spatial patterning of Early
Bronze Age barrow mounds in southern Jutland to identify social interaction and the
corresponding flow of goods. Intentional alignments of mounds, or what the
researchers call ‘‘barrow lines,’’ are analyzed for centrality using network theory.
Correspondence analysis of material culture found along these lines compared with
items recovered away from the lines suggests to Johansen and his colleagues that a
network of social interaction and the exchange of wealth goods occurred in the
region. The longevity of the regional patterning points to a cultural tradition
surrounding mound building that superseded the lives of individuals in southern
Scandinavia, while the lack of central places suggests an absence of chiefly societies
until later in the Bronze Age. Such complex spatiotemporal network and
correspondence analyses are possible with computer-based GISci technology.
Today, many if not most GISci applications in archaeology are using GIS to
assemble geospatially referenced environmental data, create paleoenvironmental
reconstructions, and compare these with the regional distribution of past human
settlement. These analyses are increasingly easy to do with the growing availability
of digital environmental data, and the results can be used to determine the impact of
environmental structure on regional human settlement, the effect of climate change
on settlement patterns, and the role of human resource use in altering the physical
landscape. Field (2004), for example, uses a GIS database to assemble regional
information on the topography and soil quality of Fiji’s Sigatoka Valley. Additional
data on seasonal and long-term climatic fluctuations are then employed to
reconstruct the cultivation risks in different parts of the region due to both drought
and floods. Field then compares these results with the settlement history to illustrate
how these spatiotemporal environmental challenges encouraged the development of
competitive strategies as represented by fortified habitations. Another compelling
example is provided by Hill (2000, 2004), who uses GISci techniques to evaluate

123
50 J Archaeol Res (2008) 16:37–81

the impact of Holocene human settlement on soil erosion in west-central Jordan.


Part of his assessment is based on a GIS database built from topographic,
precipitation, soil, and vegetation spatial information, which was input into a
universal soil loss equation to create potential soil loss raster data for the entire
region. By comparing the results of the environmental analysis with reconstructions
of the intensity of past landscape use, Hill demonstrates that soil degradation has
been a problem for the region’s agripastoral inhabitants throughout the Holocene.
Importantly, both Field and Hill note the applied value of their studies for
understanding environmental risks for today’s inhabitants of Fiji and Jordan,
respectively.
GIS functionality includes the ability to model high-resolution digital landscapes,
which can then be used to simulate the effect of topography on movement and
visibility (Conolly and Lake 2006). Most GIS software includes algorithms for
generating ‘‘cost surfaces,’’ or ‘‘friction surfaces,’’ which calculate the costs of
movement across three-dimensional space and that can identify likely pathways used
by people in the past. Digitally modeling complex landscapes also allows for much
more realistic and accurate assessments of site catchments (e.g., Varien 1999; Varien
et al. 2000). In his study of human-induced soil degradation in Jordan described
earlier, Hill (2004) establishes use territories for each site in the region according to
the cost of movement across the topographically variable landscape. The GIS then
generates a ‘‘use-intensity’’ surface by accumulating the use territories of each site,
which in turn is compared with the independently generated measures of soil loss.
Jennings and Craig (2001) employ a similar approach but a different cost-surface
algorithm to generate movement costs away from the centroids of Andean valleys
reportedly controlled by the Wari empire. This GIS-based approach is used to model
areas of spatial control in relation to administrative sites (see also Hare 2004).
Algorithms for determining the intervisibility of two points in a region as well as
the complete ‘‘viewshed’’ of a single point also are built into most GIS software
packages (Conolly and Lake 2006). In his research on Paquimé (Casas Grandes) of
northern Mexico, Swanson (2003) analyzes digital elevation data in a GIS database
to identify the intervisibility of hilltop features hypothesized to be fire-signaling
stations. While the results reveal that the features can all be linked together into a
line-of-sight network, Swanson acknowledges that the topography of the region
naturally promotes intervisibility, even if that was not the intent of the Paquimé
people. He therefore created a GIS routine to generate intervisibility estimates for
randomly selected topographic high points, using the resulting data for statistical
support.
Combining the geospatial functions of GIS software can be particularly
informative for understanding past behavior at regional scales. Kantner and
Hobgood (2003), for example, employ both cost-surface and viewshed analysis to
evaluate the cultural landscape of a region of the northern U.S. Southwest. In their
study, remnants of 11th-century roadways associated with the Chacoan tradition are
compared with a simulated network of GIS-generated ‘‘cost paths’’ that connect
Puebloan villages together. The results suggest that the roads were not part of a
regional ‘‘network’’ linking disparate villages into a unified economic system; rather
they indicate that roadways served local functions related to small-scale

123
J Archaeol Res (2008) 16:37–81 51

socioideological dynamics. Next, an alleged communication network based on


Chaco-era towers found in Puebloan villages is evaluated by assessing GIS-
computed intervisibility. Again, the analysis casts doubt on the proposed
intervillage connectivity and instead identifies the enhanced localized viewshed
provided by the towers. Similar studies include Mack’s (2004) use of both viewshed
analysis and cost-path modeling to reconstruct the dramatic landscape experiences
of pilgrims visiting 16th-century Vijayanagara, India, while Byerly and colleagues
(2005) apply the two techniques to assess the function of a purported bison jump in
Texas through an understanding of its place on the Paleoindian landscape.
GISci applications in archaeological inquiry have grown in popularity largely due
to their ease of use and the availability of so many useful, prepackaged analytical
algorithms. The seduction of this new technology, however, has too often led to its
uncritical use. Many different software packages exist, and each has different and
often invisible built-in algorithms designed to achieve similar functions. Archaeol-
ogists also develop their own formulas for their GIS analyses, as illustrated by the
numerous cost-path algorithms reported in the literature (e.g., Kantner 2004). Little
empirical work has been conducted to compare the strengths and weaknesses of
different packages and algorithms, leading to uncertainty as to how much value
should be placed in GIS-based studies. Similarly, the precision allowed in GIS
databases often becomes confused with accuracy, such that fuzzy geospatial data at
low resolutions attain enhanced value once integrated into sophisticated GIS-based
regional analyses in which variability in data quality cannot easily be accommodated.

Predictive modeling and simulation

The application of predictive modeling in regional archaeology is enjoying a


resurgence of interest. Once popular in the 1970s and 1980s (Aldenderfer 1981),
particularly for management purposes on public lands, predictive modeling and
related approaches in computer simulation were criticized for their emphasis on the
role of the physical environment in determining patterns of human settlement, a
debate that still continues today (e.g., Gaffney and van Leusen 1995; Kvamme
1997). Related to this was the difficulty of accessing computers powerful enough—
and the requisite training in programming—to do anything more than very simple
models. Today, with supercomputers on everyone’s desks, and readily available,
object-oriented modeling software, archaeologists are again exploring the utility of
modeling and simulation for both practical applications and for evaluating theory
(Brantingam 2006; Ebert 2000; Westcott and Brandon 2000). Unfortunately, as in
the case of GISci approaches, such ease of access often comes with the uncritical
use of the new technology (see Woodman and Woodward 2002).
Most computer modeling in archaeology takes advantage of the built-in
analytical capabilities of GIS software. With regional environmental information
organized into data layers in a GIS database, simple mathematical procedures can be
used to relate variables together according to a theoretical model and produce
predictions of where archaeological sites should be found (e.g., Ebert and Singer
2004; Elliott 2005). Or the reverse can be done, using known site locations to

123
52 J Archaeol Res (2008) 16:37–81

identify the environmental correlates of past human settlement (e.g., Duncan and
Beckman 2000; Perkins 2000; Warren and Asch 2000), which in turn can be used to
guide targeted regional survey. Addressing criticisms of environmental determin-
ism, many current applications consider not only physical environmental conditions
but also social variables. Stancic and Kvamme (1999) take this approach in their
GIS-based predictive modeling of Bronze Age hillforts in Croatia. In addition to the
standard suite of physical variables such as slope and soil information gleaned from
space-based remote-sensing data, the study includes four ‘‘social’’ variables. While
the latter are primarily variables related to distances between archaeological
features, begging the question of exactly what a ‘‘social variable’’ is, the resulting
predictive model reflects a growing interest in a wider variety of causative variables
impacting regional settlement. In another example, the relative contributions of men
and women to subsistence underlies Zeanah’s (2004) GIS-based computer
simulation of settlement changes in the Carson Desert of western Nevada. While
explicitly social variables are not built into the GIS database, Zeanah does use
models from human behavioral ecology to separately simulate the foraging
strategies of men and women, which are then combined together to predict
subsistence changes and thus the settlement patterns of the Late Holocene.
Several archaeologists are building simulation models in which actual ‘‘agents’’
interact with one another and the physical landscape over numerous ‘‘generations’’
according to a set of theoretically derived rules. Such agent-based modeling in
archaeology is almost always spatial, considering social interactions and environ-
mental conditions at a regional scale. In one recent case study, the landscape of
Long House Valley of northeastern Arizona was digitally recreated and agents
‘‘released’’ onto it (Axtell et al. 2002; Dean et al. 2000; Gumerman et al. 2003).
The simulation considered agents to be equivalent to households, and each was
given basic rules that determined how much food the agent consumed, when the
agent would have to move to find more food, and the points at which a household
would fission and when it would ‘‘die.’’ The simulation was allowed to run for the
equivalent of 500 years, and the spatial and demographic results were compared
with the actual archaeological patterns. Johnson and colleagues (2005; see also
Kohler et al. 2000) conduct a similar agent-based simulation for the Mesa Verde
region of southwest Colorado, but they frame their study as an historical ecological
investigation of human impacts on the landscape. Like the previous example, their
simulation considers households to be digital agents that follow simple rules over
many simulated years. The focus of that study, however, is on fuelwood demands by
the agents and the subsequent effects on the physical environment. Johnson and his
colleagues model different wood-use rates to identify their impact on both forest
depletion and regional settlement patterns, concluding that human-induced degra-
dation had serious repercussions for sustainable occupation of the landscape.

Demographic reconstruction

Much of our understanding of the past relies on reconstructions of how many people
were alive at a specific location at a particular moment in time. Methods for

123
J Archaeol Res (2008) 16:37–81 53

accurate demographic reconstructions are therefore of particular concern to


archaeologists. Recognizing that demography can be scaled along multiple spatial
and temporal dimensions, not all demographic reconstructions are for region-sized
populations. However, the obverse is true: scholars investigating regional socio-
cultural phenomena want to know how many people were alive and interacting at
varying sociopolitical scales during the time period of interest (e.g., Bandy 2004;
Cobb and Butler 2002; Kintigh et al. 2004, p. 432; Kowalewski 2003; Milner and
Oliver 1999; Nelson and Hegmon 2001; Osborne 2004). Accordingly, much recent
work in regional archaeology is considering new approaches for developing precise
and accurate demographic reconstructions.
Most archaeologists continue to rely directly or indirectly on ethnographic
analogies for building models of regional population. In his study of the Pre-
Pottery Neolithic in the Near East, Kuijt (2000) needs an accurate demographic
reconstruction to understand relationships between sedentism, agriculture, and
sociopolitical inequality. He considers ethnographic studies comparing overall
settlement area and roofed area with numbers of residents when attempting to
reconstruct regional and community-level populations. Acknowledging the
inadequacies of these approaches, Kuijt nevertheless proposes that they are still
useful for comparative purposes (Kuijt 2000, pp. 80–85). Working with a more
complete archaeological record, Kintigh and colleagues (2004) similarly recon-
struct settlement demography to understand late 13th-century aggregations at large
towns in the Zuni region of the U.S. Southwest. In their study, the researchers
consider simple demographic estimates, such as those derived by calculating the
average number of rooms occupied per year of a given temporal period—a
measure similar to that used by Kuijt. Kintigh and his colleagues, however, desire
a more precise measure of annual population size within a temporal period. They
present a modeling technique in which individual room use lives are tabulated
annually and adjusted according to the growth rate calculated across the entire
period.
Gallivan (2002) takes a different approach in his investigations of sedentism
and demography in late precontact and early contact contexts in the Middle
Atlantic region. Measures of pottery discard are used to determine duration and
size of occupancy—a method known as ‘‘accumulations research’’ (Varien and
Mills 1997). To provide a standard rate of discard, Gallivan uses ‘‘strong’’
archaeological cases where intact floor features can be used to convincingly infer
a standard household ceramic assemblage. Assuming that the ‘‘strong cases’’ are
valid, this method for reconstructing regional demography avoids the question-
able uniformitarian assumptions of direct ethnographic analogy while providing
what is perhaps the most precise approach. Demographic reconstruction for
Scarre’s (2001) recent research in Neolithic Brittany, in contrast, relies on
indirect measures of population size while advocating the use of multiple
measures from different sources to overcome the limitations of any one method.
In that study, Scarre assesses the distribution of polished stone axes, palyno-
logical records of deforestation and cereal cultivation, and labor estimates for
monument construction to provide a general assessment of regional demography
and mobility.

123
54 J Archaeol Res (2008) 16:37–81

Complementary techniques for regional archaeology

Many analytical methods used in archaeology generate data that enhance


interpretation when considered from a regional perspective, even though they are
not usually considered part of traditional regional analysis in archaeology. These
techniques allow us to look at the regional distribution of archaeological remains in
a new light by providing the means to associate materials with one another or with
the physical environment in a way that complements simple spatiotemporal
relationships. Many such archaeological techniques useful for regional archaeology
exist, but for the sake of space, only two are considered here: compositional and
osteological studies.

Compositional studies

The goal of compositional analysis is to identify the elemental and/or mineral


components of material items. When the resulting data are considered from a
regional spatial perspective, compositional analysis becomes very effective at
reconstructing how people created, modified, or moved items within a particular
landscape. Accordingly, in most recent applications in archaeology, compositional
analysis has concentrated on the production and distribution of ceramics within or
between regions. A recent investigation by Eerkens and colleagues (2002) provides
a typical example. They use instrumental neutron activation analysis (INAA) to
compare the elemental composition of ceramics recovered from a regional
archaeological record with the distribution of raw clays across the landscape to
show that pottery among the Numa of the southwestern Great Basin was produced at
the household level and distributed and consumed locally. Nichols and colleagues
(2002) similarly utilize INAA to reconstruct changing spatial relations of production
and distribution in the Basin of Mexico from ceramics, identifying the existence of a
highly localized solar market model and increasing regionalism through time. INAA
is also used in an analysis of pottery from capacocha sacrifices in Inca territory
(Bray et al. 2005), showing that the pottery consistently came from both clay
sources at Cuzco and Lake Titicaca and local areas where the sacrifices occurred,
even those located over 1,000 km apart.
Many other materials can be analyzed using INAA, and many other methods for
compositional analysis also exist (e.g., Kantner et al. 2000; Kennett et al. 2002). The
mobility patterns of Paleoarchaic foragers across the Great Basin are reconstructed
using energy dispersive X-ray fluorescence (ED-XRF) analyses of lithic artifacts
(Jones et al. 2003). In this example, locations of artifact recovery were compared with
the availability of raw stone on the landscape to help establish territorial boundaries of
different populations. Another study uses new, nondestructive portable infrared
mineral analyzer spectroscopic technology (PIMA-SPTM) to source red, flint-clay
stone figurines associated with Cahokia (Emerson et al. 2003). The assessment of
contexts of recovery combined with regional compositional data indicate that
production of these figurines occurred in Cahokia in the 12th century and that as
Cahokia deteriorated in the late 13th century, the figurines were traded away.

123
J Archaeol Res (2008) 16:37–81 55

Osteological studies

Bone isotope analysis is a well-established technique for reconstructing past diet


from skeletal remains using the proportions of stable isotopes identified with a mass
spectrometer. Applications of this method usually are used to track changing diet
over time rather than across space. Some recent studies, however, have looked at
dietary differences across a region or among different regions. Tomczak (2003), for
example, considers dietary variation, as reconstructed through carbon and nitrogen
isotopic analyses of human skeletal remains, among Chirabaya sites in southern
Peru. The pattern of variation is used to assess both horizontal and vertical models
of resource utilization, leading to support of the former model and the identification
of regional economic specialization (see also Borrero and Barberena 2006). Yesner
and colleagues (2003) attempt to confirm that the ethnohistoric record of Terra del
Fuego in South America can be extended into the past. Carbon and nitrogen isotope
analyses of skeletons from several sites largely confirm that ethnohistorically
documented ethnic groups with different subsistence economies also existed in the
past. Although the incorporation of regional relationships into bone isotope studies
presents exciting possibilities, both of the examples cited here are limited by the
sizes of their regional skeletal populations. As the cost of geochemical assays
decreases and the geospatial precision of the analytical approaches increases, the
value of these types of data for regional archaeology will grow.
Craniometric analyses also benefit from regional scales of assessment. Blom
(2005) invokes Fredrik Barth’s work (2000) on ethnic group boundaries and
ethnohistorical accounts from Spanish explorers in her osteological assessment of
cranial modification in the Tiwanaku empire of the southern Andes. Her study of
skeletal populations from three valleys, including the one containing Tiwanaku
itself, reveals clear boundaries between all the valleys. Remains recovered from the
capital, however, exhibit a mix of deformation styles, reflecting its central authority
and regional draw. Blom’s study illustrates the necessity of a regional perspective
for assessing questions of group identity, particularly ethnicity. A use for
osteological research that has seen considerable interest over the past several years
is that of long-distance migrations. Pinhasi and Pluciennik (2004), for example,
build on previous settlement pattern analyses of demographic processes of Neolithic
expansion in the Old World, specifically considering interregional craniometric data
as a means for tracing the region-by-region movement of agriculturalists from the
Levant into southeastern Europe. These and other new ways of looking at human
movement across regional and supraregional space are demonstrating great potential
for adding another technique to the toolkit of regional archaeology.

New paradigms for regional archaeology

In the last two decades, several new theoretical approaches in anthropology have
affected how regional archaeology is practiced, either because they use regional
data in remarkably new ways or because they conceptualize human spatial behavior
on a regional scale unlike ever before. These new paradigms are in various ways

123
56 J Archaeol Res (2008) 16:37–81

inspired by the traditional processual approach to regional data, either tracing their
epistemological and interpretive history directly to processual archaeology, or
reflecting the more humanistic postmodern approach that is at least in part a reaction
to the perceived problems with processual archaeology. This section summarizes
three of these new paradigms—historical ecology, landscape archaeology, and
evolutionary theory—and evaluates their impacts on regional analysis as a discrete
set of tools for archaeological inquiry and on the regional perspective more
generally. These certainly are not the only archaeological approaches benefiting
from regional scales of inquiry, but a broad spatial viewpoint is central to each of
these perspectives.

Historical ecology

Scholars long have recognized that human impact on the physical environment is
not only a recent phenomenon resulting from the Agricultural and Industrial
Revolutions (Redman 1999). From the so-called ‘‘blitzkrieg’’ model of megafauna
extinction (e.g., Martin et al. 1985) to deforestation for firewood (e.g., Kohler 1992;
Kohler and Matthews 1988), humans often are implicated for negatively impacting
the landscape, sometimes to such a degree that they unintentionally precipitate their
own economic and sociopolitical crises. Recognition of this systemic human-
environmental relationship has been formalized into an area of multidisciplinary
research that is known variously as ‘‘historical ecology,’’ ‘‘landscape history,’’
‘‘environmental history,’’ or ‘‘socioecology,’’ among others (e.g., Barton et al. 2004;
Butzer 1982; Crumley 1994a; Crumley and Marquardt 1990; Hardesty and Fowler
2001; Kim 2003; Winthrop 2001).
Historical ecology traces how human impact has created a particular landscape
and how that resulting landscape has in turn shaped human behavior. Crumley
(1994b, p. 6) provides the example of a forest:
[A]stronomically driven regional climate is modified by latitude and
topography, and by the nonuniform distribution of population and human
activity; thus the existence of a forest is the result of both location, which
determines temperature and rainfall patterns, and previous and current human
management practices.
The reconstruction of the dialectical relationship between humans and the regions in
which they live provides two kinds of information. On one hand, historical ecology
most obviously tells us how specific landscapes came to be the way they are today
and the role of human populations in creating the world in which we live. This
ostensibly can provide lessons for the future (e.g., Fisher and Feinman 2005;
Redman 1999). On the other hand, historical ecology also can tell us about human
values, attitudes, and behaviors at specific points in the past as read from their
impact on the landscape. This provides us reconstructions of the past from a
regional environmental perspective. Historical ecology is explicitly multidisciplin-
ary, drawing data from geographers and historians as well as ethnographers,
ethnohistorians, and archaeologists, and it is analytically multiscalar, considering

123
J Archaeol Res (2008) 16:37–81 57

human-environment interaction at varying spatial and temporal scales, with a


particular emphasis on the regional scale (e.g., Runnels 2000).
Recent examples of studies in historical ecology include Fisher’s (2005)
multidisciplinary landscape project that is based on settlement pattern research and
intensive geoarchaeological investigations. Fisher shows that in the history of the
Tarascan empire of Mesoamerica, episodes of land degradation were the unintended
consequences of human actions and not caused by nonhuman environmental
changes. Barton and colleagues (2004) provide another recent example, although
they represent it as socioecology rather than historical ecology. They investigate
changes in human-environment landscapes in a region of eastern Spain from the
Late Pleistocene through the Middle Holocene, employing a program of survey and
subsurface testing of geographic study units that they call ‘‘patches.’’ Measures of
land use are generated with the assistance of GIS software for four valleys, and a
reconstruction of ‘‘contingent landscapes’’ is presented in which evolutionary
changes in human cultural behavior and the environment are shaped by earlier
human-environment interactions.
Some studies that arguably fall under the rubric of historical ecology are less
interested in the recursive nature of human-environment interactions and more
focused on the ‘‘coevolution’’ of human populations and the physical environment.
Darling and colleagues (2004), for example, consider the case of the Akimel
O’odham, Gila River Pima Indians of south-central Arizona, and the ‘‘village drift’’
of their rancherias in response to high-frequency environmental changes, such as
shifts in river channels. While in this example people did not so much precipitate
environmental change, previous decisions about where to live and plant fields
shaped their possible responses to environmental changes, producing a predictable
form of drifting settlement within culturally defined territories (also see Waters and
Ravesloot 2001). Despite the primacy given to the physical environment, this case
study still embraces the role of prior decisions in contingently shaping how humans
behave on the landscape, the essential paradigmatic premise of historical ecology.

Landscape archaeology

At first glance, landscape archaeology and historical ecology appear very similar, as
both see humans as intrinsic components of landscapes, and in fact the labels are
often used interchangeably. However, historical ecology and allied approaches
generally trace their intellectual heritage to processual archaeology, and they
typically concentrate on functional-economic relationships between humans and the
regional landscapes in which they live. Some historical ecological reconstructions
are akin to parables that warn us of our impact on the environment (e.g., Redman
1999). Landscape archaeology, on the other hand, is more closely informed by
postmodern currents in anthropology as a whole, and particularly by social theory,
regarding landscape as an ideational construct of the human mind. While both
historical ecology and landscape archaeology regard human-environment interac-
tion as their topic of inquiry, the former emphasizes the interaction of humans in or
as part of the landscape while the latter emphasizes landscapes as creations of

123
58 J Archaeol Res (2008) 16:37–81

human cognition. As Bender (2002, p. S103) notes in her essay on landscape and
time, ‘‘landscape is time materializing: landscapes, like time, never stand
still…[And] landscapes and time can never be ‘out there’: they are always
subjective.’’
Knapp and Ashmore (1999) provide an oft-cited formulation of landscape
archaeology. They note that landscape exists only insofar as it is ‘‘perceived,
experienced, and contextualized by people’’ (1999, p. 1), paralleling Bender’s
phenomenological approach to landscape (also see Anschuetz et al. 2001, pp. 160–
161; Ashmore 2002). Seminal work in the phenomenology of landscape has
especially deep roots in British archaeology (Thomas 1996; Tilley 1994). From this
perspective, landscape is regarded as comprising places that no matter how physical
or essential to basic human needs, are only meaningfully constituted through human
action in reference to them. Mobile foragers, for example, move through a regional
space of trails, views, hunting points, water holes, and so on, together which
constitute a cultural landscape, since people have assigned these places culturally
situated meanings mediated by their collective experiences with them over time.
Landscape archaeology further contends that landscapes reinforce cultural values, in
accordance with the perspective’s foundation on social theories of practice and
structure. Accordingly, human response to nonhuman environmental change is
discursively shaped by the created landscape in which the change occurs. Because
not all elements of a landscape have the same meaning and creation, Knapp and
Ashmore (1999, pp. 10–12) categorize landscape in three ways: constructed
landscapes, in which culturally meaningful features are built onto the landscape;
conceptualized landscapes, in which cultural meanings are attributed to natural
features, with few constructed features; and ideological landscapes, which are emic,
imagined landscapes rife with meaning that evoke emotional responses.
The majority of archaeological research conducted under the rubric of landscape
archaeology focuses on themes of place and meaning within regional space. ‘‘Place’’
is a fundamental concept within landscape archaeology that encapsulates the central
theoretical perspective of this paradigm: places are temporal human creations within
landscapes, ‘‘the hybrid conjoining of heterogeneous semantic fields—imaginar-
ies—with the material world [at varying scales]’’ (Whitridge 2004, p. 243; also see
Alcock 2002; Thomas 2001). Examples include Potter’s study (2004) of hunting
landscapes in the U.S. Southwest, in which he uses the regional distribution and
imagery of rock art places to identify a gender-charged landscape cognized by its
human creators as a field in which men establish their maleness. In their research on
ritual landscapes associated with the 19th-century ghost dance movement in western
North America, Carroll and colleagues (2004) approach place and meaning from a
different direction. They examine the integration of regional geography and ritual
technologies into a single imaginary of landscape places, which then can be used to
identify ceremonial settings in the archaeological record.
An important arena of research in landscape archaeology is the role of cultural
memories of places in defining identities. From this perspective, places created by
people at a past point in time are continually engaged in the negotiation of meaning
at a later point in time; time and space, in fact, are envisioned differently than from
a traditional Western cultural viewpoint. In some case studies, places are charged

123
J Archaeol Res (2008) 16:37–81 59

with past events and personalities that are integral to current practices. The research
of Stewart and colleagues (2004) on Inuit oral histories and archaeological places in
northern Canada, for example, identifies how places imagined as ‘‘traditional’’ are
‘‘foregrounded’’ in the landscape. Similarly, Colwell-Chanthaphonh’s ethnohisto-
rical work (Colwell-Chanthaphonh 2003; Ferguson and Colwell-Chanthaphonh
2006) in the San Pedro Valley of southeastern Arizona notes the evocative role of
regionally distributed rock-art places for shaping the identities of several indigenous
groups today (see also King 2003). Historical archaeologists especially have
gravitated to this approach to past regional landscapes, with a particular interest in
how the meanings of places can be contested by different sociocultural groups with
different memories of the landscape (e.g., contributions in Shackel 2001; Shackel
and Chambers 2004).
Monuments represent another area of interest in landscape analysis, especially
since a ‘‘monument’’ is necessarily a construct of human perception of space and
place, meaning and memory. Johansen (2004), for example, reveals the monumental
nature of the Neolithic ashmounds of South India. As identified by their spatial
locations and stratigraphic details, these unusual constructions of decomposed and
burned cow dung, soils, and material culture are interpreted as not just enormous
middens but instead as products of episodic ritual practice that visually reinforced
sociosymbolic meanings; only analysis at a regional scale could identify this
pattern. In south-central Ohio, Bernardini’s (2004) examination of Hopewell
earthworks distinguishes the ‘‘referential meaning’’ of individual monuments from
their ‘‘experiential meaning,’’ suggesting that the earthworks were part of a single
regional landscape maintained through ceremonial action rather than the separate
creations of autonomous villages (see also Howey and O’Shea 2006). Massive
ditches found in southern Benin, West Africa also take on new meaning when
examined through the lens of landscape archaeology (Norman and Kelly 2004).
From dating from the 17th through 19th centuries, these features historically were
interpreted by Europeans as the product of Western military designs. Norman and
Kelley (2004), however, use both anthropological and archaeological data to argue
that the Hueda and Dahomey kingdoms created a built landscape in which the
ditches were symbolic barriers patrolled by supernatural members of their
pantheons. In all these examples, human understanding of and interactions with
and across regional space are shaped by how the landscapes are perceived, not only
by ‘‘objective’’ assessments of the physical environment and cost-benefit analyses of
behavioral options.

Evolutionary theory

While landscape archaeology engages social theory in considerations of human


behavior and culture on a regional scale, evolutionary theory also offers a
theoretical foundation for the regional investigation of past behavior. Advocates of
evolutionary theory in archaeology all trace their intellectual histories to Darwin’s
theory of evolution by natural selection—and, to a lesser degree, sexual selection—
as updated with a contemporary understanding of genetics and complementary

123
60 J Archaeol Res (2008) 16:37–81

evolutionary processes. This ‘‘neo-evolutionary’’ or ‘‘neo-Darwinian’’ perspective


has long been a part of regional archaeology, tracing back to the adaptationist views
inherent in cultural ecology. Most often the application of evolutionary theory is
implicit and imprecise. Nevertheless, because neo-evolutionary theory purports to
explain changing relationships among changing organisms in changing landscapes,
it remains an important means for understanding human cultural behavior at
regional scales. Contemporary evolutionary theory in archaeology occurs in three
forms: evolutionary archaeology, or what is often called ‘‘selectionism’’; cultural
transmission theory; and human behavioral ecology. Each differs significantly in
how neo-evolutionary theory is interpreted and in their understanding of how this
can be applied in archaeological research, at a regional scale or otherwise.
Selectionism most strictly interprets evolutionary theory while also pointing to
the empirical insufficiency of reconstructed past behavior. Accordingly, this
approach focuses on the direct analysis and interpretation of material culture,
regarding it as an extension of the human phenotype and therefore shaped by natural
selection. Due to this emphasis on artifacts rather than behavior and the selectionist
interest in ecological spatiotemporal settings, regional data are often collected
through distributional archaeology. Ladefoged and Graves (2000), for example,
consider agricultural strategies in the north Kohala region on the island of Hawai’i
using selectionist theory. They argue that a heterogeneous and changing environ-
ment provided the context in which selection acted on the material manifestations of
differing variants of agricultural practice. Using a GIS database of 570 km of
agricultural field border walls and 190 km of trails across a 55-km2 area, the
researchers argue that selection favored a shift from field locations and designs that
optimized total energy return to those that provide more stable, risk-averse returns.
In his investigation of steatite vessel spatiotemporal distribution in the Eastern
Woodlands of North America, Truncer (2004) uses a selectionist perspective in a
supraregional analysis. Steatite vessels were produced in low frequencies for
2,000 years, but Truncer argues that they markedly increased around 2,500 cal B.C.
because selective conditions favored their functional characteristics. His regional
perspective is reflected in the spatiotemporal correlation of vessel distributions with
mast-forest environmental contexts.
The second evolutionary approach, cultural transmission theory, concentrates on
the mechanisms by which cultural traits are passed from individual to individual,
contending that those mechanisms are quite different from how biological traits are
transmitted and are as important, if not more so, than the role of natural selection in
shaping human cultural change. With its emphasis on the role of interpersonal
communication, a consideration of space is an important part of cultural
transmission theory, and regional scales of analysis are common. Jordan and
Shennan (2003) provide a compelling example. In their study, the researchers were
interested in regional distributions of basketry styles and how these reflect cultural
transmission among indigenous Californian groups. The analysis reveals that
linguistic boundaries are not mirrored by stylistic patterning. The basketry traditions
also do not correlate with ecological patterns, suggesting that changes in basketry
style were not transmitted using the same mechanisms as language and therefore did
not accumulate trait changes in the same way. Jordan and Shennan demonstrate the

123
J Archaeol Res (2008) 16:37–81 61

value of a regional perspective for understanding changing human cultural traits


across both time and space. The scholars note, however, that teasing apart the
differential processes of basketry and linguistic trait transmission will require them
to move their analysis to the subregional scale.
In contrast to both selectionism and cultural transmission theory, human
behavioral ecology is more directly interested in human decision making, arguing—
often implicitly—either that evolutionary processes have shaped cognition to
operate as a proxy for selective forces or that these forces operate directly on human
behavior. Either way, adaptive behavior and the material culture it produces are
assessed at regional scales of ecological analysis. In another evolutionary study on
Hawai’i, this one in the Kona region, Allen (2004) employs a bet-hedging model
from behavioral ecology to show shifting farming strategies in the context of
environmental unpredictability. Her conclusions and regional data sources parallel
Ladefoged and Graves’s (2000) study, but Allen emphasizes chiefly decision-
making strategies rather than the success of farming systems in the face of selective
pressure. Many other applications of behavioral ecology similarly compare
predictions of human foraging behavior with regional archaeological records
(e.g., Bettinger 1999; Elston and Zeanah 2002).
Despite the strength of evolutionary approaches in regional archaeology, they
cannot readily explain all human behavior. Whitridge’s (2001) multiscalar analysis
of fish remains and fishing technologies among the Thule Inuit warns archaeologists
that regional distributions of resources do not always equate with the subsistence
economy. Whitridge uses the distribution of fish remains at varying scales to
conclude that fish were a minor component of classic Thule diet. That pattern
contrasts, however, with the importance of fishing in material and oral culture,
leading Whitridge to propose that fishing was used for monitoring environmental
conditions but perhaps more importantly for communal recreation. This regional
study serves as a cautionary tale illustrating that humans use environmental
resources in complex ways that have a large social component that evolutionary
approaches often miss.

Concluding comments

For many archaeologists, regional archaeology is equated with quantitative and


graphical approaches for characterizing the spatial relationships among archaeo-
logical remains and between them and the physical environment. During its
heyday in the 1970s and 1980s, regional analysis included a suite of tools, such as
fall-off analysis, central-place models, and catchment analysis, situated within an
ecological and processual epistemological and interpretive paradigm. This review
of regional approaches over the last several years suggests that the regional
analysis of the past no longer maintains such a cohesive form. Analytical tools that
abstract and/or idealize human-space relationships are not as common. Important
exceptions do exist (e.g., Blanton 2004; Brown and Witschey 2003; Drennan and
Peterson 2004; Hare 2004; Nelson and Schollmeyer 2003), but even these cases
are much more sophisticated than their progenitors. Complementing the old

123
62 J Archaeol Res (2008) 16:37–81

techniques are new tools such as GIS-based landscape modeling and agent-
oriented computer simulation that arguably provide more contingent and realistic
reconstructions of human interaction with and across regional space. Regional
analysis has become a more ubiquitous and analytically flexible regional
archaeology.
What happened to the simple models of classic regional analysis? No single
event in the last two decades can be identified as the point at which the old tools
were replaced by the newer approaches described in this article, and in fact
traditional regional analysis still does provide heuristic value. But at least two
converging factors likely led to the erosion of regional analysis as a discrete toolkit
of simple quantitative and graphical techniques. First, the growing availability of
powerful computers and their ease of use have promoted model complexity over
analytical simplicity. Now, trend surfaces stretched to accommodate least-cost-
path movement and representing three-dimensional space are more common than
fall-off analysis; complex spatiotemporal models of farming productivity with
multiyear storage simulations provide more realism than catchment analysis;
fractal geometry and complexity theory build upon the rank-size rule; and massive
relational databases now accommodate artifacts and features as basic analytical
units instead of sites and communities. These technologically sophisticated tools
provide a level of precision that make the old tools of regional analysis less
attractive. On the other hand, the new techniques do have their drawbacks: their
complexity can confound comparisons between projects, and often the adoption of
technological applications seems to be done just for the sake of being able to use
the hottest new ‘‘toy.’’
The second factor that likely eroded the unity of regional analysis is the variety of
new theoretical paradigms interested in human-space interactions, many of which
are intellectually incompatible or even explicitly opposed to one another. Historical
ecology, landscape archaeology, selectionist evolutionary theory, and human
behavioral ecology, to name a few, regard space in different ways (Daly and Lock
2004), not just in the interpretive sense of what they think humans do in the space,
but more fundamentally in the epistemological sense of what and how we can know
of this arena of human behavior. Traditional adaptationist and functionalist
approaches couched in the scientific method vie with social theory situated in
hermeneutic or dialectical epistemologies as the interpreters of past human spatial
behavior. Nevertheless, even as this has led to more diverse ‘‘regional archaeol-
ogies’’ with different thematic interests, they often share many of the newer tools,
such as GIS, that allow archaeologists to identify and describe relationships across
regional space.
Is regional analysis therefore dead? Perhaps. Consider how often ‘‘regional
analysis’’ appears as the keyword of an article today compared with 25 years ago. In
its place, however, is a dominant regional perspective in which an assessment of
multiscalar spatial context is fundamental to a great many archaeological studies, no
matter what specific analytical tools or theoretical paradigms are employed. Julian
Steward would probably appreciate how thoroughly assessment of human-space
relationships is integrated into archaeological inquiry today.

123
J Archaeol Res (2008) 16:37–81 63

References cited

Alcock, S. (2002). Archaeologies of the Greek Past: Landscape, Monuments, and Memories, Cambridge
University Press, Cambridge, UK.
Alcock, S., and Cherry, J. F. (eds.) (2004). Side by Side Survey: Comparative Regional Studies in the
Mediterranean World, Oxbow Books, Oxford.
Aldenderfer, M. S. (1981). Computer simulation in archaeology: An introductory essay. In Sabloff, J.
(ed.), Simulations in Archaeology, University of New Mexico Press, Albuquerque, pp. 11–49.
Aldenderfer, M. S. (1982). Methods of cluster validation in archaeology. World Archaeology 14: 61–72.
Aldenderfer, M. S., and Maschner, H. D. G. (eds.) (1996). Anthropology, Space, and Geographic
Information Systems, Oxford University Press, Oxford.
Allen, K. M. S., Green, S. W., and Zubrow, E. B. W. (1990). Interpreting Space: GIS and Archaeology,
Taylor and Francis, London.
Allen, M. S. (2004). Bet-hedging strategies, agricultural change, and unpredictable environments:
Historical development of dryland agriculture in Kona, Hawaii. Journal of Anthropological
Archaeology 23: 196–224.
Anschuetz, K. F., Wilshusen, R. H., and Scheick, C. L. (2001). An archaeology of landscapes:
Perspectives and directions. Journal of Archaeological Research 9: 157–211.
Ashmore, W. (2002). ‘‘Decisions and dispositions’’: Socializing spatial archaeology. American
Anthropologist 104: 1172–1183.
Axtell, R. L., Epstein, J. M., Dean, J. S., Gumerman, G. J., Swedlund, A. C., Harburger, J., Chakravarty,
S., Hammond, R., Parker, J., and Parker, M. (2002). Population growth and collapse in a multiagent
model of the Kayenta Anasazi in Long House Valley. Proceedings of the National Academy of
Sciences 99: 7275–7279.
Bailey, T. C. (1994). A review of statistical spatial analysis in geographical information systems. In
Fotheringham, S., and Rogerson, P. (eds.), Spatial Analysis and GIS, Taylor and Francis, Bristol,
PA, pp. 13–44.
Bandy, M. S. (2004). Fissioning, scalar stress, and social evolution in early village societies. American
Anthropologist 106: 322–333.
Banning, E. B., Hawkins, A. L., and Stewart, S. T. (2006). Detection functions for archaeological survey.
American Antiquity 71: 723–742.
Barth, F. (2000). Boundaries and connections. In Cohen, A. P. (ed.), Signifying Identities: Anthropo-
logical Perspectives on Boundaries and Contested Values, Routledge, New York, pp. 17–36.
Barton, C. M., Bernabeu, J., Aura, J. E., Garcia, O., Schmich, S., and Molina, L. (2004). Long-term
socioecology and contingent landscapes. Journal of Archaeological Method and Theory 11:
253–295.
Bauer, B. S., and Covey, R. A. (2002). Processes of state formation in the Inca heartland (Cuzco, Peru).
American Anthropologist 104: 846–864.
Baxter, M. J. (1994). Exploratory Multivariate Analysis in Archaeology, Edinburgh University Press,
Edinburgh.
Bell, T. L., and Church, R. L. (1985). Location-allocation modeling in archaeological settlement pattern
research: Some preliminary applications. World Archaeology 16: 354–371.
Bender, B. (2002). Time and landscape. Current Anthropology 43: S103–S112.
Bernardini, W. (2004). Hopewell geometric earthworks: A case study in the referential and experiential
meaning of monuments. Journal of Anthropological Archaeology 23: 331–356.
Bettinger, R. L. (1999). From traveler to processor: Regional trajectories of hunter-gatherer sedentism in
the Inyo-Mono region, California. In Billman, B. R., and Feinman, G. M. (eds.), Settlement Pattern
Studies in the Americas: Fifty Years Since Virú, Smithsonian Institution Press, Washington, DC, pp.
39–55.
Bevan, A., and Conolly, J. (2005). Multiscalar approaches to settlement pattern analysis. In Lock, G., and
Molyneaux, B. (eds.), Confronting Scale in Archaeology: Issues of Theory and Practice, Kluwer,
New York, pp. 217–234.
Billman, B. R., and Feinman, G. M. (eds.) (1999). Settlement Pattern Studies in the Americas: Fifty Years
Since Virú, Smithsonian Institution Press, Washington, DC.
Binford, L. R., and Binford, S. R. (1966). A preliminary analysis of functional variability in the
Mousterian of Levallois facies. American Anthropologist 69: 238–295.

123
64 J Archaeol Res (2008) 16:37–81

Blanton, R. E. (2004). Settlement pattern and population change in Mesoamerican, Mediterranean


civilizations: A comparative perspective. In Alcock, S. E., and Cherry, J. F. (eds.), Side-by-Side
Survey: Comparative Regional Studies in the Mediterranean World, Oxbow Books, Oxford, pp.
206–232.
Blom, D. E. (2005). Embodying borders: Human body modification and diversity in Tiwanaku society.
Journal of Anthropological Archaeology 24: 1–24.
Borrero, L. A., and Barberena, R. (2006). Hunter-gatherer home ranges and marine resources: An
archaeological case from southern Patagonia. Current Anthropology 47: 855–867.
Brantingham, P. J. (2006). Measuring forager mobility. Current Anthropology 47: 435–459.
Bray, T. L., Minc, L. D., Ceruti, M. C., Chavez, J. A., Perea, R., and Reinhard, J. (2005). A compositional
analysis of pottery vessels associated with the Inca ritual of capacocha. Journal of Anthropological
Archaeology 24: 82–100.
Brown, C. T., and Witschey, W. R. T. (2003). The fractal geometry of ancient Maya settlement. Journal
of Archaeological Science 30: 1619–1632.
Butzer, K. W. (1982). Archaeology as Human Ecology: Method and Theory for a Contextual Approach,
Cambridge University Press, Cambridge, UK.
Byerly, R. M., Cooper, J. R., Meltzer, D. J., Hill, M. E., and LaBelle, J. M. (2005). On Bonfire Shelter
(Texas) as a Paleoindian bison jump: An assessment using GIS and zooarchaeology. American
Antiquity 70: 595–629.
Cannon, A. (2000). Settlement and sea-levels on the central coast of British Columbia: Evidence from
shell midden cores. American Antiquity 65: 67–77.
Carroll, A. K., Zedeño, M. N., and Stoffle, R. W. (2004). Landscapes of the Ghost Dance: A cartography
of Numic ritual. Journal of Archaeological Method and Theory 11: 127–156.
Chang, K.-C. (1958). Study of the Neolithic social grouping: Examples from the New World. American
Anthropologist 60: 298–334.
Church, T., Brandon, R. J., and Burgett, G. R. (2000). GIS applications in archaeology: Method in search
of theory. In Westcott, K. L., and Brandon, R. J. (eds.), Practical Applications of GIS for
Archaeologists: A Predictive Modeling Kit, Taylor and Francis, Philadelphia, pp. 135–155.
Cobb, C. R., and Butler, B. M. (2002). The Vacant Quarter revisited: Late Mississippian abandonments of
the lower Ohio Valley. American Antiquity 67: 625–641.
Coffey, G. (2006). Reevaluating regional migration in the northern San Juan during the Late Pueblo I
period: A reconnaissance survey of the east Dove Creek area. Kiva 72: 57–72.
Collins, J. M., and Molyneaux, B. L. (2003). Archaeological Survey, AltaMira, Walnut Creek, CA.
Colwell-Chanthaphonh, C. (2003). Native American perspectives of the past in the San Pedro Valley of
southeastern Arizona. Kiva 69: 5–29.
Conolly, J., and Lake, M. (2006). Geographical Information Systems in Archaeology, Cambridge
University Press, Cambridge, UK.
Conyers, L. B., Ernenwein, E. G., and Bedal, L.-A. (2002). Ground penetrating radar (GPR) mapping as a
method for planning excavation strategies, Petra, Jordan. E-tiquity 1. Journal online. Available from
http://www.saa.org/publications/E-tiquity/e-tiquity.html. Accessed 28 July 2006.
Coulam, N. J., and Schroedl, A. R. (2004). Late Archaic totemism in the Greater American Southwest.
American Antiquity 69: 41–62.
Covey, R. A. (2003). A processual study of Inka state formation. Journal of Anthropological Archaeology
22: 333–357.
Cowgill, G. L. (1989). Formal approaches in archaeology. In Lamberg-Karlovsky, C. C. (ed.),
Archaeological Thought in America, Cambridge University Press, Cambridge, UK, pp. 74–88.
Crawford, O. G. S. (1929). Air Photography for Archaeologists, Ordnance Survey, Professional Series
No. 12, Southampton, UK.
Crumley, C. L. (ed.) (1994a). Historical Ecology: Cultural Knowledge and Changing Landscapes, School
of American Research Press, Santa Fe, NM.
Crumley, C. L. (1994b). Historical ecology: A multidimensional ecological orientation. In Crumley, C. L.
(ed.), Historical Ecology: Cultural Knowledge and Changing Landscapes, School of American
Research Press, Santa Fe, NM, pp. 1–13.
Crumley, C. L., and Marquardt, W. H. (1990). Landscape: A unifying concept in regional analysis. In
Allen, K. M. S., Green, S. W., and Zubrow, E. B. W. (eds.), Interpreting Space: GIS and
Archaeology, Taylor and Francis, London, pp. 73–79.

123
J Archaeol Res (2008) 16:37–81 65

Daly, P., and Lock, G. (2004). Time, space, and archaeological landscapes: Establishing connections in
the first millennium BC. In Goodchild, M. F., and Janelle, D. G. (eds.), Spatially Integrated Social
Science, Oxford University Press, Oxford, pp. 349–365.
Darling, J. A., Ravesloot, J. C., and Waters, M. R. (2004). Village drift and riverine settlement: Modeling
Akimel O’odham land use. American Anthropologist 106: 282–295.
de Montmollin, O. (1989). The Archaeology of Political Structure: Settlement Analysis in a Classic Maya
Polity, Cambridge University Press, Cambridge, UK.
Dean, J. S., Gumerman, G. J., Epstein, J. M., Axtell, R. L., Swedlund, A. C., Parker, M. T., and
McCarroll, S. (2000). Understanding Anasazi culture change through agent-based modeling. In
Kohler, T. A., and Gumerman, G. J. (eds.), Dynamics in Human Societies, Oxford University Press,
Oxford, pp. 179–205.
Dewar, R. E. (1991). Incorporating variation in occupation span into settlement-pattern analysis.
American Antiquity 56: 604–620.
Dewar, R. E., and McBride, K. A. (1992). Remnant settlement patterns. In Rossignol, J., and Wandsnider,
L. (eds.), Space, Time, and Archaeological Landscapes, Plenum, New York, pp. 227–256.
Dixon, E. J., Manley, W. F., and Lee, C. M. (2005). The emerging archaeology of glaciers and ice
patches: Examples from Alaska’s Wrangell-St. Elias National Park and Preserve. American
Antiquity 70: 129–143.
Drennan, R. D., and Peterson, C. E. (2004). Comparing archaeological settlement systems with rank-size
graphs: A measure of shape and statistical confidence. Journal of Archaeological Science 31:
533–549.
Duff, A. I. (1996). Ceramic micro-seriation: Types or attributes? American Antiquity 61: 89–101.
Duff, A. I. (2000). Scale, interaction, and regional analysis in late Pueblo prehistory. In Hegmon, M. (ed.),
The Archaeology of Regional Interaction: Religion, Warfare, and Exchange across the American
Southwest, University Press of Colorado, Boulder, pp. 71–98.
Duncan, R. B., and Beckman, K. A. (2000). The application of GIS predictive site location models within
Pennsylvania, West Virginia. In Westcott, K. L., and Brandon, R. J. (eds.), Practical Applications of
GIS for Archaeologists: A Predictive Modeling Kit, Taylor and Francis, Philadelphia, pp. 33–58.
Dunnell, R. C. (1992). The notion site. In Rossignol, J., and Wandsnider, L. (eds.), Space, Time, and
Archaeological Landscapes, Plenum, New York, pp. 21–42.
Dunnell, R. C., and Dancey, W. S. (1983). The siteless survey: A regional scale data collection strategy.
In Schiffer, M. B. (ed.), Advances in Archaeological Method and Theory, Academic Press, New
York, pp. 267–287.
Earle, T. K. (1976). A nearest-neighbor analysis of two Formative settlement systems. In Flannery, K. V.
(ed.), The Early Mesoamerican Village, Academic Press, New York, pp. 196–222.
Ebdon, D. (1985). Statistics in Geography, Basil Blackwell, Oxford.
Ebert, D., and Singer, M. (2004). GIS, predictive modeling, erosion, site monitoring. Assemblage 8.
Journal online. Available from http://www.assemblage.group.shef.ac.uk/issue8/ebertandsinger.html.
Accessed 28 January 2007.
Ebert, J. I. (1992). Distributional Archaeology, University of New Mexico Press, Albuquerque.
Ebert, J. I. (2000). The state of the art in ‘‘inductive’’ predictive modeling: Seven big mistakes (and lots of
smaller ones). In Westcott, K. L., and Brandon, R. J. (eds.), Practical Applications of GIS for
Archaeologists: A Predictive Modeling Kit, Taylor and Francis, London, pp. 129–134.
Ebert, J. I., Camilli, E. L., and Berman, M. J. (1996). GIS in the analysis of distributional archaeological
data. In Maschner, H. D. G. (ed.), New Methods, Old Problems: Geographic Information Systems in
Modern Archaeological Research, Occasional Paper No. 23, Center for Archaeological Investiga-
tions, Southern Illinois University, Carbondale, pp. 25–37.
Eddy, F. W., Lightfoot, D. R., Welker, E. A., Wright, L. L., and Torres, D. C. (1996). Air photographic
mapping of San Marcos Pueblo. Journal of Field Archaeology 23: 1–13.
Eerkens, J. W., Neff, H., and Glascock, M. D. (2002). Ceramic production among small-scale and mobile
hunters and gatherers: A case study from the southwestern Great Basin. Journal of Anthropological
Archaeology 21: 200–229.
Elliott, M. (2005). Evaluating evidence for warfare and environmental stress in settlement pattern data
from the Malpaso valley, Zacatecas, Mexico. Journal of Anthropological Archaeology 24: 297–315.
Elston, R. G., and Zeanah, D. W. (2002). Thinking outside the box: A new perspective on diet breadth and
sexual division of labor in the Prearchaic Great Basin. World Archaeology 34: 103–130.

123
66 J Archaeol Res (2008) 16:37–81

Emerson, T. E., Hughes, R. E., Hynes, M. R., and Wisseman, S. U. (2003). The sourcing and
interpretation of Cahokia-style figurines in the trans-Mississippi South and Southeast. American
Antiquity 68: 287–313.
Faught, M. K. (2004). The underwater archaeology of paleolandscapes, Apalachee Bay, Florida.
American Antiquity 69: 275–289.
Feinman, G. M., and Nicholas, L. M. (1999). Reflections on regional survey: Perspectives from the
Guirún area, Oaxaca, Mexico. In Billman, B. R., and Feinman, G. R. (eds.), Settlement Pattern
Studies in the Americas: Fifty Years Since Virú, Smithsonian Institution Press, Washington, DC, pp.
172–190.
Ferguson, T. J., and Colwell-Chanthaphonh, C. (2006). History is in the Land: Multivocal Tribal
Traditions in Arizona’s San Pedro Valley, University of Arizona Press, Tucson.
Field, J. S. (2004). Environmental and climatic considerations: A hypothesis for conflict and the
emergence of social complexity in Fijian prehistory. Journal of Anthropological Archaeology 23:
79–99.
Finsten, L., and Kowalewski, S. A. (1999). Spatial scales and process: In and around the valley of Oaxaca.
In Billman, B. R., and Feinman, G. M. (eds.), Settlement Pattern Studies in the Americas: Fifty
Years Since Virú, Smithsonian Institution Press, Washington, DC, pp. 22–35.
Fish, S. K., and Kowalewski, S. A. (eds.) (1990). The Archaeology of Regions: A Case for Full-Coverage
Survey, Smithsonian Institution Press, Washington, DC.
Fisher, C. T. (2005). Demographic and landscape change in the Lake Patzcuaro Basin, Mexico:
Abandoning the garden. American Anthropologist 107: 87–95.
Fisher, C. T., and Feinman, G. M. (2005). Introduction to ‘‘Landscapes over Time.’’ American
Anthropologist 107: 62–69.
Flannery, K. V. (1968). Archaeological systems theory in early Mesoamerica. In Meggers, B. J. (eds.),
Anthropological Archaeology in the Americas, Anthropological Society of Washington, Washing-
ton, DC, pp. 67–87.
Flannery, K. V. (1976). Evolution of complex settlement systems. In Flannery, K. V. (eds.), The Early
Mesoamerican Village, Academic Press, New York, pp. 162–173.
Fowler, M. J. F. (2002). Satellite remote sensing and archaeology: A comparative study of satellite
imagery of the environs of Figsbury Ring, Wiltshire. Archaeological Prospection 9: 55–69.
Fox, C. (1923). The Archaeology of the Cambridge Region, Cambridge University Press, Cambridge, UK.
Fox, C. (1932). The Personality of Britain, National Museum of Wales, Cardiff.
Gaffney, V. L., and van Leusen, M. (1995). Postscript: GIS, environmental determinism, and
archaeology: A parallel text. In Lock, G., and Stancic, Z. (eds.), Archaeology and Geographical
Information Systems: A European Perspective, Taylor and Francis, Oxford, pp. 367–376.
Gaffney, V. L., Stancic, Z., and Watson, H. (1996). Moving from catchments to cognition: Tentative steps
toward a larger archaeological context for GIS. In Aldenderfer, M. S., and Maschner, M. D. G.
(eds.), Anthropology, Space, and Geographic Information Systems, Occasional Paper No. 23, Center
for Archaeological Investigations, Oxford University Press, New York, pp. 132–154.
Galaty, M. L. (2005). European regional studies: A coming of age? Journal of Archaeological Research
13: 291–336.
Gallivan, M. D. (2002). Measuring sedentariness and settlement population: Accumulations research in
the middle Atlantic region. American Antiquity 67: 535–557.
Geib, P. R. (2000). Sandal types and Archaic prehistory on the Colorado Plateau. American Antiquity 65:
509–524.
Gillings, M., and Wheatley, D. (2005). Geographic information systems. In Maschner, H. D. G., and
Chippendale, C. (eds.), Handbook of Theories and Methods in Archaeology, AltaMira, Walnut
Creek, CA, pp. 373–422.
Given, M. (2004). Mapping and manuring: Can we compare sherd density figures? In Alcock, S. E., and
Cherry, J. F. (eds.), Side-by-Side Survey: Comparative Regional Studies in the Mediterranean
World, Oxbow Books, Oxford, pp. 13–21.
Goodchild, M. F., and Janelle, D. G. (eds.) (2004). Spatially Integrated Social Science: Examples in Best
Practice, Oxford University Press, Oxford.
Green, S. W., and Perlman, S. M. (eds.) (1985). The Archaeology of Frontiers and Boundaries, Academic
Press, Orlando, FL.
Gumerman, G. J., Swedlund, A. C., Dean, J. S., and Epstein, J. M. (2003). The evolution of social
behavior in the prehistoric American Southwest. Artificial Life 9: 435–444.

123
J Archaeol Res (2008) 16:37–81 67

Hardesty, D. L., and Fowler, D. D. (2001). Archaeology and environmental changes. In Crumley, C. L.
(ed.), New Directions in Anthropology and Environment: Intersections, AltaMira, Walnut Creek,
CA, pp. 72–89.
Hare, T. S. (2004). Using measures of cost distance in the estimation of polity boundaries in the
Postclassic Yautepec Valley, Mexico. Journal of Archaeological Science 31: 799–814.
Hargrave, M. L., Somers, L. E., Larson, T., Shields, R., and Dendy, J. (2002). The role of resistivity
survey in historic site assessment and management: An example from Fort Riley, Kansas. Historical
Archaeology 36: 89–110.
Harmon, J. M., Leone, M. P., Prince, S. D., and Snyder, M. (2006). LiDAR for archaeological landscape
analysis: A case study of two eighteenth-century Maryland plantation sites. American Antiquity 71:
649–670.
Hegmon, M., Hays-Gilpin, K. A., McGuire, R. H., Rautman, A. E., and Schlanger, S. H. (2000).
Changing perceptions of regional interaction in the prehistoric Southwest. In Hegmon, M. (ed.), The
Archaeology of Regional Interaction: Religion, Warfare, and Exchange Across the American
Southwest and Beyond, University of Colorado Press, Boulder, pp. 1–21.
Henderson, H., and Ostler, N. (2005). Muisca settlement organization and chiefly authority at Suta, Valle
de Leyva, Colombia: A critical appraisal of native concepts of house for studies of complex
societies. Journal of Anthropological Archaeology 24: 148–178.
Hill, J. B. (2000). Decision making at the margins: Settlement trends, temporal scale, and ecology in the
Wadi al Hasa, west-central Jordan. Journal of Anthropological Archaeology 19: 221–241.
Hill, J. B. (2004). Land use and an archaeological perspective on socio-natural studies in the Wadi
Al-Hasa, west-central Jordan. American Antiquity 69: 389–412.
Hill, J. B., Clark, J. J., Doelle, W. H., and Lyons, P. D. (2004). Prehistoric demography in the Southwest:
Migration, coalescence, and Hohokam population decline. American Antiquity 69: 689–716.
Hill, J. N. (1970). Broken K Pueblo: Prehistoric Social Organization in the American Southwest,
University of Arizona Press, Tucson.
Hodder, I. (1979). Simulating the growth of hierarchies. In Renfrew, C., and Cooke, K. L. (eds.),
Transformations: Mathematical Approaches to Culture Change, Academic Press, New York, pp.
117–144.
Hodder, I., and Orton, C. (1976). Spatial Analysis in Archaeology, Cambridge University Press,
Cambridge, UK.
Holdaway, S., and Fanning, P. (2004). Developing a landscape history as part of a survey strategy:
Examples from western New South Wales, Australia. Paper presented at the 69th Annual Meeting of
the Society for American Archaeology, Montreal, Canada.
Howey, M. C. L., and O’Shea, J. M. (2006). Bear’s journey and the study of ritual in archaeology.
American Antiquity 71: 261–282.
Jennings, J. (2006). Core, peripheries, and regional realities in Middle Horizon Peru. Journal of
Anthropological Archaeology 25: 346–370.
Jennings, J., and Craig, N. (2001). Politywide analysis and imperial political economy: The relationship
between valley political complexity and administrative centers in the Wari empire of the central
Andes. Journal of Anthropological Archaeology 20: 479–502.
Johansen, K. L., Laursen, S. T., and Holst, M. K. (2004). Spatial patterns of social organization in the
Early Bronze Age of south Scandinavia. Journal of Anthropological Archaeology 23: 33–55.
Johansen, P. G. (2004). Landscape, monumental architecture, and ritual: A reconsideration of the South
Indian ashmounds. Journal of Anthropological Archaeology 23: 309–330.
Johnson, C. D., Kohler, T. A., and Cowan, J. (2005). Modeling historical ecology, thinking about
contemporary systems. American Anthropologist 107: 96–107.
Johnson, G. A. (1977). Aspects of regional analysis in archaeology. Annual Review of Anthropology 6:
479–508.
Johnson, G. A. (1980). Rank-size convexity and system integration: A view from archaeology. Economic
Geography 56: 234–247.
Jones, G. T., Beck, C., Jones, E. E., and Hughes, R. E. (2003). Lithic source use and paleoarchaic foraging
territories in the Great Basin. American Antiquity 68: 5–38.
Jordan, P., and Shennan, S. (2003). Cultural transmission, language, and basketry traditions amongst the
California Indians. Journal of Anthropological Archaeology 22: 42–74.
Kantner, J. (1996). Settlement pattern analysis. In Fagan, B. (ed.), The Oxford Companion to
Archaeology, Oxford University Press, New York, pp. 636–638.

123
68 J Archaeol Res (2008) 16:37–81

Kantner, J. (2004). Geographical approaches for reconstructing past human behavior from prehistoric
roadways. In Goodchild, M. F., and Janelle, D. G. (eds.), Spatially Integrated Social Science:
Examples in Best Practice, Oxford University Press, Oxford, pp. 323–344.
Kantner, J. (2005). Regional analysis in archaeology. In Maschner, H. D. G., and Chippendale, C. (eds.),
Handbook of Theories and Methods in Archaeology, AltaMira, Walnut Creek, CA, pp. 1179–1223.
Kantner, J., and Hobgood, R. (2003). Digital technologies and prehistoric landscapes in the American
Southwest. In Forte, M., Williams, P. R., and Wiseman, J. (eds.), The Reconstruction of
Archaeological Landscapes through Digital Technologies, Archaeopress, Oxford, pp. 117–123.
Kantner, J., Bower, N., Ladwig, J., Perlitz, J., Hata, S., and Greve, D. (2000). Interaction between great
house communities: An elemental analysis of Cibolan ceramics. In Kantner, J., and Mahoney, N. M.
(eds.), Great House Communities across the Chacoan Landscape, University of Arizona Press,
Tucson, pp. 130–146.
Kelso, W. M., and Most, R. (eds.) (1990). Earth Patterns: Essays in Landscape Archaeology, University
of Virginia Press, Charlottesville.
Kennett, D. J., Sakai, S., Neff, H., Gossett, R., and Larson, D. O. (2002). Compositional characterization
of prehistoric ceramics: A new approach. Journal of Archaeological Science 29: 443–455.
Kim, J. (2003). Land-use conflict and the rate of the transition to agricultural economy: A comparative
study of southern Scandinavia and central-western Korea. Journal of Archaeological Method and
Theory 10: 277–321.
King, T. (2003). Places that Count: Traditional Cultural Properties, Rowman and Littlefield, Lanham,
MD.
Kintigh, K. W. (1988). The effectiveness of subsurface testing: A simulation approach. American
Antiquity 53: 686–707.
Kintigh, K. W., Glowacki, D. M., and Huntley, D. L. (2004). Long-term settlement history and the
emergence of towns in the Zuni area. American Antiquity 69: 432–456.
Knapp, A. B., and Ashmore, W. (1999). Archaeological landscapes: Constructed, conceptualized,
ideational. In Ashmore, W., and Knapp, A. B. (eds.), Archaeologies of Landscape: Contemporary
Perspectives, Blackwell, Oxford, pp. 1–31.
Kohler, T. A. (1992). Prehistoric human impact on the environment in the upland North American
Southwest. Population and Environment: A Journal of Interdisciplinary Studies 13: 255–268.
Kohler, T. A., Kresl, J., Van West, C. R., Carr, E., and Wilshusen, R. H. (2000). Be there then: A
modeling approach to settlement determinants and spatial efficiency among late Ancestral Pueblo
populations of the Mesa Verde region, U.S. Southwest. In Kohler, T. A., and Gumerman, G. J.
(eds.), Dynamics in Human Societies, Oxford University Press, Oxford, pp. 145–178.
Kohler, T. A., and Matthews, M. H. (1988). Long-term Anasazi land use and forest reduction: A case
study from southwest Colorado. American Antiquity 53: 537–564.
Kolb, M. J., and Snead, J. E. (1997). It’s a small world after all: Comparative analyses of community
organization in archaeology. American Antiquity 62: 609–628.
Kooyman, B. (2006). Boundary theory as a means to understanding social space in archaeological sites.
Journal of Anthropological Archaeology 25: 424–436.
Kowalewski, S. A. (2003). Scale and explanation of demographic change: 3,500 years in the Valley of
Oaxaca. American Anthropologist 105: 313–325.
Kuijt, I. (2000). People and space in early agricultural villages: Exploring daily lives, community size,
and architecture in the Late Pre-Pottery Neolithic. Journal of Anthropological Archaeology 19:
75–102.
Kulisheck, J. (2003). Pueblo population movements, abandonment and settlement change in sixteenth and
seventeenth century New Mexico. Kiva 69: 30–54.
Kvamme, K. L. (1989). Geographic information systems in regional archaeological research, data
management. In Schiffer, M. B. (ed.), Archaeological Method and Theory, University of Arizona
Press, Tucson, pp. 139–203.
Kvamme, K. L. (1996). Investigating chipping debris scatters: GIS as an analytical tool. In Maschner, H.
D. G. (ed.), New Methods, Old Problems: Geographic Information Systems in Modern Archae-
ological Research, Occasional Paper No. 23, Center for Archaeological Investigations, Southern
Illinois University, Carbondale, pp. 38–71.
Kvamme, K. L. (1997). Ranter’s corner: Bringing the camps together: GIS and ED. Archaeological
Computing Newsletter 47: 1–5.

123
J Archaeol Res (2008) 16:37–81 69

Kvamme, K. L. (1999). Recent directions and developments in geographical information systems.


Journal of Archaeological Research 7: 153–201.
Kvamme, K. L. (2003). Geophysical surveys as landscape archaeology. American Antiquity 68: 435–457.
Ladefoged, T. N., and Graves, M. W. (2000). Evolutionary theory and the historical development of dry-
land agriculture in north Kohala, Hawai’i. American Antiquity 65: 423–448.
Lake, M. W., Woodman, P. E., and Mithen, S. J. (1998) Tailoring GIS software for archaeological
applications: An example concerning viewshed analysis. Journal of Archaeological Science 25:
27–38.
Lee, Y. K. (2004). Control strategies and polity competition in the lower Yi-Luo Valley, north China.
Journal of Anthropological Archaeology 23: 172–195.
Lekson, S. H. (1999). The Chaco Meridian: Centers of Political Power in the Ancient Southwest,
AltaMira, Walnut Creek, CA.
Lekson, S. H., and Cameron, C. M. (1995). The abandonment of Chaco Canyon, the Mesa Verde
migrations, and the reorganization of the Pueblo world. Journal of Anthropological Archaeology 14:
184–202.
Lekson, S. H., Nepstad-Thornberry, C. P., Yunker, B. E., Laumbach, T. S., Cain, D. P., and Laumbach, K.
W. (2002). Migrations in the Southwest: Pinnacle Ruin, southwestern New Mexico. Kiva 68: 73–101.
Lewarch, D. E., and O’Brien, M. J. (1981). The expanding role of surface assemblages in archaeological
research. In Schiffer, M. B. (ed.), Advances in Archaeological Method and Theory, Academic Press,
New York, pp. 297–342.
Lewis, R. B. (2000). Sea-level rise and subsidence effects on Gulf Coast archaeological site distributions.
American Antiquity 65: 525–541.
Lightfoot, K. G., and Martinez, A. (1995). Frontiers and boundaries in archaeological perspective. Annual
Review of Anthropology 24: 471–492.
Lloyd, C. D., and Atkinson, P. M. (2004). Archaeology and geostatistics. Journal of Archaeological
Science 31: 151–156.
Lock, G. (2003). Using Computers in Archaeology: Towards Virtual Pasts, Routledge, London.
Mack, A. (2004). One landscape, many experiences: Differing perspectives of the temple districts of
Vijayanagara. Journal of Archaeological Method and Theory 11: 59–81.
Madry, S. L. H., and Crumley, C. L. (1990). An application of remote sensing and GIS in a regional
archaeological settlement pattern analysis: The Arroux River valley, Burgundy, France. In Allen, K.
M. S., Green, S. W., and Zubrow, E. B. W. (eds.), Interpreting Space: GIS and Archaeology, Taylor
and Francis, London, pp. 364–380.
Madry, S. L. H., and Rakos, L. (1996). Line-of-sight and cost-surface techniques for regional research in
the Arroux River valley. In Maschner, H. D. G. (ed.), New Methods, Old Problems: Geographic
Information Systems in Modern Archaeological Research, Occasional Paper No. 23, Center for
Archaeological Investigations, Southern Illinois University, Carbondale, pp. 104–126.
Mahoney, N. M. (2000). Redefining the scale of Chacoan communities. In Kantner, J., and Mahoney, N.
M. (eds.), Great House Communities Across the Chacoan Landscape, University of Arizona Press,
Tucson, pp. 17–27.
Mahoney, N. M., Adler, M. A., and Kendrick, J. W. (2000). The changing scale and configuration of
Mesa Verde communities. Kiva 66: 67–90.
Martin, P. S., Thompson, R. S., and Long, A. (1985). Shasta ground sloth extinction: A test of the blitzkrieg
model. In Mead, J. I., and Meltzer, D. J. (eds.), Environments and Extinctions: Man in Late Glacial
North America, Center for the Study of Early Man, University of Maine, Orono, pp. 5–14.
Maschner, H. D. G. (ed.) (1996). New Methods, Old Problems: Geographic Information Systems in
Modern Archaeological Research, Occasional Paper No. 23, Center for Archaeological Investiga-
tions, Southern Illinois University, Carbondale.
McGuire, R. H., Adams, E. C., Nelson, B. A., and Spielmann, K. A. (1994). Drawing the Southwest to
scale: Perspectives on macroregional relations. In Gumerman, G. J. (ed.), Themes in Southwest
Prehistory, School of American Research Press, Santa Fe, NM, pp. 239–266.
Milner, G. R., and Oliver, J. S. (1999). Late prehistoric settlements and wetlands in the central Mississippi
Valley. In Billman, B. R., and Feinman, G. M. (eds.), Settlement Pattern Studies in the Americas:
Fifty Years Since Virú, Smithsonian Institution Press, Washington, DC, pp. 79–95.
Muir, R. J., and Driver, J. C. (2002). Scale of analysis and zooarchaeological interpretation: Pueblo III
faunal variation in the northern San Juan region. Journal of Anthropological Archaeology 21:
165–199.

123
70 J Archaeol Res (2008) 16:37–81

Neitzel, J. E. (1994). Boundary dynamics in the Chacoan regional system. In Wills, W. H., and Leonard,
R. D. (eds.), The Ancient Southwestern Community, University of New Mexico Press, Albuquerque,
pp. 209–240.
Nelson, M. C., and Hegmon, M. (2001). Abandonment is not as it seems: An approach to the relationship
between site and regional abandonment. American Antiquity 66: 213–235.
Nelson, M. C., and Schollmeyer, K. G. (2003). Game resources, social interaction, and the ecological
footprint in southwest New Mexico. Journal of Archaeological Method and Theory 10: 69–110.
Nichols, D. L., Brumfiel, E. M., Neff, H., Hodge, M., Charlton, T. H., and Glascock, M. D. (2002).
Neutrons, markets, cities, and empires: A 1000-year perspective on ceramic production and
distribution in the Postclassic Basin of Mexico. Journal of Anthropological Archaeology 21: 25–82.
Norman, N. L., and Kelly, K. G. (2004). Landscape politics: The serpent ditch and the rainbow in West
Africa. American Anthropologist 106: 98–110.
Obenauf, M. S. (1980). A history of research on the Chacoan roadway system. In Lyons, T. R., and
Mathien, F. J. (eds.), Cultural Resources Remote Sensing, National Park Service, Cultural Resources
Management Division, Washington, DC, pp. 123–167.
Osborne, R. (2004). Demography and survey. In Alcock, S. E., and Cherry, J. F. (eds.), Side-by-Side
Survey: Comparative Regional Studies in the Mediterranean World, Oxbow Books, Oxford, pp.
163–172.
Parker, B. J. (2002). At the edge of empire: Conceptualizing Assyria’s Anatolian frontier ca. 700 BC.
Journal of Anthropological Archaeology 21: 371–395.
Parsons, J. R. (1972). Archaeological settlement patterns. Annual Review of Anthropology 1: 127–150.
Parsons, J. R. (2004). Critical reflections on forty years of ‘‘systematic regional survey.’’ Paper presented
at the 69th Annual Meeting of the Society for American Archaeology, Montreal, Canada.
Paynter, R. W. (1983). Expanding the scope of settlement analysis. In Moore, J. A., and Keene, A. S.
(eds.), Archaeological Hammers and Theories, Academic Press, New York, pp. 233–275.
Perkins, P. (2000). A GIS investigation of site locations and landscape relationships in the Albegna
Valley, Tuscany. In Lockyear, K., Sly, T. J. T., and Mihailescu-Birliba, V. (eds.), Computer
Application and Quantitative Methods in Archaeology, Archaeopress, Oxford, pp. 133–140.
Peterson, C. E., and Drennan, R. D. (2005). Communities, settlements, sites, and surveys: Regional-scale
analysis of prehistoric human interaction. American Antiquity 70: 5–30.
Pinhasi, R., and Pluciennik, M. (2004). A regional biological approach to the spread of farming in Europe.
Current Anthropology 45: S59–S82.
Potter, J. M. (2004). The creation of person, the creation of place: Hunting landscapes in the American
Southwest. American Antiquity 69: 322–338.
Redman, C. L. (1999). Human Impact on Ancient Environments, University of Arizona Press, Tucson.
Rhoades, R. E. (1978). Archaeological use and abuse of ecological concepts and studies: The ecotone
example. American Antiquity 43: 608–614.
Roper, D. C. (1979). The method and theory of site catchment analysis: A review. In Schiffer, M. B. (ed.),
Advances in Archaeological Method and Theory, Academic Press, New York, pp. 119–140.
Ruggles, A. J., and Church, R. L. (1996). Spatial allocation in archaeology: An opportunity for
reevaluation. In Maschner, H. D. G. (ed.), New Methods, Old Problems: Geographic Information
Systems in Modern Archaeological Research, Occasional Paper No. 23, Center for Archaeological
Investigations, Southern Illinois University, Carbondale, pp. 147–176.
Ruggles, C. L. N., and Medyckyj-Scott, D. J. (1996). Site location, landscape visibility, and symbolic
astronomy: A Scottish case study. In Maschner, H. D. G. (ed.), New Methods, Old Problems:
Geographic Information Systems in Modern Archaeological Research, Occasional Paper No. 23,
Center for Archaeological Investigations, Southern Illinois University, Carbondale, pp. 127–146.
Runnels, C. (2000). Anthropogenic soil erosion in prehistoric Greece: The contribution of regional
surveys to the archaeology of environmental disruptions, human response. In Bawden, G., and
Reycraft, R. M. (eds.), Environmental Disaster and the Archaeology of Human Response, Maxwell
Museum of Anthropology, Albuquerque, NM, pp. 11–20.
Sanders, W. T. (1956). The central Mexican symbiotic region: A study in prehistoric settlement patterns.
In Willey, G. R. (ed.), Prehistoric Settlement Patterns in the New World, Viking Fund Publications
in Anthropology, New York, pp. 115–127.
Sanders, W. T. (1965). The cultural ecology of the Teotihuacán Valley. Unpublished ms. on file,
Department of Anthropology, Pennsylvania State University, University Park.

123
J Archaeol Res (2008) 16:37–81 71

Sanders, W. T. (1999). Three valleys: Twenty-five years of settlement archaeology in Mesoamerica. In


Billman, B. R., and Feinman, G. M. (eds.), Settlement Pattern Studies in the Americas: Fifty Years
Since Virú, Smithsonian Institution Press, Washington, DC, pp. 12–21.
Sauer, C. O. (1925). The morphology of landscape. University of California Publications in Geography 2:
19–53.
Scarre, C. (2001). Modeling prehistoric populations: The case of Neolithic Brittany. Journal of
Anthropological Archaeology 20: 285–313.
Sever, T. L., and Irwin, D. E. (2003). Landscape archaeology: Remote-sensing investigation of the
ancient Maya in the Peten rainforest of northern Guatemala. Ancient Mesoamerica 14: 113–122.
Shackel, P. A. (ed.) (2001). Myth, Memory, and the Making of the American Landscape, University Press
of Florida, Gainesville.
Shackel, P. A., and Chambers, E. J. (eds.) (2004). Places in Mind: Public Archaeology as Applied
Anthropology, Routledge, London.
Silliman, S. W., Farnsworth, P., and Lightfoot, K. G. (2000). Magnetometer prospecting in historical
archaeology: Evaluating survey options at a 19th century rancho site in California. Historical
Archaeology 34: 89–109.
Smith, C. A. (ed.) (1976). Regional Analysis, Academic Press, New York.
Smith, M. E., and Montiel, L. (2001). The archaeological study of empires and imperialism in pre-
Hispanic central Mexico. Journal of Anthropological Archaeology 20: 245–284.
Spencer, C. S., and Redmond, E. M. (2001). Multilevel selection and political evolution in the Valley of
Oaxaca, 500–100 B.C. Journal of Anthropological Archaeology 20: 195–229.
Stafford, C. R., and Hajic, E. R. (1992). Landscape scale: Geoenvironmental approaches to prehistoric
settlement strategies. In Rossignol, J., and Wandsnider, L. (eds.), Space, Time, and Archaeological
Landscapes, Plenum, New York, pp. 137–166.
Stancic, Z., and Kvamme, K. L. (1999). Settlement patterns modelling through Boolean overlays of social
and environmental variables. In Barcelo, J. A., Briz, I., and Vila, A. (eds.), New Technique for Old
Times, Archaeopress, Oxford, pp. 231–237.
Stanish, C. (2001). Regional research on the Inca. Journal of Archaeological Research 9: 213–241.
Stein, G. J. (1999). Rethinking World-Systems: Diaspora, Colonies, and Interaction in Uruk
Mesopotamia, University of Arizona Press, Tucson.
Steponaitis, V. P., and Kintigh, K. W. (1993). Estimating site occupation spans from dated artifact types:
Some new approaches. In Stoltman, J. B. (ed.), Archaeology of Eastern North America: Papers in
Honor of Stephen Williams, Mississippi Department of Archives and History, Jackson, pp. 349–361.
Steward, J. H. (1937). Ecological aspects of Southwestern society. Anthropos 32: 87–104.
Steward, J. H. (1938). Basin-Plateau Aboriginal Sociopolitical Groups, Bulletin 120, Bureau of American
Ethnology, Washington, DC.
Stewart, A. M., Keith, D., and Scottie, J. (2004). Caribou crossings and cultural meanings: Placing
traditional knowledge and archaeology in context in an Inuit landscape. Journal of Archaeological
Method and Theory 11: 183–211.
Swanson, S. (2003). Documenting prehistoric communication networks: A case study in the Paquimé
polity. American Antiquity 68: 753–767.
Thomas, J. (1996). Time, Culture, and Identity, Routledge, London.
Thomas, J. (2001). Archaeologies of place and landscape. In Hodder, I. (ed.), Archaeological Theory
Today, Polity, Cambridge, pp. 165–186.
Thompson, S. (2004). Side-by-side and back-to-front: Exploring intra-regional latitudinal and longitu-
dinal comparability in survey data—Three case studies from Metaponto, Italy. In Alcock, S. E., and
Cherry, J. F. (eds.), Side-by-Side Survey: Comparative Regional Studies in the Mediterranean
World, Oxbow Books, Oxford, pp. 65–85.
Tilley, C. F. (1994). A Phenomenology of Landscape: Places, Paths, and Monuments, Berg, Oxford.
Tomczak, P. D. (2003). Prehistoric diet and socioeconomic relationships within the Osmore Valley of
southern Peru. Journal of Anthropological Archaeology 22: 262–278.
Truncer, J. (2004). Steatite vessel age and occurrence in temperate eastern North America. American
Antiquity 69: 487–513.
Varien, M. D. (1999). Sedentism and Mobility in a Social Landscape: Mesa Verde and Beyond,
University of Arizona Press, Tucson.
Varien, M. D., and Mills, B. J. (1997). Accumulations research: Problems and prospects for estimating
site occupation span. Journal of Archaeological Method and Theory 4: 141–191.

123
72 J Archaeol Res (2008) 16:37–81

Varien, M. D., Van West, C. R., and Patterson, G. S. (2000). Competition, cooperation, and conflict:
Agricultural production and community catchments in the central Mesa Verde region. Kiva 66:
45–66.
Viti-Finzi, C., and Higgs, E. S. (1970). Prehistoric economy in the Mount Carmel area of Palestine: Site
catchment analysis. Proceedings of the Prehistoric Society 36: 1–37.
Wandsnider, L. (1992). The spatial dimension of time. In Rossignol, J., and Wandsnider, L. (eds.), Space,
Time, and Archaeological Landscapes, Plenum, New York, pp. 257–284.
Wandsnider, L. (1998). Regional scale processes and archaeological landscape units. In Ramenofsky, A.
F., and Steffen, A. (eds.), Unit Issues in Archaeology: Measuring Time, Space, and Material,
University of Utah Press, Salt Lake City, pp. 87–102.
Wandsnider L. (2004). Solving the puzzle of the archaeological labyrinth: Time perspectivism in
Mediterranean surface archaeology. In Alcock, S. E., and Cherry, J. F. (eds.), Side-by-Side Survey:
Comparative Regional Studies in the Mediterranean World, Oxbow Book, Oxford, pp. 49–62.
Warren, R. E., and Asch, D. L. (2000). A predictive model of archaeological site location in the eastern
Prairie Peninsula. In Westcott, K. L., and Brandon, R. J. (eds.), Practical Applications of GIS for
Archaeologists: A Predictive Modeling Kit, Taylor and Francis, Philadelphia, pp. 5–32.
Washburn, D. K. (1974). Nearest neighbor analysis of Pueblo I–III settlement patterns along the Rio
Puerco of the East, New Mexico. American Antiquity 39: 315–335.
Waters, M. R., and Ravesloot, J. C. (2001). Landscape change and the cultural evolution of the Hohokam
along the middle Gila River and other river valleys in south-central Arizona. American Antiquity 66:
285–299.
Wells, E. C., Rice, G. E., and Ravesloot, J. C. (2004). Peopling landscapes between villages in the middle
Gila River valley of central Arizona. American Antiquity 69: 627–652.
Westcott, K. L., and Brandon, R. J. (eds.) (2000). Practical Applications of GIS for Archaeologists: A
Predictive Modeling Kit, Taylor and Francis, London.
Whalen, M. E., and Minnis, P. E. (2003). The local and the distant in the origin of Casas Grandes,
Chihuahua, Mexico. American Antiquity 68: 314–332.
Whallon, R. E. (1968). Investigations of late prehistoric social organization in New York state. In
Binford, S. R., and Binford, L. R. (eds.), New Perspectives in Archeology, Aldine, Chicago, pp.
223–244.
Whallon, R. E. (1984). Unconstrained clustering for the analysis of spatial distributions in archaeology. In
Hietala, H. J. (ed.), Intrasite Spatial Analysis in Archaeology, Cambridge University Press,
Cambridge, UK, pp. 242–277.
Wheatley, D., and Gillings, M. (2002). Spatial Technology and Archaeology: The Archaeological
Applications of GIS, Taylor and Francis, New York.
Whitridge, P. (2001). Zen fish: A consideration of the discordance between artifactual and zooarchae-
ological indicators of Thule Inuit fish use. Journal of Anthropological Archaeology 20: 3–72.
Whitridge, P. (2004). Landscapes, houses, bodies, things: ‘‘Place’’ and archaeology of Inuit imaginaries.
Journal of Archaeological Method and Theory 11: 213–250.
Wilkinson, T. J., Ur, J., and Casana, J. (2004). From nucleation to dispersal: Trends in settlement pattern
in the northern Fertile Crescent. In Alcock, S. E., and Cherry, J. F. (eds.), Side-by-Side Survey:
Comparative Regional Studies in the Mediterranean World, Oxbow Books, Oxford, pp. 189–205.
Willey, G. R. (1949). Prehistoric Settlement Patterns in the Virú Valley, Peru, Bulletin 115, Bureau of
American Ethnology, Washington, DC.
Willey, G. R. (1999). The Virú Valley project and settlement archaeology: Some reminiscences and
contemporary comments. In Billman, B. R., and Feinman, G. M. (eds.), Settlement Pattern Studies
in the Americas: Fifty Years Since Virú, Smithsonian Institution Press, Washington, DC, pp. 9–11.
Winthrop, K. R. (2001). Historical ecology: Landscapes of change in the Pacific Northwest. In Crumley,
C. L. (ed.), New Directions in Anthropology and Environment: Intersections, AltaMira, Walnut
Creek, CA, pp. 203–222.
Wobst, H. M. (1974). Boundary conditions for paleolithic social systems: A simulation approach.
American Antiquity 39: 147–178.
Wonderley, A. (2005). Effigy pipes, diplomacy, and myth: Exploring interaction between St. Lawrence
Iroquoians and Eastern Iroquois in New York state. American Antiquity 70: 211–240.
Woodman, P. E., and Woodward, M. (2002). The use and abuse of statistical methods in archaeological
site location modelling. In Wheatley, D., Earl, G., and Poppy, S. (eds.), Contemporary Theory in
Archaeological Computing, Oxbow Books, Oxford, pp. 22–27.

123
J Archaeol Res (2008) 16:37–81 73

Yesner, D. R., Figuerero Torres, M. J., Guichon, R. A., and Borrero, L. A. (2003). Stable isotope analysis
of human bone and ethnohistoric subsistence patterns in Tierra del Fuego. Journal of Anthropo-
logical Archaeology 22: 279–291.
Zeanah, D. W. (2004). Sexual division of labor and central place foraging: A model for the Carson Desert
of western Nevada. Journal of Anthropological Archaeology 23: 1–32.
Zubrow, E. B. W. (1994). Knowledge representation and archaeology: A cognitive example using GIS. In
Renfrew, C., and Zubrow, E. B. W. (eds.), The Ancient Mind: Elements of Cognitive Archaeology,
Cambridge University Press, Cambridge, UK, pp. 107–118.

Bibliography of recent literature

Anaya Hernández, A., Guenter, S. P., and Zender, M. U. (2003). Sak Tz’i’, a Classic Maya center: A
locational model based on GIS and epigraphy. Latin American Antiquity 14: 179–191.
Anderson, D. G., and Faught, M. K. (2004). Paleoindian climate, dating and artifact distribution on a very
large scale: Evidence and implications. Antiquity 74: 507–513.
Andreou, S. (2001). Exploring the patterns of power in the Bronze Age settlements of northern Greece.
In: Branigan, K. (ed.), Urbanism in the Aegean Bronze Age, Sheffield Centre for Aegean
Archaeology, University of Sheffield, Sheffield, pp. 160–173.
Aporta, C. (2003). New ways of mapping: Using GPS mapping software to plot place names and trails in
Igloolik (Nunavut). Arctic 56: 321–327.
Arnold, B. (2002). A landscape of ancestors in southwest Germany. Antiquity 76: 321–322.
Arnold, J. E. (2001). The Chumash in world and regional perspectives. In Arnold, J. E. (ed.), The Origins
of a Pacific Coast Chiefdom, University of Utah Press, Salt Lake City, pp. 1–20.
Aswani, S., and Sheppard, P. (2003). The archaeology and ethnohistory of exchange in precolonial and
colonial Roviana: Gifts, commodities, and inalienable possessions. Current Anthropology 44:
S51–S78.
Attema, P., Burgers, G.-J., and Van Joolen, E. (eds.) (2002). New Developments in Italian Landscape
Archaeology, Archaeopress, Oxford.
Balkansky, A. K., Kowalewski, S. A., Rodriguez, V. P., Pluckhahn, T. J., Smith, C. A., Stiver, L. R.,
Beliaev, D., Chamblee, J. F., Espinoza, V. Y. H., and Perez, R. S. (2000). Archaeological survey in
the Mixteca Alta of Oaxaca, Mexico. Journal of Field Archaeology 27: 365–390.
Barker, G. (2002). A tale of two deserts: Contrasting desertification histories on Rome’s desert frontiers.
World Archaeology 33: 488–507.
Barrett, E. M. (2002). Conquest and Catastrophe: Changing Rio Grande Pueblo Settlement Patterns in
the Sixteenth and Seventeenth Centuries, University of New Mexico Press, Albuquerque.
Barton, C. M., Bernabeu Auban, J., Aura Tortosa, J. E., Garcia, O., and La Roca, N. (2002). Dynamic
landscapes, artifact taphonomy, and landuse modeling in the western Mediterranean. Geoarchae-
ology 17: 155–190.
Barton, C. M., Schmich, S., and James, S. R. (2004). The ecology of human colonization in pristine
landscapes. In Barton, C. M., Clark, G. A., Yesner, D. R., and Pearson, G. (eds.), The Settlement of
the American Continents: A Multidisciplinary Approach to Human Biogeography, University of
Arizona Press, Tucson, pp. 138–161.
Beck, R. A. J. (2003). Consolidation and hierarchy: Chiefdom variability in the Mississippian Southeast.
American Antiquity 68: 641–661.
Beekman, C. S. (2000). The correspondence of regional patterns and local strategies in Formative to
Classic period West Mexico. Journal of Anthropological Archaeology 19: 385–412.
Bell, T., and Lock, G. (2000). Topographic and cultural influences on walking the ridgeway in later
prehistoric times. In: Lock, G. (ed.), Beyond the Map: Archaeology and Spatial Technologies, IOS
Press, Amsterdam, pp. 85–100.
Bender, B. (2001). Landscapes on-the-move. Journal of Social Archaeology 1: 75–89.
Bender, B., and Winer, M. (eds.) (2001). Contested Landscapes: Movement, Exile, and Place, Berg,
Oxford.
Bevan, A., and Conolly, J. (2004). GIS, archaeological survey, and landscape archaeology on the island of
Kythera, Greece. Journal of Field Archaeology 29: 123–138.

123
74 J Archaeol Res (2008) 16:37–81

Binder, D. (2000). Mesolithic and Neolithic interaction in southern France and northern Italy: New data
and current hypotheses. In Price, T. D. (ed.), Europe’s First Farmers, Cambridge University Press,
Cambridge, UK, pp. 117–143.
Bintliff, J., Kuna, M., and Venclová, N. (eds.) (2000). The Future of Surface Artefact Survey in Europe,
Sheffield Academic Press, Sheffield.
Blanton, R. E. (2000). Hellenistic, Roman and Byzantine Settlement Patterns of the Coast of Western
Rough Cilicia, British Archaeological Reports, Oxford.
Boyle, K. (2001). Middle Paleolithic settlement patterns in Mediterranean France: Human geography and
archaeology. In Conard, N. J. (ed.), Settlement Dynamics of the Middle Paleolithic and Middle Stone
Age, Kerns Verlag, Tübingen, pp. 519–543.
Bradley, R. J. (2000). Networks of shell ornament exchange: A critical assessment of prestige economies
in the North American Southwest. In Hegmon, M. (ed.), The Archaeology of Regional Interaction:
Religion, Warfare, and Exchange across the American Southwest and Beyond, University of
Colorado Press, Boulder, pp. 167–188.
Brookes, A., and Smith, A. (eds.) (2001). Holy Ground: Theoretical Issues Relating to the Landscape and
Material Culture of Ritual Space, Archaeopress, Oxford.
Brown, C. T., and Witschey, W. R. T. (2003). The fractal geometry of ancient Maya settlement. Journal
of Archaeological Science 30: 1619–1632.
Brück, J. (ed.) (2001). Bronze Age Landscapes: Tradition and Transformation, Oxbow Books, Oxford.
Buck, P. E., Sabol, D. E., and Gillespie, A. R. (2003). Sub-pixel artifact detection using remote sensing.
Journal of Archaeological Science 30: 973–989.
Burke, A. (2006). Neanderthal settlement patterns in Crimea: A landscape approach. Journal of
Anthropological Archaeology 25: 510–524.
Burmeister, S. (2000). Archaeology and migration: Approaches to an archaeological proof of migration.
Current Anthropology 41: 539–608.
Butzer, K. W. (2005). Environmental history in the Mediterranean world: Cross-disciplinary investigation of
cause-and-effect for degradation and soil erosion. Journal of Archaeological Science 32: 1773–1800.
Campana, S., and Francovich, R. (2003). Landscape archaeology in Tuscany: Cultural resource
management, remotely sensed techniques, GIS based data integration and interpretation. In Forte,
M., and Williams, P. R. (eds.), The Reconstruction of Archaeological Landscapes through Digital
Technologies, Archaeopress, Oxford, pp. 15–28.
Carlson, J. (ed.) (2007). Pilgrimage and Ritual Landscape in Pre-Columbian America, University of
Texas Press, Austin (in press).
Carter, J. C. (ed.) (2001). The Study of Ancient Territories: Chersonesos and Metaponto, Institute of
Classical Archaeology, University of Texas, Austin.
Challis, K., Priestnall, G., Gardner, A., Henderson, J., and O’Hara, S. (2004). Corona remotely-sensed
imagery in dryland archaeology: The Islamic city of al-Raqqa, Syria. Journal of Field Archaeology
29: 139–153.
Chapman, H., and Gearey, B. (2000). Palaeoecology and the perception of prehistoric landscapes: Some
comments on visual approaches to phenomenology. Antiquity 74: 316–319.
Cherry, J. F. (2003). Archaeology beyond the site: Regional survey and its future. In Papadopoulos, J. K.,
and Leventhal, R. M. (eds.), Theory and Practice in Mediterranean Archaeology, Cotsen Institute of
Archaeology, University of California, Los Angeles, pp. 137–160.
Clark, J. J. (2001). Tracking Prehistoric Migrations: Pueblo Settlers Among the Tonto Basin Hohokam,
University of Arizona Press, Tucson.
Cochrane, E. E., and Neff, H. (2006). Investigating compositional diversity among Fijian ceramics with
laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-MS): Implications for
interaction studies on geologically similar islands. Journal of Archaeological Science 33: 378–390.
Conard, N. J. (ed.) (2001). Settlement Dynamics of the Middle Paleolithic and Middle Stone Age, Kerns
Verlag, Tübingen.
Cooney, G. (2000). Landscapes of Neolithic Ireland, Routledge, London.
Cunningham, T. (2001). Variations on a theme: Divergence in settlement patterns, spatial organization in
the Far East of Crete during the Proto- and Neopalatial periods. In Branigan, K. (ed.), Urbanism in
the Aegean Bronze Age, Sheffield Centre for Aegean Archaeology, University of Sheffield,
Sheffield, pp. 72–86.
Daniel, I. R. J. (2001). Stone raw material availability and Early Archaic settlement in the southeastern
United States. American Antiquity 66: 237–265.

123
J Archaeol Res (2008) 16:37–81 75

Derry, L. (2000). Southern town plans, story telling, and historical archaeology. In Young, A. L. (ed.),
Archaeology of Southern Urban Landscapes, University of Alabama Press, Tuscaloosa, pp. 14–29.
Dickinson, W. R., and Shutler, R. Jr. (2000). Implications of petrographic temper analysis for Oceania
prehistory. Journal of World Prehistory 14: 203–266.
Doelle, W. H. (2000). Tonto Basin demography in a regional perspective. In Dean, J. S. (ed.), Salado,
Amerind Foundation, Dragoon, AZ, and University of New Mexico Press, Albuquerque, pp. 81–105.
Doolittle, W. E. (2000). Cultivated Landscapes of Native North America, Oxford University Press,
Oxford.
Dorner, B., Lertzman, and K., Fall, J. (2002). Landscape pattern in topographically complex landscapes:
Issues and techniques for analysis. Landscape Ecology 17: 729–743.
Drennan, R. D., and Peterson, C. E. (2006). Early chiefdom communities compared: The settlement
pattern record for Chifeng, the Alto Magdalena, and the Valley of Oaxaca. In Blanton, R. E. (ed.),
Settlement, Subsistence, and Social Complexity: Essays Honoring the Legacy of Jeffrey R. Parsons,
Cotsen Institute of Archaeology, University of California, Los Angeles, pp. 119–154.
Driessen, J. (2001). History and hierarchy: Preliminary observations on the settlement pattern in Minoan
Crete. In Branigan, K. (ed.), Urbanism in the Aegean Bronze Age, Sheffield Centre for Aegean
Archaeology, University of Sheffield, Sheffield, pp. 51–71.
Duff, A. I. (2000). Scale, interaction, and regional analysis in late Pueblo prehistory. In Hegmon, M. (ed.),
The Archaeology of Regional Interaction: Religion, Warfare, and Exchange across the American
Southwest, University of Colorado Press, Boulder, pp. 71–98.
Duff, A. (2002). Western Pueblo Identities: Regional Interaction, Migration, and Transformation,
University of Arizona Press, Tucson.
Duff, A. I., and Wilshusen, R. H. (2000). Prehistoric population dynamics in the northern San Juan region,
A.D. 950–1300. Kiva 66: 167–190.
Eerkens, J. W., and Rosenthal, J. S. (2004). Are obsidian subsources meaningful units of analysis?:
Temporal and spatial patterning of subsources in the Coso Volcanic Field, southeastern California.
Journal of Archaeological Science 31: 21–29.
Emerson, T. E., Hughes, R. E., Hynes, M. R., and Wisseman, S. U. (2002). Implications of sourcing the
Cahokia-style flint clay figures in the American Bottom and the upper Mississippi River valley.
Midcontinental Journal of Archaeology 27: 309–338.
Erickson, C. L. (2000). The Lake Titicaca basin: A Precolumbian built landscape. In Lentz, D. L. (ed.),
Imperfect Balance: Landscape Transformations in the Pre-Columbian Americas, Columbia
University Press, New York, pp. 311–357.
Erickson, C. L. (2003). Agricultural landscapes as world heritage: Raised field agriculture in Bolivia,
Peru. In Teutonico, J. M., and Matero, F. (eds.), Managing Change: Sustainable Approaches to the
Conservation of the Built Environment, Getty Conservation Institute, Los Angeles, pp. 181–204.
Faught, M. K. (2007). Evidence for fluted point Paleoindian and Early Archaic settlement continuity in
peninsular Florida. In Huckell, B. (ed.), The Land and the People: Explorations of Late Pleistocene,
Early Holocene Human and Environmental History in North America, University of Arizona Press,
Tucson (in press).
Field, J. S. (2002). GIS-based analyses of agricultural production and habitation in the Sigatoka Valley,
Fiji. In Ladefoged, T. N., and Graves, M. (eds.), Pacific Landscape: Archaeological Approaches,
Easter Island Foundation Press, Los Osos, CA, pp. 97–124.
Fisher, C. T., Pollard, H. P., Irade-Alcántara, I., Gardño-Monroy, V. H., and Banerjee, S. (2003). A
reexamination of human-induced environmental change within the Lake Patzcuaro Basin,
Michoacan, Mexico. Proceedings of the National Academy of Sciences 100: 4957–4962.
Forte, M., and Kay, S. (2003). Remote sensing, GIS and virtual reconstruction of archaeological
landscapes. In Forte, M., and Williams, P. R. (eds.), The Reconstruction of Archaeological
Landscapes through Digital Technologies, Archaeopress, Oxford, pp. 109–116.
Forte, M., and Williams, P. R. (eds.) (2003). The Reconstruction of Archaeological Landscapes through
Digital Technologies, Archaeopress, Oxford.
Fyfe, R. M. (2006). GIS and the application of a model of pollen deposition and dispersal: A new
approach to testing landscape hypotheses using the POLLANDCAL models. Journal of
Archaeological Science 33: 483–493.
Fyfe, R. M., Brown, A. G., and Rippon, S. J. (2004). Characterising the late prehistoric, ‘‘Romano-
British’’ and medieval landscape, and dating the emergence of a regionally distinct agricultural
system in South West Britain. Journal of Archaeological Research 31: 1699–1714.

123
76 J Archaeol Res (2008) 16:37–81

Goldstein, P. S. (2000). Communities without borders—The vertical archipelago and diaspora


communities in the southern Andes. In Yaeger, J., and Canuto, M. (eds.), The Archaeology of
Communities: A New World Perspective, Routledge Press, New York, pp. 182–209.
Goldstein, P. S. (2000). Exotic goods and everyday chiefs: Long-distance exchange and indigenous
sociopolitical development in the south-central Andes. Latin American Antiquity 11: 335–362.
Greene, S. E. (2002). Sacred Sites and the Colonial Encounter: A History of Meaning and Memory and
Meaning in Ghana, Indiana University Press, Bloomington.
Grier, C. (2003). Dimensions of regional interaction in the prehistoric Gulf of Georgia. In Matson, R. G.,
Coupland, G., and Mackie, Q. (eds.), Emerging from the Mist: Studies in Northwest Coast Culture
History, University of British Columbia Press, Vancouver, pp. 170–187.
Hare, G. P., Greer, A., Gotthardt, R., Farnell, R., Bowyer, V., and Schweger, C. (2004). Ethnographic and
archaeological investigations of alpine ice patches in southwest Yukon, Canada. Arctic 57: 260–272.
Harrower, M., McCorriston, J., and Oches, E. A. (2002). Mapping the roots of agriculture in southern
Arabia: The application of satellite remote sensing, global position system, and geographic
information system technologies. Archaeological Prospection 9: 35–42.
Hawkins, A. L., Stewart, S. T., and Banning, E. B. (2003). Interobserver bias in enumerated data from
archaeological survey. Journal of Archaeological Science 30: 1503–1512.
Hazelwood, L., and Steele, J. (2004). Spatial dynamics of human dispersals: Constraints on modelling and
archaeological validation. Journal of Archaeological Science 31: 669–679.
Hegmon, M. (ed.) (2000). The Archaeology of Regional Interaction: Religion, Warfare, and Exchange
Across the American Southwest and Beyond, University of Colorado Press, Boulder.
Hegmon, M. M., Nelson, M. C., and Ennes, M. J. (2000). Corrugated pottery, technological style, and
population movement in the Mimbres region of the American Southwest. Journal of Anthropo-
logical Research 56: 217–240.
Herr, S. A., and Clark, J. J. (2002). Mobility and the organization of Prehispanic Southwest communities.
In Parkinson, W. E. (ed.), The Archaeology of Tribal Societies, International Monographs in
Prehistory, Ann Arbor, MI, pp. 123–154.
Hicks, R. (2002). Ways of inhabiting the world: Landscape archaeology. American Anthropologist 104:
315–339.
Hockett, B. (2005). Middle and Late Holocene hunting in the Great Basin: A critical review of the debate
and future prospects. American Antiquity 70: 713–731.
Hoffecker, J. F. (2002). Desolate Landscapes: Ice Age Settlement in Eastern Europe, Rutgers University
Press, New Brunswick, NJ.
Holcomb, D. W. (2001). Imaging radar and archaeological survey: An example from the Gobi Desert of
southern Mongolia. Journal of Field Archaeology 28: 131–141.
Horning, A. (2001). Of saints and sinners: Mythic landscapes of the Old and New South. In Shackel, P. A.
(ed.), Myth, Memory, and the Making of the American Landscape, University Press of Florida,
Gainesville, pp. 21–46.
Houston, S., Escobedo, H., Child, M., Golden, C., and Muñoz, R. (2003). The moral community: Maya
settlement transformation at Piedras Negras, Guatemala. In Smith, M. L. (ed.), The Social
Construction of Ancient Cities, Smithsonian Institution Press, Washington, DC, pp. 254–268.
Huntley, D. L., and Kintigh, K. W. (2004). Archaeological patterning and organizational scale of late
prehistoric settlement clusters in the Zuni region of New Mexico. In Adams, E. C., and Duff, A. I. (eds.),
The Protohistoric Pueblo World: A.D. 1275–1600, University of Arizona Press, Tucson, pp. 62–74.
Huysecom, E. (2002). Palaeoenvironment and human population in West Africa: An international
research project in Mali. Antiquity 76: 335–336.
Huysecom, E., Ozainne, S., Raeli, F., Ballouche, A., Rasse, M., and Stokes, S. (2004). Ounjougou (Mali):
A history of Holocene settlement at the southern edge of the Sahara. Antiquity 78: 579–593.
Isbell, W. H. (2000). What we should be studying: The ‘‘imagined community’’ and the ‘‘natural
community.’’ In Canuto, M. A., and Yaeger, J. (eds.), The Archaeology of Communities: A New
World Perspective, Routledge, London, pp. 243–366.
Janusek, J. W. (2002). Out of many, one: Style and social boundaries in Tiwanaku. Latin American
Antiquity 13: 35–61.
Jones, G. (2005). Garden cultivation of staple crops and its implications for settlement location and
continuity. World Archaeology 37: 164–176.
Keightley, D. N. (2000). The Ancestral Landscape: Time, Space, and Community in Late Shang China
(ca. 1200–1045 B.C.), Institute of East Asian Studies, University of California, Berkeley.

123
J Archaeol Res (2008) 16:37–81 77

Knudson, K. J., Price, T. D., Buikstra, J. E., and Blom, D. E. (2004). The use of strontium isotope
analyses to investigate Tiwanaku migration and mortuary ritual in Bolivia and Peru. Archaeometry
46: 5–18.
Kolata, A. L. (ed.) (2003). Tiwanaku and Its Hinterland: Archaeology and Paleoecology of an Andean
Civilization, Smithsonian Institution Press, Washington, DC.
Kolb, M. J., and Dixon, B. (2002). Landscapes of war: Rules and conventions of conflict in ancient
Hawai’i (and elsewhere). American Antiquity 67: 514–534.
Kolb, M. J., and Speakman, R. J. (2005). Elymian regional interaction in Iron Age western Sicily: A
preliminary neutron activation study of incised/impressed tablewares. Journal of Archaeological
Science 32: 795–804.
Koontz, R., Reese-Tayler, K., and Headrick, A. (eds.) (2001). Landscape and Power in Ancient
Mesoamerica, Westview Press, Boulder, CO.
Kowalewski, S. A. (2003). What is a community? The long view from Oaxaca, Mexico. Social Evolution
and History 2: 4–34.
Krakker, J. J. (2001). The physical-environmental context of Powers phase settlements. In O’Brien, M. J.
(ed.), Mississippian Community Organization: The Powers Phase in Southeastern Missouri, Kluwer
Academic/Plenum Publishers, New York, pp. 56–76.
Ladefoged, T. L., and Graves, M. W. (eds.) (2002). Pacific Landscapes: Archaeological Approaches to
Oceania, Easter Island Foundation, Los Osos, CA.
Ladefoged, T. L., Graves, M. W., and McCoy, M. D. (2003). Archaeological evidence for agricultural
development in Kohala, Island of Hawaii. Journal of Archaeological Science 30: 923–940.
Lahr, M., Foley, R., and Pinhasi, R. (2000). Expected regional patterns of Mesolithic-Neolithic human
population admixture in Europe based on archaeological evidence. In Renfrew, C., and Boyle, K.
(eds.), Archaeogenetics: DNA and the Population Prehistory of Europe, McDonald Institute,
Cambridge, UK, pp. 81–88.
Lake, M. W. (2000). MAGICAL computer simulation of Mesolithic foraging. In Kohler, T. A., and
Gumerman, G. J. (eds.), Dynamics in Human Societies, Oxford University Press, Oxford, pp.
107–143.
Lentz, D. L. (ed.) (2000). Imperfect Balance: Landscape Transformations in the Pre-Columbian
Americas, Columbia University Press, New York.
Lepofsky, D., Lertzman, K., Hallett, D., and Mathews, R. (2005). Climate change and culture change on
the southern coast of British Columbia 2400–1200 Cal. B.P.: An hypothesis. American Antiquity 70:
267–293.
Levi, L. J. (2002). An institutional perspective on prehispanic Maya residential variation: Settlement and
community at San Estevan, Belize. Journal of Anthropological Archaeology 21: 120–141.
Levy, T. E., and Holl, F. C. (2002). Migrations, ethnogenesis, and settlement dynamics: Israelites in Iron
Age Canaan and Shuwa-Arabs in the Chad Basin. Journal of Anthropological Archaeology 21:
83–118.
Liu, L., Chen, X., Lee, Y. K., Wright, H. T., and Miller-Rosen, A. (2004). Settlement patterns and
development of social complexity in the Yiluo region, North China. Journal of Field Archaeology
29: 1–26.
Llobera, M. (2001). Building past landscape perception with GIS: Understanding topographic
prominence. Journal of Archaeological Science 28: 1005–1014.
Lloyd, C. D., and Atkinson, P. M. (2004). Archaeology and geostatistics. Journal of Archaeological
Science 31: 151–165.
Louwagie, G., Stevenson, C. M., and Langohr, R. (2006). The impact of moderate to marginal land
suitability on prehistoric agricultural production and models of adaptive strategies for Easter Island
(Rapa Nui, Chile). Journal of Anthropological Archaeology 25: 290–317.
Lovis, W. A., Donahue, R. E., and Holman, M. B. (2005). Long-distance logistic mobility as an
organizing principle among northern hunter-gatherers: A Great Lakes middle Holocene settlement
system. American Antiquity 70: 669–693.
Lucero, L. J. (2003). The politics of ritual: The emergence of Classic Maya rulers. Current Anthropology
44: 523–558.
Lycett, M. (2001). Transformation of place: Occupational history and differential persistence in
seventeenth century New Mexico. In Preucell, R. W. (ed.), Archaeologies of the Pueblo Revolt:
Identity, Meaning, and Renewal in the Pueblo World, University of New Mexico Press,
Albuquerque, pp. 61–74.

123
78 J Archaeol Res (2008) 16:37–81

Lyons, P. D. (2003). Ancestral Hopi Migrations, Anthropological Paper No. 68, University of Arizona
Press, Tucson.
Malim, T. (2000). The ritual landscape of the Neolithic and Bronze Age along the middle and lower Ouse
Valley. In Dawson, M. (ed.), Prehistoric, Roman and Post-Roman Landscapes of the Great Ouse
Valley, Council for British Archaeology, York, pp. 57–88.
Malim, T. (2001). Place and space in the Cambridgeshire Bronze Age. In Brück, J. (ed.), Bronze Age
Landscapes: Tradition and Transformation, Oxbow Books, Oxford, pp. 9–22.
Malville, N. J. (2001). Long-distance transport of bulk goods in the pre-Hispanic American Southwest.
Journal of Anthropological Archaeology 20: 230–243.
Maxham, M. D. (2000). Rural communities in the Black Warrior Valley, Alabama: The role of
commoners in the creation of the Moundville I landscape. American Antiquity 65: 337–354.
Mayor, A., Huysecom, E., Gallay, A., Rasse, M., and Ballouche, A. (2005). Population dynamics and the
paleoclimate over the past 3000 years in the Dogon Country, Mali. Journal of Anthropological
Archaeology 24: 25–61.
McCorriston, J., and Weisberg, S. (2002). Spatial and temporal variation in Mesopotamian agricultural
practices in the Khabur Basin, Syrian Jazira. Journal of Archaeological Science 29: 485–498.
Millett, M. (2001). Roman interaction in northwestern Iberia. Oxford Journal of Archaeology 20:
157–170.
Neff, H. (2003). Analysis of Mesoamerican plumbate pottery surfaces by laser ablation-inductively
plasma-mass spectrometry (LA-ICP-MS). Journal of Archaeological Science 30: 21–35.
Nelson, M. C. (2000). Abandonment: Conceptualization, representation, and social change. In Schiffer,
M. B. (ed.), Social Theory in Archaeology, University of Utah Press, Salt Lake City, pp. 52–62.
Nelson, M. C., and Schachner, G. (2002). Understanding abandonments in the North American
Southwest. Journal of Archaeological Research 10: 167–206.
Niven, L. B., Egeland, C. P., and Todd. L. C. (2004). An inter-site comparison of enamel hypoplasia in
bison: Implications for paleoecology and modeling Late Plains Archaic subsistence. Journal of
Archaeological Science 31: 1783–1794.
Nocete, F., Alex, E., Nieto, J. M., Sáez, R., and Bayona, M. R. (2005). An archaeological approach to
regional environmental pollution in the south-western Iberian Peninsula related to third millennium
BC mining and metallurgy. Journal of Archaeological Science 32: 1566–1576.
Nocete, F., Saez, R., Nieto, J. M., Cruz-Aunon, R., Cabrero, R., Alex, E., and Bayona, M. R. (2005).
Circulation of silicified oolitic limestone blades in South-Iberia (Spain and Portugal) during the third
millennium B.C.: An expression of a core/periphery framework. Journal of Anthropological
Archaeology 24: 62–81.
O’Brien, M. J. (ed.) (2001). Mississippian Community Organization: The Powers Phase in Southeastern
Missouri, Kluwer Academic/Plenum Publishers, New York.
Ogundiran, A. O. (2001). Ceramic spheres and regional networks in the Yoruba-Edo region, Nigeria,
13th–19th centuries A.C. Journal of Field Archaeology, 28: 27–43.
Pauketat, T. R. (2003). Resettled farmers and the making of a Mississippian polity. American Antiquity
68: 39–66.
Peterson, R. (2003). Settlements, Kinship, and Hunting Grounds in Traditional Greenland: A
Comparative Study of Local Experiences from Upernavik and Ammassalik, Man and Society No.
27, Meddelelser om Grønland, Copenhagen.
Pettegrew, D. K. (2001). Chasing the Classical farmstead: Assessing the formation and signature of rural
settlement in Greek landscape archaeology. Journal of Mediterranean Archaeology 14: 189–209.
Pickering, M. (2003). Modelling Hunter-Gatherer Settlement Patterns. An Australian Case Study,
Archaeopress, Oxford.
Pinhasi, R., Foley, R., and Lahr, M. (2000). Spatial and temporal patterns in the Mesolithic-Neolithic
archaeological record of Europe. In Renfrew, C., and Boyle, K. (eds.), Archaeogenetics: DNA and
the Population Prehistory of Europe, McDonald Institute, Cambridge, UK, pp. 45–56.
Pool, C. A. (2006). Current research on the Gulf Coast of Mexico. Journal of Archaeological Research
14: 189–241.
Possehl, G. L. (2000). The drying up of the Sarasvati: Environmental disruption in South Asian
prehistory. In Bawden, G., and Reycraft, R. M. (eds.), Environmental Disaster and the Archaeology
of Human Response, Maxwell Museum of Anthropology, Albuquerque, NM, pp. 63–74.
Premo, L. S. (2004). Local spatial autocorrelation statistics quantify multi-scale patterns in distributional
data: An example from the Maya Lowlands. Journal of Archaeological Science 31: 855–866.

123
J Archaeol Res (2008) 16:37–81 79

Price, T. D., Bentley, R. A., Gronenborn, D., Lüning, J., and Wahl, J. (2001). Human migration in the
Linearbandkeramik of central Europe. Antiquity 75: 593–603.
Quinn, R., Dean, M., Lawrence, M., Liscoe, S., and Boland, D. (2005). Backscatter responses and
resolution considerations in archaeological side-scan sonar surveys: A control experiment. Journal
of Archaeological Science 32: 1252–1264.
Ravesloot, J. C., and Waters, M. R. (2004). Geoarchaeology and archaeological site patterning on the
Middle Gila River, Arizona. Journal of Field Archaeology 29: 203–214.
Reinhard, J., and Ceruti, M. C. (2007). Pilgrimage, sacred mountains, and human sacrifice among the
Inca. In Carlson, J. (ed.), Pilgrimage and Ritual Landscape in Pre-Columbian America, University
of Texas Press, Austin (in press).
Richter, J. (2001). For lack of a wise old man? Late Neanderthal land-use patterns in the Altmül River
valley, Bavaria. In Conard, N. J. (ed.), Settlement Dynamics of the Middle Paleolithic and Middle
Stone Age, Kerns Verlag, Tübingen, pp. 205–219.
Roberts, B. K., and Wrathmall, S. (2002). Region and Place: A Study of English Rural Settlement, English
Heritage Publications, Oxford.
Rolland, N. (2001). Determinants of Middle Paleolithic settlement organization: A review of evidence,
based on the record from Western Europe. In Conard, N. J. (ed.), Settlement Dynamics of the Middle
Paleolithic and Middle Stone Age, Kerns Verlag, Tübingen, pp. 545–571.
Rosenswig, R. M. (2000). Some political processes of ranked societies. Journal of Anthropological
Archaeology 19: 413–460.
Ruby, B., Carr, C., and Charles, D. (2005). Community organization in the Scioto, Mann, and Havan
Hopewell Regions: A comparative perspective. In Carr, C., and Case, D. T. (eds.), Gathering
Hopewell: Society, Ritual, and Ritual Interaction, Plenum, New York.
Sanders, W. (1999). Twenty-five years of settlement archaeology in Mesoamerica. In Billman, B., and
Feinman, G. (eds.), Settlement Pattern Studies in the Americas: Fifty Years Since Virú, Smithsonian
Institution Press, Washington, DC, pp. 12–21.
Santley, R. S., Nelson, S. A., Reinhardt, B. K., Pool, C. A., and Arnold, P. J., III. (2000). When day turned
to night: Volcanism and the archaeological record from the Tuxtla Mountains, southern Veracruz,
Mexico. In Bawden, G., and Reycraft, R. M. (eds.), Environmental Disaster and the Archaeology of
Human Response, Maxwell Museum of Anthropology, Albuquerque, NM, pp. 143–162.
Sarris, A., and Jones, R. E. (2000). Geophysical and related techniques applied to archaeological survey
in the Mediterranean: A review. Journal of Mediterranean Archaeology 13: 3–75.
Sarris, A., Galaty, M. L., Yerkes, R. W., Parkinson, W. A., Gyucha, A., Billingsley, D., and Tate, R.
(2004). Geophysical prospection and soil chemistry at the Early Copper Age settlement of Vésztõ-
Bikeri, southeastern Hungary. Journal of Archaeological Science 31: 927–939.
Scarborough, V. (2003). The Flow of Power: Ancient Water Systems and Landscapes, School of
American Research Press, Santa Fe, NM.
Scarborough, V. (2003). How to interpret an ancient landscape. Proceedings of the National Academy of
Sciences 100: 4366–4368.
Scarre, C. (2000). Forms and landforms: The design and setting of Neolithic monuments in western
France. In Ritchie, A. (ed.), Neolithic Orkney in its European Context, McDonald Institute for
Archaeological Research, Cambridge, UK, pp. 309–320.
Schaepe, D. M. (2006). Rock fortifications: archaeological insights into pre-contact warfare and
sociopolitical organization among the Stólo of the lower Fraser River Canyon, B.C. American
Antiquity 71: 671–706.
Schroeder, S. (2000). Settlement patterns and cultural ecology in the southern American Bottom. In
Ahler, S. R. (eds.), Mounds, Modoc, and Mesoamerica: Papers in Honor of Melvin L. Fowler,
Scientific Papers Vol. 28, Illinois State Museum, Springfield, pp. 179–191.
Schuldenrein, J., Wright, R. P., Mughal, M. R., and Khan, M. A. (2004). Landscapes, soils, and mound
histories of the Upper Indus Valley, Pakistan: New insights on the Holocene environments near
ancient Harappa. Journal of Archaeological Science 31: 777–797.
Sluyter, A. (2002). Colonialism and Landscape: Postcolonial Theory and Applications, Rowman and
Littlefield, Lanham, MD.
Smith, A. T. (2003). The Political Landscape: Constellations of Authority in Early Complex Societies,
University of California Press, Berkeley.
Snead, J. E. (2002). Ancestral Pueblo trails and the cultural landscape of the Pajarito Plateau, New
Mexico. Antiquity 76: 756–765.

123
80 J Archaeol Res (2008) 16:37–81

Spikens, P. (2000). GIS models of past vegetation: An example from northern England, 10,000–5000 B.P.
Journal of Archaeological Science 27: 219–234.
Spikens, P., Conneller, C., Ayestaran, H., and Scaife, B. (2002). GIS based interpolation applied to
distinguishing occupation phases of early prehistoric sites. Journal of Archaeological Science 29:
1235–1245.
Stanish, C., and Bauer, B. (2007). Pilgrimage and geography of power in the Inca state. In Morris, C.,
Burger, R., and Matos, R. (eds.), Variability in the Expressions of Inka Power, Dumbarton Oaks,
Washington, DC (in press).
Stoffle, R. W., Loendorf, L., Austin, D., Halmo, D., and Bulletts. A. (2000). Ghost dancing the Grand Canyon:
Southern Paiute rock art, ceremony, and cultural landscapes. Current Anthropology 41: 11–38.
Stoltman, J. B. (2000). A reconsideration of the cultural processes linking Cahokia to its northern
hinterlands during the period A.D. 1000–1200. In Ahler, S. R. (ed.), Mounds, Modoc, and
Mesoamerica: Papers in Honor of Melvin L. Fowler, Scientific Papers Vol. 28, Illinois State
Museum, Springfield, pp. 439–467.
Sutton, S. B. (2000). Contingent Countryside: Settlement, Economy, and Land Use in the Southern
Argolid since 1700, Stanford University Press, Stanford, CA.
Thurston, T. L. (2001). Landscapes of Power, Landscapes of Conflict: State Formation in the Danish Iron
Age, Kluwer Academic/Plenum Publishing, New York.
Tringham, R. (2000). Southeastern Europe in the transition to agriculture in Europe: Bridge, buffer, or
mosaic. In Price, T. D. (ed.), Europe’s First Farmers, Cambridge University Press, Cambridge, UK,
pp. 19–56.
Tschan, A. P., Raczkowski, W., and Latalowa, M. (2000). Perception and viewsheds: Are they mutually
inclusive? In Lock, G. (ed.), Beyond the Map: Archaeology and Spatial Technologies, IOS Press,
Amsterdam, pp. 28–48.
Ucko, P. J., and Layton, R. (eds.) (1999). The Archaeology and Anthropology of Landscapes: Shaping
Your Landscape, Routledge, New York.
Usman, A. A. (2004). On the frontier of empire: Understanding the enclosed walls in northern Yoruba,
Nigeria. Journal of Anthropological Archaeology 23: 119–132.
van der Leeuw, S. E. (2000). Land degradation as a socionatural process. In McIntosh, R., Tainter, J., and
McIntosh, S. K. (eds.), The Way the Wind Blows: Climate, History, and Human Action, University
of Columbia Press, New York.
Van de Velde, P. (2001). An extensive alternative to intensive survey: Point sampling in the Riu Mannu
Survey Project, Sardinia. Journal of Mediterranean Archaeology 14: 24–52.
VanPool, T. L., VanPool, C. S., Cruz-Antillón, R., Leonard, R. D., and Harmon, M. J. (2000). Flaked
stone and social interaction in the Casas Grandes region, Chihuahua, Mexico. Latin American
Antiquity 11: 163–174.
Vermeersch, P. (2001). Middle Paleolithic settlement patterns in West European open-air sites:
Possibilities, problems. In Conard, N. J. (ed.), Settlement Dynamics of the Middle Paleolithic and
Middle Stone Age, Kerns Verlag, Tübingen, pp. 251–263.
Whalen, M. E., and Minnis, P. E. (2001). Casas Grandes and Its Hinterland: Prehistoric Regional
Organization in Northwest Mexico, University of Arizona Press, Tucson.
Wheatley, D., and Gillings, M. (2000). Vision, perception, and GIS: Developing enriched approaches to
the study of archaeological visibility. In Lock, G. (ed.), Beyond the Map: Archaeology and Spatial
Technologies, IOS Press, Amsterdam, pp. 1–27.
Whittle, A. (2002). The coming of agriculture: People, landscapes, and change c. 4000–1500 B.C. In
Slack, P., and Ward, R. (eds.), The Peopling of Britain: The Shaping of a Human Landscape. The
Linacre Lectures, 1999, Oxford University Press, Oxford, pp. 77–109.
Whitmore, T. M., and Turner, B. L. II. (2001). Cultivated Landscapes of Middle America on the Eve of
Conquest, Oxford University Press, Oxford.
Wilkinson, T. J. (2000). Regional approaches to Mesopotamian archaeology: The contribution of
archaeological surveys. Journal of Archaeological Research 8: 219–267.
Wisseman, S. U., Moore, D. M., Hughes, R. E., Hynes, M. R., and Emerson, T. E. (2002). Mineralogical
approaches to sourcing Eastern Woodlands pipes and figurines. Geoarchaeology 17: 689–715.
Yacobaccio, H. D., Escola, P. S., Pereyra, F. X., Lazzari, M., and Glascock, M. (2004). Quest for ancient
routes: Obsidian sourcing research in Northwestern Argentina. Journal of Archaeological Science
31: 193–204.

123
J Archaeol Res (2008) 16:37–81 81

Zarins, J. (2000). Environmental disruption and human response: An archaeological-historical example


from South Arabia. In Bawden, G., and Reycraft, R. M. (eds.), Environmental Disaster and the
Archaeology of Human Response, Maxwell Museum of Anthropology, Albuquerque, NM, pp.
35–49.
Zedeño, M. N. (2000). On what people make of places: A behavioral cartography. In Schiffer, M. B. (ed.),
Social Theory in Archaeology, University of Utah Press, Salt Lake City, pp. 97–111.

123

Potrebbero piacerti anche