You are on page 1of 53

Is it Intensification Yet?

Current
Archaeological Perspectives on the
Evolution of Hunter-Gatherer Economies

Christopher Morgan

Journal of Archaeological Research

ISSN 1059-0161

J Archaeol Res
DOI 10.1007/s10814-014-9079-3

1 23
Your article is protected by copyright and all
rights are held exclusively by Springer Science
+Business Media New York. This e-offprint is
for personal use only and shall not be self-
archived in electronic repositories. If you wish
to self-archive your article, please use the
accepted manuscript version for posting on
your own website. You may further deposit
the accepted manuscript version in any
repository, provided it is only made publicly
available 12 months after official publication
or later and provided acknowledgement is
given to the original source of publication
and a link is inserted to the published article
on Springer's website. The link must be
accompanied by the following text: "The final
publication is available at link.springer.com”.

1 23
Author's personal copy
J Archaeol Res
DOI 10.1007/s10814-014-9079-3

Is it Intensification Yet? Current Archaeological


Perspectives on the Evolution of Hunter-Gatherer
Economies

Christopher Morgan

 Springer Science+Business Media New York 2014

Abstract Originally designed to explain causes of increased productivity in


agricultural systems, the concept of intensification has become widely linked to
hunter-gatherer archaeology. Worldwide, recent applications show that progress has
been made with regard to recognizing, describing, and modeling the declining
foraging efficiency predicted by traditional intensification models that take into
account confounding factors like taphonomy, environmental change, and differen-
tial foraging goals. Less progress has been made in explaining intensification due to
problems of identifying primacy in the environmental, demographic, technological,
and social mechanisms that lead to increased production. These problems are
confounded by imprecise usage of the concept ‘‘intensification,’’ which runs the
gamut from behaviors that either increase or decrease efficiency as the means of
increasing productivity. Resolving these problems hinges on unpacking the very
concept of intensification as currently applied to hunter-gatherer archaeology. This
requires much greater specificity with regard to efficiency and adherence to a
Boserupian perspective that declining efficiency marks intensification processes.
Alternative modes of increasing production that do not necessarily entail declining
efficiency—specialization, diversification, and innovation—also must be taken into
account to explain the evolution of hunter-gatherer economies.

Keywords Intensification  Hunter-gatherer  Efficiency  Cultural evolution

C. Morgan (&)
Department of Anthropology, University of Nevada, Reno, 1664 N. Virginia St, Reno,
NV 89557-0096, USA
e-mail: ctmorgan@unr.edu

123
Author's personal copy
J Archaeol Res

Introduction

Despite its initial development as a means of explaining the drivers of increased


agricultural production in terms of resource-population dynamics (Boserup 1965),
the concept of intensification has become widely linked to hunter-gatherer research,
used to explain everything from the evolution of modern human behavior (Manne
and Bicho 2009), to the origin of economies prefiguring the inception of farming
(Henry 1985), to the development of ranking in complex hunter-gatherer societies
(Fitzhugh 2003b). This is perhaps not surprising given the common perception that
what characterizes macroscale changes in human economies over the last
100,000 years or so is increasing diversification, technological sophistication, and
more intensive landscape and resource use (Marlowe 2005). This idea is echoed in
the sentiments of a wide range of research that sees the long-term trajectory of
human cultural evolution, most of it occurring within the realm of hunter-gatherer
economic systems, as one of increase on economic, demographic, and even social
scales (Ellis et al. 2013; Holling 2001; Kirch 2005; Redding 1988). This has resulted
in a vast corpus of literature on the archaeology of hunter-gatherers that explicitly
uses the concept of intensification in its analytical framework and an even larger
amount of research that implicates intensification processes in cultural evolution and
culture change. In short, the term and its related concepts have become nearly
ubiquitous in the archaeological literature on hunter-gatherers. This has led to
diverse and even contradictory applications of the concept of intensification, to the
point that a critical assessment of its current methodological and theoretical
applications to the archaeology of hunter-gatherers is overdue.
In light of this, I provide a critical review of intensification as the concept has
been applied to global hunter-gatherer archaeological research over about the last
decade, since the publication of Binford’s (2001) seminal treatise on hunter-gatherer
ecology and Smith’s (2001) essential contribution toward understanding the
continuum operating between intensive foraging and farming economies. This
research comprises a compendious literature. As of this writing, a search on the
Anthropology Plus database for titles that include the term ‘‘intensification’’
published between 2002 and 2013 produces 118 sources, nearly all of them related
to hunter-gatherers. Increasing search parameters to include words like ‘‘intense,’’
‘‘intensive,’’ ‘‘foraging efficiency,’’ or ‘‘population pressure’’ raises the total to 286
records. A similar search on Google Scholar using ‘‘intensification,’’ ‘‘hunter
gatherer,’’ and ‘‘archaeology’’ for this same time period produces 8580 results, a
volume of literature clearly beyond the scope of all but the most insomniac of
readers. Therefore, I sample this literature in seven main categories where the
concept of intensification has been widely appropriated both historically and in
current usage to hunter-gatherer archaeology. I developed this sample to provide a
representative snapshot rather than a comprehensive review of the current status of
applications of intensification to hunter-gatherer archaeology in each of these
subcategories, many of which have been subject to recent review articles of their
own. These categories are (1) Upper Paleolithic, (2) Terminal Pleistocene-Early
Holocene (TPEH), (3) global coastal settings and Oceania, and (4) North America,
broadly defined. Two settings in North America where the literature on hunter-

123
Author's personal copy
J Archaeol Res

gather intensification is particularly well established are (5) the northwestern Pacific
coast (hereafter ‘‘Pacific Northwest’’), and (6) California. The final category
examines (7) theoretical and quantitative models (Fig. 1).
I derived three main results from this review. First, the main debates regarding
intensification, as relevant now as ever, are implied in much of the recent literature.
One of the original debates regarding intensification was whether or not increasing
population was required to initiate the process (Cowgill 1975). This subject is either
addressed directly or implied in the diversity of opinion as to whether increasing
population, climate change, or changes in social behavior drove intensification in
the TPEH and especially along the western coast of North America, where long-
running debates over whether abundance or scarcity drove the development of these
societies speaks to this fundamental population-pressure issue. Second, other
debates also are alive and well, with little evidence of resolution. Specifically, there
is a long-running but often overlooked schism regarding the fundamental definition
of intensification. One side of this divide uses intensification in a descriptive sense,
where it is simply taken to mean any increase in economic productivity. The other
uses intensification in an explanatory sense, where it is employed to explain how
people solved population-resource imbalances by the addition of more labor to
increase economic output, resulting in a net decrease in foraging efficiency. As my
review indicates, this is much more than a semantic disagreement; rather it is a
fundamental theoretical divergence with important implications for the roles that
labor, capital, specialization, diversification, and innovation play in the development
of increasingly complex hunter-gatherer economic systems. Lastly, significant
headway has been made over the last decade in terms of how we think about and
model foraging efficiency (much of it based on how to interpret zooarchaeological
data) that shows potential to resolve some of the longer-running theoretical debates
over population pressure and efficiency, the critical components of the intensifi-
cation argument, at least as initially conceived. A synthesis of this material indicates
that there are multiple pathways to increased productivity in hunter-gatherer
economies—specialization, diversification, intensification, and innovation—each
entailing differential effects on efficiency, a critical and at times overlooked
component of intensification research since the concept was first elucidated by
Boserup.

A necessarily brief history of intensification and hunter-gatherer archaeology

As originally developed by Boserup (1965, 1970, 1981), intensification entails the


addition of labor (e.g., field preparation and tending, fertilization, irrigation) to
increase the yield of agricultural plots to feed increasingly large populations, which
solves the Malthusian (Malthus 1798) problem of growing populations outstripping
their food supply (Cohen 1977). The long-term stability of intensified modes of
production aside, the model has at its core the assumption that increased production
entails declining efficiency due to the greater labor required to effect this change
(Zvelebil 1986). Boserup’s model and its criteria engendered considerable debate
(e.g., Leach 1999; Morrison et al. 1996; Thurston and Fisher 2007a), hinging in

123
Author's personal copy
J Archaeol Res

North Canada
(N. America)

Mediterranean (Paleolithic)
Mesolithic
Northwest Coast (TPEH)
Northern California Epipaleolithic
East Coast & Woodlands
Northwest China
(N. America)
Hawaii (Paleolithic & TPEH)
(Coast) Texas (N. America)

Southern California/ Papua New Guinea (Coast)


California Bight

Great Basin
(N. America) Australia
(Coast)

South African Coast


New Zealand
Tierra del Fuego (Paleolithic & Coast)
(Coast)
(Coast)

Fig. 1 Map showing locations of studies described in the text

large part on Cowgill’s (1975) criticism that intrinsic rates of population increase
cannot be assumed and that increases in productivity may result by seizing ‘‘new
opportunities’’ (Hassan 1977). Building on this criticism, others saw environmental
characteristics (mainly abundance) and innovations (mainly technological but also
social) as means of increasing productivity without necessarily decreasing
efficiency (Bender 1978, 1981, 1985; Kirch 1994). In the realm of hunter-gatherer
studies, this type of thinking is antedated by the early evolutionary idea that
increasing efficiency characterized the development of Archaic lifeways in the
eastern woodlands (Caldwell 1958) and desert west of North America (Jennings
1968). Debate also centered on whether intensification was a ‘‘top-down’’ approach
driven by elite leadership (e.g., Carneiro 1970; Hayden 2001; Wittfogel 1957) or
whether it was driven from the ‘‘bottom-up’’ by small landholders and householders
who initiated increased productivity (Erickson 1993; Thurston and Fisher 2007b).
These debates clearly prefigure many of those associated with intensification and its
applications to hunter-gatherer archaeology today.
The roots of intensification’s application to hunter-gatherer archaeology began in
earnest with research focusing on the antecedents to plant and animal domestication
in the terminal Pleistocene (Binford 1968; Flannery 1969) and on the recognition of
‘‘complex’’ hunter-gatherers in the late 1970s and 1980s (Price and Brown 1985).
Flannery’s (1969) ‘‘broad spectrum revolution’’ (BSR) model served as a counter to
Braidwood’s (1963) ‘‘settling-in’’ hypothesis for domestication. Flannery argued
that it was the Epipaleolithic budding off of populations into less productive habitats
in the Near East that challenged environmental carrying capacities, forcing such
groups into exploiting a wider array of abundant, predictable, but also smaller-
bodied prey and plants. Critical to this argument is the idea that population pressure
ultimately drove subsistence diversification, if not outright intensification (Earle
1980). In the second case, and counter to the arguments of Lee (1968), was the
acknowledgment that hunter-gatherers in places like the Pacific Northwest,
California, Japan, and Mesolithic Europe stored food, were at least semisedentary,
had higher-than-expected population densities, and sometimes exhibited

123
Author's personal copy
J Archaeol Res

hierarchical social structures (Arnold 1996; Binford 1980; Price and Brown 1985).
This recognition culminated in a fundamental rethinking of hunter-gatherers, in
large part because these behaviors were more commonly thought to be associated
with agriculturalists and, critically, entailed food production that matched or
exceeded that of some agricultural societies (Baumhoff 1963; Smith 2001). Further
reevaluation of the very utility of hunter-gatherers as an analytical type has led to
observations that the only legitimate difference between hunter-gatherers and
farmers was their subsistence base (Morrison 1994; Sassaman 2004; Terrell et al.
2003). The analytical trope has persisted and even flourished in light of this essential
criticism, but the intrinsic linkage between complex hunter-gatherers and intensi-
fication had been made (Arnold 1993).
In light of these developments, collectors, processors, chieftainships, and other
forms of ‘‘intensified,’’ complex hunter-gatherers were identified and added to
global hunter-gatherer research agenda (Bettinger 1991; Kelly 2013). Much of this
work was geared to resolving the Braidwood–Flannery debate by reconstructing
Mesolithic, Epipaleolithic, and Natufian settlement and subsistence strategies in
Europe and the Near East (Bar-Yosef 1998; Bar-Yosef and Belfer-Cohen 1989;
Gould 1985; Jochim 1976), much of it incorporating optimal foraging models.
Many indeed found evidence for increased diet breadth and sedentism at the dawn
of the Holocene (e.g., Henry 1985), but the extent to which this truly represented
intensification (in the Boserupian sense) or even association with the development
of agriculture is currently in doubt as much now as ever (Zeder 2012).
The extension of the concept of intensification into the realm of Holocene
hunter-gatherer archaeology soon developed in places like Australia (Lourandos
and Ross 1994; Smith 2013) but became particularly well established in western
North American (Bettinger 2001). In the Pacific Northwest, for example,
archaeologists used intensification to explain burgeoning populations, intensive
salmon exploitation, storage, and the development of mostly sedentary, hierarchi-
cal societies (Ames 1981, 1994; Croes and Hackenberger 1988; Matson 1992).
Disagreement developed (and continues to this day) as to whether these
phenomena hinged on environmental abundance or population increase, whether
intensive storage economies developed before or after population increases, and on
whether social rather than ecological factors were the prime movers behind these
phenomena (Ames 2003; Fitzhugh 2003b; Hayden 2001; Matson 1992).
Overlooked in these debates was the fact that intensification began to be defined
in an explicitly non-Boserupian sense, meaning it was applied to any means or
evidence for increased production (diversification, specialization, storage, etc.),
whether this entailed declining efficiency or not (Ames 1985; Matson 1983).
Essentially, increased productivity became conflated with increased efficiency in
some of the Northwest Coast literature, in direct contrast to the Boserupian model
(see also Binford 2001, p. 188; Morrison 1994, p. 130).
Similarly, in California archaeology, concepts of intensification were applied
throughout the 1980s and 1990s to explain not only the development of intensive
acorn and small-seed processing in the mid-to-late Holocene (Basgall 1987; Beaton
1991; Bouey 1979, 1987; Wohlgemuth 1996) but also the development of
chiefdoms on the southern California Bight (Arnold 1991, 1992b; Arnold et al.

123
Author's personal copy
J Archaeol Res

1997). The degree to which this was the result of population increase, environmental
decline (both entailing population pressure), or the environmental abundance of
California was never resolved (e.g., see the dichotomy between King 1990 and Raab
and Larson 1997). Also never resolved was the degree to which intensification was a
‘‘top-down’’ process initiated by entrepreneurial chiefs taking advantage of periodic
environmental downturns (Arnold 1992a) or whether it was a ‘‘bottom-up’’
progression affiliated more with household economic decision making in light of
resource stress (Raab and Larson 1997). A substantial portion of research in the
region explicitly used optimal foraging models (i.e., diet breadth/prey choice) to
equate declining prey size with declining foraging efficiency over time and hence
Boserupian intensification, especially in the San Francisco Bay region (Broughton
1994a, b, 1997).
In sum, intensification became linked to hunter-gatherer archaeology mainly
through its potential utility for explaining the BSR and ‘‘complex’’ hunter-
gatherers. In many instances, this linkage came to incorporate aspects of
behavioral ecology (Bird and O’Connell 2006), especially the diet breadth/prey
choice optimal foraging model (MacArthur and Pianka 1966) as a means to
model declining foraging efficiency. The meaning of the term itself became
conflated with both a strict Boserupian definition that entails declining foraging
efficiency (hereafter ‘‘intensification sensu stricto [s.s.]’’) or alternatively as any
means of increasing productivity (e.g., diversification, specialization, innovation),
including those that ostensibly increased efficiency (hereafter ‘‘intensification
sensu lato [s.l.]’’). When applied in the Boserupian sense, its application has been
vexed by the chicken-or-the-egg problem of whether environmental productivity
results in opportunities for intensification, or whether environmental perturbations
or increasing population densities (i.e., scarcity) leads to intensification processes.
Further, a dichotomy developed regarding the social means associated with
intensification processes and whether it was more often mediated by leaders or
by intensifying populations at large. These latter two issues hinge on the subject
of process and whether intensification, like cultural evolution more generally,
entails unilinear versus multilinear trajectories (Boone and Smith 1998; Morrison
1994). All of these issues are well expressed in the literature of the last decade,
with little progress made in identifying the drivers of intensification (unless they
are indeed as multilinear as Morrison [1994] proposes) and the oftentimes
unsystematic and even contradictory uses of the concept of intensification. But
significant progress has been made in how we model and think about process and
in ways to identify declining foraging efficiency in the hunter-gatherer
archaeological record. The evidence for these assertions is reviewed under the
subheadings below.

Intensification in the Paleolithic

Questions regarding intensification in the Paleolithic prior to the Epipaleolithic (i.e.,


before the terminal Pleistocene and Early Holocene, depending on location) focus
on subsistence during Middle Paleolithic (MP)—Upper Paleolithic (UP) transitions

123
Author's personal copy
J Archaeol Res

(ca. 50,000–40,000 cal BP), the origins of intensification, and the degree to which
these transitions were associated with anatomically modern humans. Research
mainly focuses on the Mediterranean region and Africa, with regional summaries
provided Stiner et al. (2000) and Stiner and Munro (2002) for the Mediterranean and
by Steele and Klein (2009), Steele (2012), and Klein and Steele (2013) for Africa.
Most of this research attempts to determine when exactly human and hominin diets
expanded using zooarchaeological datasets.
In East Asia, for example, Prendergast et al. (2009) present evidence for Late
Upper Paleolithic (ca. 18,000–14,000 BP) faunal intensification from Yuchanyan
Cave, in the Yangtze River basin of South China. Using standard zooarchaeological
quantitative measures—minimum number of individuals (MNI) and number of
identified specimens (NISP)—they show that diet breadth increased to include
freshwater fish, turtles, smaller mammals, and especially aquatic birds, along with
more common Paleolithic prey like cervids. They argue, like many, that
diversification such as this indicates a shift to including lower-ranked resources
in the diet and thus declining hunting efficiency. Less conclusively, based on bone
breakage patterns, they argue bone greasing and marrow extraction in the cervid
assemblage provides yet another marker of more labor input used to extract
diminishing caloric returns, a common marker of declining efficiency. The relative
dearth of cut marks on and crushing of the bones, however, may suggest otherwise.
In South Africa, Klein et al. (2004) use data from Ysterfontein rockshelter and
data from other nearby sites on the Atlantic coast to argue that subsistence practices
between premodern, Middle Stone Age (MSA, ca. 300,000–50,000 cal BP) people
and modern, Late Stone Age (LSA, after ca. 50,000 cal BP) people where
qualitatively different. Pre-50,000-year-old MSA sites like Ysterfontein are
dominated by large tortoise and larger shellfish, whereas LSA sites like Eland’s
Bay contain fish, seabirds, and smaller-bodied shellfish and tortoise. They take this
to represent ‘‘more effective’’ (Klein et al. 2004, p. 5708) exploitation of the local
resource base and more intensive use of previously used resources. Steele and Klein
(2005) and Steele (2012) come to similar conclusions, the former making the claim
that as LSA population densities increased, prey size diminished on South Africa’s
Atlantic and Indian Ocean coasts as functions of predation and harvest pressure (the
latter in the case of shellfish). Steele (2012) tracks similar patterns on the coast, but
notes a MSA–LSA switch from hunting mostly eland to larger (but also more
dangerous) animals like buffalo and pig, which she attributes to innovations in
projectile weaponry. If true, there appears to be evidence for increasing intensity of
marine resource exploitation along the coast, marked by a trend to exploiting
smaller-bodied prey (following Broughton et al. 2011). But inland technological
innovations allowed successful hunts of larger-bodied prey, arguably increasing
returns and hunting efficiency in the face of ostensibly burgeoning LSA human
populations.
In the Mediterranean region, similar patterns pertain. At Vale Boi, an UP site in
Portugal with a faunal record extending from the Gravettian through the
Magdalenian, Manne et al. (2011) and Manne and Bicho (2009) argue that
targeting marine resources (especially shellfish) and rabbits signals diversification
and evidence for greasing large mammal bones (e.g., Equus spp.) signals

123
Author's personal copy
J Archaeol Res

intensification, which they apply in its strictest sense, marked by increased labor
input and decreased efficiency. They contend that the evidence for greasing, at ca.
28,000 BP, is the earliest yet for intensification in Europe.
In Spain, Cortés-Sánchez et al. (2008) and Bicho and Haws (2008) come to
similar conclusions: that increasing dietary diversity and the addition of ‘‘low-
ranked’’ prey like shellfish, fish, birds, and rabbits to Paleolithic diets around
30,000 cal BP signals the initiation of intensification processes in Europe (see
Starkovich 2009 for a similar pattern in Greece). Similarly, Stiner and Munro (2002)
pool data from well-stratified and dated caves and rockshelters throughout the
region to track changes in diet breadth over the course of the last 200,000 years.
They find that MP, ostensibly Neanderthal populations preyed mainly on slow-
moving, slow-growing tortoises and mollusks while UP, ostensibly modern human
populations, increasingly exploited faster-moving, faster-maturing, smaller-bodied
prey like birds (and in the later UP, hares and rabbits), implying that population-
mediated intensification began earlier than the Natufian in the UP. Hockett (2009)
comes to similar conclusions in Portugal—that anatomically modern humans had
more diverse diets than Neanderthals, but argues instead that the greater nutritional
value of these diets gave modern humans a competitive adaptive advantage (see also
Haws and Hockett 2004; Hockett and Haws 2003, 2005). Implied in this latter,
nutritional ecology perspective is that intensification plays at best a subsidiary role
in explaining increased dietary diversity in the UP.
This runs counter to others who see different temporal contexts for intensification
in Europe and the Near East. Speth (2004), for example, uses MNI, NISP, tooth
eruption and wear, and utility indices from Kebara Cave near modern Israel’s coast
to argue that transgressive of the late MP, Neanderthal hunting pressure forced a
shift to eating smaller species (gazelle and fallow deer in response to declining
numbers of red deer and auroch) and younger and smaller gazelle. Evidence of
increased transport distance, marked by high-utility body parts in later deposits,
shows hunters traveling farther to obtain prey through time; both lines of evidence,
Speth argues, mark intensification in the MP. At the other end of the temporal
spectrum, Álvarez-Fernández (2011) and Gutiérrez-Zugasti et al. (2013) agree there
is indeed evidence for increasing exploitation of marine resources, especially
shellfish, in Spain during the UP. But these behaviors, Álvarez-Fernández argues,
represent bead manufacture by anatomically modern humans rather than a
fundamental change in subsistence orientation. Both sets of researchers argue,
however, that intensification of marine and other resources did not occur until the
Mesolithic (early Holocene), in association with increasing population densities (see
also Aura Tortosa et al. 2002a, b; González-Sainz and González-Urquijo 2007).
This sentiment is more or less in agreement with Piperno et al. (2004), who
recovered evidence for processing wild barley (Hordeum spontaneum), perhaps
emmer wheat (Tritium dicoccoides), and other wild cereals ca. 23,000 cal BP at
Ohalo II in Israel, in the very late UP. They also identified a burned feature they
claim resembles an earth oven that may have been used for baking bread made from
processed cereal grains. If so, this would be the earliest direct evidence not only for
wild cereal processing but also for the type of cooking that increases cereal’s

123
Author's personal copy
J Archaeol Res

glycemic index, the energetic value of the food, and perhaps such food’s net return
rate.
Either directly or indirectly, applications of the concept of intensification in
Paleolithic research reveal several consistencies with intensification in hunter-
gatherer archaeology more generally. In nearly all cases, increasing diversity in
zooarchaeological assemblages (measured by frequency [MNI and NISP] and
large to small-bodied abundance indices [AI]) tends to be seen as a marker of
increases in animal exploitation (intensification s.l.), if not outright intensification
(s.s.). Increased effort may be marked by taking either smaller fauna, where
reduced package size is equated with lower caloric return per unit of hunting
effort, or by taking faster prey (like rabbits), which require more effort to catch.
The exception to this is work by Hockett (2009), who focuses rather on the
nutritional ecology of diverse diets, obviating their association with intensification,
at least in the strictest sense. Second, there is a fundamental disagreement as to
when intensification processes began. To many, it did not appear until the
Epipaleolithic or Mesolithic. For many others, however, it is associated with
anatomically modern humans and the latter portion of the Aurignacian (ca.
30,000 cal BP). Resolving this latter conflict of course relies to some degree on
refining ways to identify intensification, both in terms of modeling and
zooarchaeological quantification; it also hinges on being explicit with regard to
whether intensification entails simply quantitative increases in yield or whether it
also requires declining hunting efficiency. It also must take into account the
possibility that the qualitative differences Klein (2008), for instance, sees between
anatomically modern humans and Archaic Homo (sensu Dennell 2009) may entail
fundamentally different species behaviors, with anatomically modern humans
exhibiting a marine focus that was at best dabbled in by Neanderthals and other
Archaic peoples (Álvarez-Fernández 2011; Gutiérrez-Zugasti et al. 2013; Jones
2013, Figs. 2–4; Klein et al. 2004; Mannino and Thomas 2002). If this were the
case, increased faunal diversity marked in large part by a marine focus may be
less a sign of Boserupian intensification (where population pressure drives the
exploitation of lower-return resources) but rather difference in ecological niche
between two different species.

Terminal Pleistocene-Early Holocene intensification

The subject of Mesolithic, Epipaleolithic, Natufian, and other TPEH intensification


is driven mainly by concerns for the origins of agriculture, making it outside the
already broad scope of this review. The subject also has been thoroughly
summarized elsewhere fairly recently (Smith 1998; Stiner 2001; Zeder 2012). I
thus focus on three main issues that are critical to the broader subject of
understanding intensification in hunter-gatherer economies during the Pleistocene–
Holocene transition: (1) how intensification is identified in TPEH contexts, (2)
whether and to what degree intensification is documented in the TPEH, and (3) the
causes of TPEH intensification.

123
Author's personal copy
J Archaeol Res

In identifying intensification, Munro and Atici (2009b) recognize three main


patterns in Late Pleistocene subsistence change in the Mediterranean basin:
intensification, diversification, and specialization. The latter is straightforward and
simple to measure via changes in diet breadth: the restriction of diet to one or two
preferred prey species, as was apparently the case with ibex and reindeer hunting
during western Europe’s Early UP (Grayson and Delpech 2002; Grayson et al.
2001). Broughton et al. (2011) tend to equate diversification (also straightforward
to measure via increased diet breadth) with intensification, using prey size as a
proxy for resource rank. Following this perspective, Munro and Atici (2009a)
present a compelling case for Epipaleolithic faunal intensification in the Levant by
their analysis of zooarchaeological data from 11 sites. Using taxonomic diversity,
mortality profiles, body size, and bone fragmentation data, they show increasing
diet breadth, the targeting of generally younger animals, decreasing body size, and
increasing bone fragmentation and argue that diet breadth expanded and more
energy was expended in the capture and processing of prey animals from the
Kebaran to the Natufian, ca. 21,000–11,500 cal BP (see Stutz et al. [2009] for
similar arguments made based on data from eight Epipaleolithic sites dating
between 19,000 and 12,000 cal BP in the Levant and Aura Tortosa et al. [2002a,
b] for similar arguments for Late Pleistocene marine resource intensification in
coastal Spain). Importantly, they attribute this pattern to increasing population
densities through the Epipaleolithic, though they do not directly measure this
critical variable.
In a similar vein, Atici (2009) uses faunal assemblages from two sites in Turkey
to argue that Epipaleolithic hunters specialized in targeting large-bodied, ostensibly
high-ranked prey like caprines (wild sheep and goat) at Karain B and fallow deer at
Öküzini. But there was a shift ca. 13,500 cal BP in diets to include roe deer, boar,
hare, partridge, and tortoise, which he argues represents a shift to a broad-spectrum
subsistence strategy. In contrast to Munro and Atici (2009a), however, Atici
suggests this was the result of environmental amelioration and increased biotic
productivity during the Bølling–Allerød (14,600–12,900 cal BP) and was not
caused by increased human population density. Atici (2009, p. 12) terms the smaller
fauna exploited later in time ‘‘high yield tertiary taxa’’ due to their abundance and
predictability rather than per-unit yield.
Munro (2004) makes a similar climate change argument with regard to
subsistence changes between the Early and Late Natufian in the southern Levant
(12,800–11,000 cal BP and 11,000–10,200 cal BP, respectively). Using optimal
foraging models (Charnov 1976; Stephens and Krebs 1986) and equating increased
prey diversity, increased abundance of small animals, and more intensive degrees of
carcass processing with intensification, Munro argues that through the Natufian,
hunting pressure was fairly intensive and characterized by high degrees of animal
processing (e.g., bone fragmentation and greasing); in contrast the Late Natufian
was marked by a greater frequency of larger, ostensibly higher-return fauna. She
attributes this shift to declining human population densities during the Younger
Dryas (YD, 12,900–11,600 cal BP) and shorter-term site stays brought about by
demands for greater mobility during this period of reduced and increasingly patchy
environmental productivity. The supposed Late Natufian antecedents to agriculture

123
Author's personal copy
J Archaeol Res

were thus marked by increasing mobility and decreasing diet breadth, a far cry from
the ostensibly more intensive economies that might be expected to prefigure
domestication.
Jones (2006, 2009), for Europe, and Elston et al. (2011), for East Asia, make
similar arguments: that the YD forced a shift toward more intensive subsistence
practices but not necessarily intensification in its strictest sense (see Eren 2012 for
discussions about the YD’s effects on global hunter-gatherers). Jones presents
faunal evidence that populations in southern France shifted to eating a high
number of intensively processed rabbits (Oryctolagus cuniculus) during that time.
She argues, however, that though climate forced the shift, the mass capture of
rabbits from warrens likely increased rather than decreased hunting efficiency. In
China, Elston and colleagues argue that technology can indicate intensification,
with microliths (for hunting), pottery (used for bone greasing), and milling tools
(signaling plant food intensification) as examples. They argue that microliths,
which are nonquantitatively assessed to represent large investments in lithic
procurement and tool manufacture, actually ‘‘enhance hunting efficiency.’’
Similarly, they argue that pottery ‘‘increases the efficiency of bone grease
extraction’’ (Elston et al. 2011, pp. 405–406). Both are hypothesized to represent
ways of increasing caloric yield in response to the reduced and less predictable
resource base of northeast Asia during the YD. This provides a thought-provoking
case where increased rather than decreased efficiency marks intensification as a
response to environmental stress.
Clear in the preceding is that TPEH intensification, as with the Paleolithic, is
most often associated with diversification and, when linked to optimal foraging
models, to increased diet breadth. Per the TPEH literature, intensification can thus
arguably be marked by increased bone fragmentation and carcass processing
(signaling increased labor), by greater taxonomic diversity, and by increasing
exploitation of smaller-bodied prey. Unlike research associated exclusively with the
Paleolithic, however, there also is a critique of the application of optimal foraging
theory to studies of TPEH intensification. Zeder (2012) makes this clear in her
recent review of the BSR, arguing that differential and gendered foraging goals
(Hawkes 1991), risk and uncertainty (Winterhalder et al. 1999), problems linking
diversification with intensification (Broughton and Grayson 1993; Grayson and
Cannon 1999), and different currencies (e.g., using overall nutritional value rather
than calories [see Hockett and Haws 2005]) all result in different interpretations of
BSR subsistence data and, by association, whether intensification is truly identified
in the zooarchaeological record (see also Hockett and Haws 2003). There is more
variability in the causes and results of TPEH intensification; some attribute
widening diets and more intensive processing to population increase, others to
climate change during the YD. Even more important is that researchers like Munro
and Atici (2009a) see intensification in the Boserupian sense of declining efficiency,
whereas others (Elston et al. 2011) argue that technological innovations, though
entailing their own new costs, may actually increase net efficiency. These issues
play out in greater detail in the global hunter-gather record, especially during the
Holocene.

123
Author's personal copy
J Archaeol Res

Coastal intensification in South Africa, Australia, Oceania, and South America

The subject of intensification based on marine resources in the Holocene builds


mainly on data that show increased exploitation of shellfish over time. This marine
and intertidal subsistence focus is thus of considerable consequence to understand-
ing the extent of humankind’s exploitation of diverse ecological niches and,
critically, the signals and causes of marine resource intensification. Mannino and
Thomas (2002) set the stage for this discussion in their review of the literature on
hunter-gatherer shellfish exploitation to 2001, where they argue that marine
resources, especially intertidal shellfish, are a resource prone to a ‘‘tragedy of the
commons’’ (Hardin 1968). From this perspective, lack of management leads to
overexploitation marked by declining age profiles, declining size, and shifts to
smaller, harder-to-procure or harder-to-process (i.e., ‘‘lower ranked’’) species in
shell middens over time. They do note, however, that taphonomic and behavioral
processes can affect the record as well, with children’s shellfish foraging, for
example, tending to appear to broaden the diet (Bird and Bliege Bird 2000).
Discounting this consideration as the exception rather than the rule, the authors
argue that the world’s coasts represent a linear, contiguous resource base that, due to
sequential overpredation, likely led to human population dispersals, with groups
moving on into familiar but unexploited habitats as they depleted the resources in
already exploited ones (see Allen and O’Connell [2008] and O’Connell and Allen
[2012] for how this process may have driven the rapid colonization of Australia).
Evidence for population-mediated marine resource intensification comes from
the same place where the process ostensibly began: South Africa. Jerardino (2012)
makes this case based on her analysis of ‘‘megamiddens’’ in and around Eland’s Bay
dating to 3000–2000 BP. Intensification, she argues, is marked most clearly by the
middens themselves, which consistently date to this time, with little evidence for the
accumulation of large deposits either before 3000 or after 2000 BP. Intensification,
she argues, also is identified in midden faunal assemblages, which are typically
about 80% shellfish (mostly black mussel [Choromytilus meridionalis]) but include
some shorebirds, fish, and terrestrial fauna, mostly small bovids. Data from nearby
Lambert’s Bay show increasing reliance on smaller and smaller-bodied prey,
reflected mainly by massive increases in the exploitation of black mussels and
limpets (Patella sp.) between 3000 and 2200 BP (Jerardino 2010). She argues this
was brought about by population circumscription (marked by distinctive inland
versus coastal isotopic signatures in human bone), more evidence for interpersonal
violence, and less inland lithic material on coastal sites.
A different perspective is evident in recent research in Australia, which has a
long history of controversy regarding hunter-gatherer intensification, much of which
is concerned more with social rather than explicitly economic intensification
(Hiscock 2008; Lilley 2000; Lourandos 1988; Lourandos and Ross 1994).
Controversy over economic intensification is reflected, nonetheless, in the debate
between Walters (2001) and Ulm (2002). Walters links expansions into new
territory, more fishing sites, and increased discard of fish bone on southeast
Queensland’s coast during the late Holocene to the development of more intensive
social networks, increased sedentism, and, critically, higher population densities.

123
Author's personal copy
J Archaeol Res

Ulm questions the validity of these claims on taphonomic and methodological


grounds. In his review of data from sites in the region, Ulm (2002) notes that fish
remains are rare at many excavated sites and that some actually show decreasing
quantities of such in the late Holocene. He argues that inconsistent sampling and
recovery measures, small samples, few AMS dates, coastal erosion, and sea-level
transgression are the main problems in identifying coastal intensification there and
concludes that there is no clear evidence of fishery intensification in Queensland in
the late Holocene (there also is no real evidence that there is not). The dilemma Ulm
encounters in identifying intensification in Australia’s archaeological record may,
however, reflect a more general problem in Australian archaeology. As Lilley
(2000) notes, despite Australia’s prominence in the literature on hunter-gatherer
intensification, degrees of intensification are so low across much of the continent
relative to other places where hunter-gatherers appear to have intensified their
subsistence economies that it is indeed quite difficult to identify in Australia’s
archaeological record. Australia’s desert interior has, however, its own record of
economic intensification, much of it arguably occurring in the strict Boserupian
sense in light of increasing population densities after the middle Holocene (Smith
2013; Williams 2013).
In New Zealand, Nagaoka (2002, 2005) presents an arguably clearer picture of
intensification. Nagaoka uses element utility indices, degrees of bone fragmentation,
and prey and patch choice models to analyze and interpret faunal assemblages from
Shag Harbor and Shag Mouth, two late Holocene coastal sites on New Zealand’s
south island. At Shag Mouth, she generates a Moa-quail index from the site’s faunal
assemblage and NTAXA, a measure of diversity, to show that prey body size
declined while taxonomic diversity increased over time. She claims this shows
increasing diet breadth and decreasing foraging efficiency. At Shag Harbor, she tests
two hypotheses: (1) that as humans extirpated moa from near the site they would
have to travel farther afield to hunt these large birds, which should be reflected in
field processing to remove low-utility elements (meaning only higher-utility
elements should be found at the Shag Harbor site), and (2) evidence of bone
greasing (marked by increased breakage) should increase over time with greater
hunting pressure causing people to try to extract as much energy as possible from
their prey. In general, the expectations derived from these hypotheses are met,
though I question that if populations experience significant resource stress, is it also
not likely that low-utility elements would be transported back to camp without field
processing, regardless of distance? Transporting very low utility parts would thus
also be a clear sign of declining efficiency in light of the field processing and
transport behaviors observed by Binford (1978) and predicted by a variety of central
place foraging models (e.g., Bettinger et al. 1997; Metcalfe and Barlow 1992).
Somewhat equivocal evidence for coastal intensification comes from Hawaii and
South America. Morrison and Hunt (2007), for instance, use taxonomic diversity
and evenness values to argue that shellfish-oriented diet breadth expanded between
AD 1400 and historic times for fisher-farmers living on the northwest shore of
Kauai. They attribute this mainly to overpredation, a case of ‘‘fishing down the
web’’ (see also Reitz 2004 for a similar situation in Florida), but they are ambivalent
in their conclusion, noting that a change in settlement and subsistence in late

123
Author's personal copy
J Archaeol Res

prehistory toward more terrestrial gardening (which would have required staying
closer to fields and therefore more shellfish exploitation nearshore, as opposed to the
more distant coral reef) may account for this pattern as much as population-
mediated resource intensification. In Tierra del Fuego, Zangrando (2009) presents
ichthyoarchaeolofaunal evidence for widening diet breadth between 6500 BP and
150 BP, with the most clear evidence after 1000 BP. He makes this conclusion
based on evenness, diversity, and indices of land mammal: pinniped: marine fish
abundance to show resource diversification occurred while terrestrial mammal
exploitation remained static. Zangrando argues this may have been due to increases
in population farther inland driving, via circumscription, intensification on the coast,
rather than coastal population increase (see Kennedy 2005 for a similar example
from California). While presenting another clear case of subsistence shifts
consistent with the definitions of intensification s.l and arguably s.s., he indicates
that ‘‘increased foraging efficiency’’ (Zangrando 2009, p. 590) is a hallmark of this
process, in clear refutation of the Boserupian perspective.
In sum, Holocene intensification in global coastal contexts reflects several
semantic and methodological problems common to hunter-gatherer intensification
research in general and brings up new issues—settlement and circumscription—that
play important roles in understanding hunter-gatherer intensification elsewhere. In
applying the concept of intensification, most researchers tend toward the optimal
foraging perspective that increased labor is a marker of declining foraging efficiency
(i.e., in the Boserupian sense), but that at least in some cases increased efficiency
may be affiliated with intensification (s.l.) as well. In terms of method, as in all the
research reviewed thus far, standard zooarchaeological quantification applies (e.g.,
NISP, evenness, and AI) but arguably more attention is paid to taphonomic process
and recovery techniques as problems facing the identification of intensification in
coastal settings (not surprising given the size of many fish bones and such active
geomorphic contexts). Lastly, Holocene-focused coastal researchers add to the
intensification debate the dimension of how subsistence articulates with settlement
patterns (but see Stiner et al. [2000] on mobility’s effect on diet during the TPEH)
and intergroup relations, noting that changes in the former may result in increased
dietary diversity or other markers of intensification. In terms of the latter, by the
time the Holocene rolled around, coastal hunter-gatherers (as elsewhere) operated in
less uncircumscribed vacuums than in continua of competing and cooperating
groups both along the coast and farther inland.

Intensification in Pre-Columbian North America

The subject of hunter-gatherer intensification in North America is broad, even when


the considerable literature from the continent’s west coast (covered in succeeding
sections) is omitted. Clear in the recent literature is the fact that the concept of
intensification has come to be employed in increasingly diverse ways, at times
outside the constraints of optimal foraging theory (but still usually within the
bounds of ecology) and increasingly in such a broad sense as to obviate some of its
utility for modeling and theory. The bulk of this literature, however, focuses on

123
Author's personal copy
J Archaeol Res

identifying the presence, causes, and effects of declining foraging efficiency in the
continent’s hunter-gatherer archaeological record.
Starting in the east, Raber (2010) employs intensification in its widest sense,
presenting evidence for and exploring the causes of increasing rockshelter use in the
Late Prehistoric (AD 900–1650) in Pennsylvania. Looking specifically at Mykut
rockshelter, he argues that these increases, rather than simply marking increased
population densities, instead indicate a diversified approach to managing risk
associated with increased climatic and environmental variability during the Little
Ice Age (LIA; 650–150 cal BP) (see also Morgan 2009 for a similar perspective
from California). Hunter-gatherers facilitated the shift to foraging and especially
deer hunting as an alternative or supplement to farming by operating out of upland
rockshelters in and near deer hunting grounds. In contrast to the broad focus of
Raber, White (2013) analyzes a temporally controlled database of house sizes over
time in the eastern woodlands. He uses these as a proxy for household size, finding
that size peaked between the Late Archaic and Middle Woodland periods (3800
BC–AD 500). He attributes this increase to population-level intensification marked
by the lower age at which children began to make substantive contributions to
household economies, mainly by gathering low-yield, costly to process, but also
abundant wild plant resources. This labor, in turn, helped support larger populations.
For Raber, hunting intensification simply means increase (as a way to reduce risk),
with rockshelters’ proximity to deer hunting grounds increasing net returns on
hunting due to decreased carcass transport costs (see also McGuire et al. 2007;
Morgan et al. 2012). For White, intensification is mediated by increased population
density, which requires additional work by children to offset population-resource
imbalances.
Two studies in Texas point to how mortuary data can be used in Boserupian-type
intensification studies and how such studies speak to the conditions under which
agriculture should or should not be adopted by hunter-gatherers. For the former,
Hard and Katzenberg (2011) present isotopic data from 198 human skeletons
recovered from 16 coastal mortuary sites and from faunal bone that dates between
7000 and 300 BP. Carbon signatures remain more or less stable through most of the
Holocene, save for at three sites with post 2500 BP components. These show
increases in the exploitation of C3 and C4 plants after that time, which Hard and
Katzenberg, in Boserupian (and Binfordian) mode, attribute to population packing
that forced a switch to lower-ranking plant resources. Johnson and Hard (2008)
build on Hard and Katzenberg’s work using expectations derived from Binford’s
(2001) Frames of Reference to explain why maize agriculture never really took
hold, despite being well established nearby in the American Southwest, Southeast,
and Mexico. Binford (2001) indicates that when population growth leads to
densities that constrain mobility (really just circumscription, at about 9 people/
100 km2) and when environmental characteristics (measured unsurprisingly by
effective temperature) entail high resource productivity, hunter-gatherers should
shift from hunting large fauna to aquatic resources (where available) or, in the
absence of aquatic resources, low-return plant foods. In their analysis, Johnson and
Hard (2008) show that distinct hunter-gatherer populations were constrained mainly
to specific ecological niches (coasts, rivers, inland areas) and that coastal and river-

123
Author's personal copy
J Archaeol Res

oriented people focused in the Late Prehistoric on lower-return but abundant aquatic
resources, as expected. The frequency of large cemeteries, they argue, attests to
reduced mobility, population packing, and intensification (Ricklis and Weinstein
2005). Inland, a focus on low-return plant foods is identified mainly by the ubiquity
of earth ovens used to process these costly-to-process resources. They touch on but
never really develop the idea that intensive hunting and gathering can support
population densities as high as some prehistoric agriculturalists, which could
explain why maize was not adopted there, an idea first explicitly formulated by
Baumhoff (1963; but see Bettinger 2015) as an explanation for why maize
agriculture failed be adopted in prehistoric California.
On the Plateau, Lepofsky and Peacock (2004) develop an alternative ranking
system to the diet breadth model and compare the expectations of this model to
direct (i.e., macrobotanical) and indirect (i.e., technological) markers of plant food
use over the last 4500 years. The model ranks plant foods in terms of their
availability, abundance, palatability, and persistence, the latter a measure of
resistance to overharvesting. They conclude that most roots, many types of greens,
and a variety of berries would most likely be subject to ‘‘intensification.’’ Seeds and
nuts rank very low. Their strongest archaeological dataset comes in the form of the
frequency of earth ovens used to process roots, which first appeared around
4500 BP. Their use peaked at 2400–1500 BP and after 800 BP, which the authors
claim marks intensification of geophyte exploitation. They define intensification in
its broadest sense as ‘‘the process of increasing plant food production (i.e., increased
output) via various mechanisms that may or may not involve increased energy
costs,’’ or as intensification s.l.
Thoms (2009) takes the analyses of geophyte exploitation one step farther in his
overview of the evolution of hot rock cookery across North America. Drawing in
large part on his prior work (Thoms 1989, 2003, 2008), he begins with the bias that
increases in hot rock cookery signal ‘‘population packing’’ and ‘‘land-use
intensification,’’ clearly conceiving of intensification in the Boserupian sense. He
then provides a wide range of data to show that cooking with hot rocks started
around 10,000 cal BP, was common by 8500 cal BP, increased sharply at
4000 cal BP, and in most areas increased thereafter. Due to the reduced efficiency
of cooking with hot rocks relative to open fires (Thoms 2003), he argues this process
is a clear case of population-meditated declining foraging efficiency and thus
intensification s.s.
Yu (2006) extends the analysis of geophytes and pit cooking with an eye to
identifying commonalties in the intensification process across the American West.
Like Johnson and Hard (2008), Yu draws on the ecologically and ethnologically
derived expectations for hunter-gatherer subsistence developed in Binford’s (2001)
Frames of Reference to analyze the causes behind variability in frequency of earth-
oven cooking in the American Southwest, Basin and Range, and Pacific Northwest.
In her diachronic analyses of pit frequency and morphology across these regions,
she concludes that (1) prehistoric hunters who used pits, like those in the Basin and
Range, were unlikely to adopt cultivars without first intensifying their use of wild
plants (predicated on the greater returns of hunting); (2) gatherers who intensively
used pits, like those in the pre-Puebloan Southwest, were prone to adopting cultivars

123
Author's personal copy
J Archaeol Res

because the returns on pit-processed foodstuffs were so low; and (3) aquatically
oriented groups who used pits, like those in the Pacific Northwest, were less likely
to adopt cultivars until their mobility became constrained by circumscription.
Though clearly an example of using the concept of intensification s.s (it explains
how declines in foraging efficiency affiliated with behavioral shifts toward the
adoption of cultivars helped mediate population-resource imbalances in groups that
operated in the gray zone between foraging and farming), Yu is equivocal about
what intensification entails, opting for Binford’s (2001, p. 188) simple definition of
intensification as ‘‘any practice(s) that increase food productivity per unit area,’’
implying that decreasing efficiency is not necessary for intensification (s.l.) to occur.
Finally, Betts and Friesen (2004) use specialization, diversification, and
investment to analyze diachronic subsistence change from roughly 700 BP and
into very early historic times on the Mackenzie River delta in Canada’s Northwest
Territories. They base their analysis on faunal assemblages from three sites—Cache
Point, Pond, and Kuupak—which together contain six different components. Using
evenness indices (Reitz and Wing 1999) as proxies for specialization, they argue the
Thule/Inuit specialized in the pursuit of Beluga whale (Delphinapterus leucas) in all
components. Using richness indices, however, they note greater diversification in
the subsistence economy in the latest components, with increases in birds and fish
driving the change. Lastly, by looking at investment through the lens of netted vs.
non-netted fish and seasonality data as proxies for occupational intensity, they argue
investment (both technological and settlement) increased very late in the sequence
as well, particularly at the onset of the historic period. They mostly avoid discussion
of the historically mediated context of culture contact and disruption, however,
choosing instead to focus on population pressure as a causal mechanism for this
change. They do, however, conflate increased productivity with increased
efficiency, which may or may not be the case given the difference between
intensification s.s. and s.l.

Great Basin

Outside the west coast, the Great Basin has arguably generated the greatest amount
of literature on hunter-gatherer intensification in North America, the result, like
Australia, of the region’s long history of contributions to hunter-gatherer theory
(and, like Australia, a somewhat surprising circumstance given the region’s
relatively low aboriginal population densities). Recent research not surprisingly runs
the gamut from using intensification in its broadest sense, as simply entailing doing
more of something (even if this entails increased foraging efficiency) to drawing
into the archaeological fold the confounding factor of differential foraging goals and
their effect on interpreting the behavioral proxies usually used as markers for
intensification.
Recent research that identifies intensification s.l. is rampant in the region. For
instance, Garfinkel et al. (2007) use the term in its broadest sense, as simply
signifying increase. They argue that increases in the quantity of petroglyphs during
AD 600–1300 and pictographs in historic times in the deserts of southeastern
California occurred during periods of resource and cultural stress. Morgan et al.

123
Author's personal copy
J Archaeol Res

(2012) employ a superficially similar perspective, that increasing residential use of


high-altitude environments is intensification in a broad sense, undertaken in part to
decrease carcass transport costs (which increased hunting efficiency) as a way to
supplement the caloric intake of Fremont part-time horticulturalists. Similarly,
Eerkens (2004) makes the case that pottery use in very late prehistoric eastern
California was linked to increasing emphasis on the exploitation of small seeds.
Pots, he argues, resulted in more efficient and less-labor-intensive cooking.
In contrast, in much of his work, Eerkens employs a much more Boserupian take
on intensification. For instance, Eerkens et al. (2004) provide data from test
excavations of rock rings and burn features in eastern California’s Owens Valley to
argue that these features were indeed associated with storing and processing green
as opposed to brown piñon pine (Pinus monophylla) cones, a strategy that developed
with regional population increases ca. 500 BP. The former is a more intensive way
of subsisting on piñon nuts because it entails more processing costs, mainly due to
the charring needed to get the pine nuts out of the cone. Similarly, Eerkens and
Rosenthal (2002) describe sampling and AMS dating of four different types of plant
processing features in southeastern California’s Mojave Desert: pits, pit-hearths,
hearths containing small seeds, and hearths lacking seeds. They find that earlier late
Holocene features were used for processing geophytes and the latest ones (those
dating to 300 or fewer years ago) were used for processing small seeds. They
attribute this to the ‘‘small seed intensification’’ endemic to the region in late
prehistory, but they are equivocal as to whether this was the result of population
pressure, mainly due to the difficulty of identifying population expansion in the
archaeological record (Eerkens 2003).
Turning many of the assumptions about what marks intensification on its head is
the debate, mainly in the pages of American Antiquity over the last decade, over
foraging goal variability and increases in artiodactyl hunting during the Middle
Archaic (5000–1000 BP) (Morgan and Bettinger 2012). The debate centers on
whether hunting intensification (s.l.), marked by increased frequency of artiodactyl
remains, increases in biface manufacture, and the florescence of rock art depicting
sheep and deer hunting in the Great Basin and California signals the development of
long-distance, logistical, prestige-oriented hunting by males seeking to increase
their status and by dint their reproductive success (Hildebrandt and McGuire 2002;
McGuire and Hildebrandt 2005; McGuire et al. 2007). Dissenters ask whether such
phenomena were driven by late Holocene environmental amelioration, concomitant
increases in artiodactyl abundance, or increased hunting of high-ranking resources
due to more frequent encounter rates and how costly long-range logistical hunting
was relative to other alternatives (Broughton and Bayham 2003; Byers and
Broughton 2004; Byers et al. 2005; Codding and Jones 2007; Grimstead 2012;
Whitaker and Carpenter 2012). The subject is unresolved but germane; even though
a shift to eating more high-ranking prey might be thought of as de-intensification in
the Boserupian sense, it points to the difficulties of using zooarchaeological data,
ostensibly the clearest marker of subsistence, to interpret causes of change in
economic behavior, a problem that clearly prefigures more recent efforts at
intensification-informed middle-range modeling.

123
Author's personal copy
J Archaeol Res

Intensification is thus broadly conceived in North American hunter-gatherer


archaeology. Methodologically, it shows how sources of data not usually seen in
intensification studies—stable isotopes, site structure, and features—can be used to
identify intensification in the archaeological record. This research also shows,
however, how identifying intensification can be confounded by middle-range
problems associated with how foraging goal variability affects zooarchaeological
deposits and how to identify population pressure in the archaeological record.
Finally, in terms of theory, though much of this research hinges on the idea that
‘‘population packing’’ (Binford 2001) is a cause of intensification, there are hints
that offsetting risk also may encourage such behaviors. Risk-sensitive foraging has
been shown to tend to meet women’s foraging goals (Bliege Bird 2007; Codding
et al. 2011; Jochim 1988), meaning, if true, that intensification processes may vary
according to gender. Further, risk models employ not only mean but also variance of
return rates and thus require greater knowledge of expected encounter rates to be
successfully applied in a quantitative sense, which increases the complexity (and
quite arguably the explanatory power) of such approaches (Stephens 1990;
Winterhalder and Goland 1997). These models also imply that subsistence goals
may be driven more by satisficing than optimizing decision-making objectives, a
dichotomy with important implications regarding what ultimately causes changes in
labor input, innovation, and yield (Morgan 2009; Winterhalder and Leslie 2002).
Put simply, risk-sensitive foragers may forgo maximizing immediate returns to
ensure they meet minimum caloric or nutritional thresholds over the long term,
especially in contexts where incomplete information or environmental variability
result in uncertainty with regard to economic decision making.

The complexity of complexity: Intensification in the Pacific Northwest

Perhaps nowhere have concepts of intensification been so linked to the evolution of


complex hunter-gatherers than in the Pacific Northwest. As alluded to earlier,
fundamental questions hinge on identifying when exactly intensification-related
sociocultural complexity began to develop, whether resource stress, abundance, or
aggrandizing behaviors led to these processes, and whether population increased
before or after storage, sedentism, and other markers of cultural complexity began to
develop (Ames 1981, 1994, 2003; Ames and Maschner 1999; Fitzhugh 2003b;
Hayden 2001; Matson 1992; Prentiss and Kuijt 2004). Critically, answering these
questions hinges on environmental characteristics, technological developments,
social relations, resource management, and the ways these phenomena support or
refute more efficiency-oriented perspectives on intensification.
Such perspectives are found in the more-or-less Boserupian approach of Butler
(2000), who provides zooarchaeological evidence for the impact of human predation
on mammals and fish near the Columbia River mouth. She uses AI and NISP at
eight multicomponent sites to show a shift in consumption from more and smaller-
bodied fish in very late prehistory to more mammals and larger-bodied fish after
Euroamerican contact. Butler convincingly argues, despite fairly small sample sizes,
that the historic pattern represents a shift from eating low-ranked to high-ranked

123
Author's personal copy
J Archaeol Res

resources due to prey population rebound associated with indigenous population


decline and reduced predation pressure, as predicted by the diet breadth model.
Importantly, she implies that prehistoric economies were intensive enough to
suppress prey populations.
A somewhat similar population-based explanation for intensification comes
from Fitzhugh’s (2003b) perspective on the evolution of complex hunter-gatherer
societies in the Gulf of Alaska. He tracks diachronic behavioral changes from the
early middle Holocene to culture contact with Russian and American traders in the
19th century with a robust sample of house features, artifacts, faunal remains, and
radiocarbon dates. Using these data as proxies for population density and degrees
of sociocultural complexity, he argues there was a critical period between 3500
and 800 BP when population densities and cultural complexity (marked by larger
houses, social stratification, and evidence for corporate control of resources)
increased in response to the wholesale adoption of a delayed-return, storage-based
economy. Fitzhugh uses intensification to explain this shift, arguing that increasing
population densities required fundamental behavioral transitions, in particular
adopting storage to offset winter resources and targeting lower-ranked prey to
ensure that increasingly circumscribed populations were fed. But in this and other
work Fitzhugh (2001, 2002, 2003a) also focuses on technological innovations such
as toggling harpoons and fish nets that would have increased fishing and hunting
returns, arguably increasing predation efficiency. He maintains, however, that the
processing and storage of netted fish that was so essential to generating winter
stores of food required increased labor input as well as new forms of resource
distribution that hinged on ascribed leadership roles, private property, potlatching,
extensive trade, warfare, and slaving. In sum, population increase in the face of
pronounced seasonality appears to have driven the labor increases, technological
innovations, and changes in how resources were redistributed in prehistoric
southern Alaska.
In contrast to Fitzhugh’s ecological perspective, Hayden (1981, 1990, 1995,
2001) argues instead that it was aggrandizing behaviors by the region’s (and
global) transegalitarian hunter-gatherers that led to increasingly intensive modes of
economic production. Though he published most of this work more than a decade
ago, though his concern is mainly with the development of inequality, and though
he draws much of his evidence from archaeological evidence from much farther
afield, this work is important for the theoretical alternative it provides relative to
long-running ecological debates over population, resource abundance, climate
change, and their effects on the evolution of intensive hunter-gatherer economies.
He based his thinking in part on his long-running investigation of the Keatley
Creek site on the Fraser River in southwestern British Colombia, where substantial
differences in house pit size (there are at least 115 recorded at the site) and content
between approximately 2600 and 1000 BP allude to the development of
pronounced socioeconomic disparities; larger houses contained evidence for
higher-quality food items, storage, extralocal artifacts, tools, and other markers
indicating wealth and prestige (Hayden 2000, 2005, 2007). Prestige-seeking
individuals, Hayden argues, were the drivers of the economy that underwrote these
behaviors (for an alternative, climate-change explanation for Keatley Creek

123
Author's personal copy
J Archaeol Res

socioeconomic intensification see Prentiss et al. 2003). Though wealth accumu-


lation, private property, and differential access to resources are considered rare in
most hunter-gatherer societies, Hayden theorizes these behaviors confer benefit to
both their practitioners and to the corporate groups of which they are an integral
part. This is because of the redistributive nature of these economies, the risk-
reducing effects of large-scale storage, the efficiency of being able to trade
nonfood items for food, and, most importantly, the social obligations of giving and
receiving—a form of ‘‘social storage’’ (O’Shea 1981) that makes such economies
more resilient and competitive in the long run. Once established, aggrandizing-
based economies become progressively more intensive due to the positive-
feedback relationships in such systems. These feedback mechanisms, Hayden
proposes, provide the impetus for increased economic production, or intensifica-
tion s.s. and perhaps s.l.; either, of course, generates the surplus necessary for such
systems to operate.
Alternative perspectives are seen in research on more diverse intensification
pathways. In his summary of Pacific Northwest prehistory, for instance, Ames
(2005) argues that intensification can occur two ways, through either reduced or
increased efficiency, and implies that the net result is ultimately the same: increased
food production. Butler and Campell (2004) adopt a complementary perspective,
ultimately arguing against the idea of declining foraging efficiency and Boserupian
intensification in the Pacific Northwest (see also Campbell and Butler 2010). In their
review of the region’s zooarchaeology, they generate a comprehensive, time-
transgressive archaeofaunal dataset (based on diversity, evenness, and AI) to argue
that (1) ostensibly burgeoning human populations over the course of the Holocene
reduced neither salmon nor cervid prey populations, (2) salmon were intensively
exploited from the beginning of the sequence, and (3) high-ranking cervid predation
increased toward the latter end of the sequence. The only substantial change was
between 2500 and 700 BP, when herring fishing substantially increased, which they
claim indicates a shift to more logistical strategies. It also seems plausible that it
represents a shift to eating lower-ranked species and therefore declining fishing
efficiency. They basically argue in favor of an abundance as opposed to a resource-
stress explanation for the development of Pacific Northwest coast sociocultural
complexity where the early adoption of a specialized, narrow-spectrum, high-return
salmon economy supported increasingly large populations. On the theory side of the
equation, they argue against Ames’s (2005) and Matson’s (1983) linkage of both
declining and increasing efficiency to intensification, maintaining that this
dichotomy is, ‘‘…more than a semantic confusion, it is also a theoretical schism’’
(Butler and Campell 2004, p. 336).
Moss (2012) takes a similar view, arguing against intensification explanations for
cultural evolution in the Pacific Northwest on mainly middle-range grounds. She
makes a polemical argument that intensification models do not take into account
ecological parameters at multiple scales and therefore cannot unequivocally entail
intensification. She supports her argument by identifying how local environmental
factors, diachronic environmental changes, and seasonality can skew zooarchaeo-
logical assemblages to appear to represent intensification when they, instead,
indicate abundance in the surrounding environment. Seasonality of salmon runs,

123
Author's personal copy
J Archaeol Res

‘‘hot spots’’ for herring, and habitat preference for different shellfish species, she
argues, all can result in overrepresentation of these taxa in archaeofaunal
assemblages. In other words, local environmental context and environmental
productivity matter.
Cannon and Burchell (2009) deal with this problem in a manner prescribed by
Moss, by looking at shellfish mortality profiles to account for human rather than
environmental shell midden formation processes (sensu Schiffer 1987). They
provide convincing evidence that coastal hunter-gatherers in British Columbia
actively managed their shellfish resource base by only harvesting senile clams at
main residential bases, though younger (but still ‘‘mature’’) clams were harvested
at less-intensively used camps. This counterintuitive scenario is explained as
representing a rare instance of controlled harvest and management of resources
affiliated with privatization of the resource base at larger, more permanent sites.
This incentivized ensuring sustainable harvests from year to year, hence avoiding
the ‘‘tragedy of the commons’’ (Smith and Wishnie 2000). In this case the
behavior of larger human populations resulted in increases in prey size, in
contrast to the way Broughton et al. (2011) and many others interpret markers of
intensive resource exploitation. If true, this is quite notable in that manage-
ment—deliberately forfeiting maximal immediate returns to ensure long-term
sustainability—might be able to support larger populations over the long haul,
more in accordance with risk-sensitive, satisficing economic decision making
than optimization.
In the Pacific Northwest then, the key to understanding the economic basis for
how large, complex, sedentary populations developed hinges mainly on the idea of
efficiency. Work in southern Alaska and on the Columbia River points to the
stressing effect large populations had on the Pacific Northwest resource base and
focuses on the labor required to make large-scale, delayed-return subsistence
economies work. In contrast, on the western Plateau, Hayden argues it was
aggrandizing behaviors that led to surplus production and that these behaviors,
once entrenched, resulted in a positive feedback loop that resulted in increasingly
intensive work, storage, and redistributive behaviors. Broader regional perspec-
tives suggest the early adoption of intensive, salmon-based systems and the
development of technologies to more effectively exploit the abundant resource
base of the region led to the large, arguably sustainable population densities of the
late Holocene. To this end, it is quite clear that Butler and Campell (2004) are
correct in their assertion that there is a fundamental theoretical and methodolog-
ical divide between seeing intensification in the Boserupian sense—as entailing
declining efficiency—and the broader perspective that it also may entail increased
efficiency. The former predicts not only predation/harvest pressure but also greater
demands on labor to increase yield. The latter implies that innovation, risk-
sensitive management, and other behavioral changes might offset or even reduce
demands on labor while still supporting large and even increasing populations.
The end result, as Ames (2003) claims, might indeed be the same, but the process
and explanation for how increasing economic yield was accomplished is anything
but.

123
Author's personal copy
J Archaeol Res

Land of plenty or scarcity? Intensification in California

As in the Pacific Northwest, the topic of intensification in California is a broad one


with a long history of contention over the roles that population, resource abundance
versus scarcity, and economic intensification play in the development of sociocul-
tural complexity (Codding et al. 2012; Hull 2012). It hinges in large part on research
on the evolution of marine resource use and Chumashan sociocultural complexity in
southern California (Arnold 2004) and on the late prehistoric period in central and
northern California, again with an emphasis on coastal economic processes. A
related topic, burning and other landscape management practices (Cuthrell et al.
2012; Jordan 2003; Lewis 1973), is covered in a 2013 issue of California
Archaeology (vol. 5, no. 2) devoted to the subject and is not reviewed here. The bulk
of the research on burning, however, points to just how ubiquitous this low-
investment but potentially high-return practice (in large part because the relatively
simple act of burning increases abundance and hence reduces search and
procurement times for desirable species) was in aboriginal California and just
how large a role burning played in the development of Precolumbian vegetation
communities across the state.
On the southern California coast and Channel Islands, recent research not only
incorporates traditional approaches to understanding population-mediated marine
resource intensification akin to those seen for Paleolithic and Holocene coastal
research but also exhibits how confounding factors like local contexts and the
temporal dimension of marine resource extraction affect the identification of
intensification. It also shows how important innovation is to processes of
augmenting economic production, processes that may increase rather than decrease
foraging efficiency. In terms of declining foraging efficiency, Erlandson et al.
(2008) use metric data on shell size recorded from over 11,000 specimens collected
from 41 archaeological components that document 10,000 years of human
occupation on San Miguel Island, one of southern California’s northern Channel
Islands. Shell size for California mussel (Mytilus californianus) and red abalone
(Haliotis rufescens) show a clear trend toward smaller sizes over the Holocene. The
data for black abalone (Haliotis cracherodii) are more equivocal, showing a slight
increase during the mid-Holocene, which the authors attribute to increased human
predation on sea otter, a major abalone predator. They interpret these data in light of
noncultural factors that may have caused changes in shellfish size (e.g., sea surface
temperature) and find that human predation most likely accounts for the decrease in
shellfish size (see also Braje 2007; Braje et al. 2007; Erlandson 2001; Kennett 2005;
Rick 2007).
Local and temporal factors, however, can affect these types of more-or-less
straightforward assessments of declining foraging efficiency. On the mainland (and
in an argument reminiscent of Moss’ [2012] for the Pacific Northwest), Perry (2004)
asserts that prey frequency in shell middens can be predicated on local
environmental factors. She notes that highly erodible shale in the high surf zone
of her study area prevents mussel colony establishment, leading to a greater focus
on harvesting lower-ranked black turban snail (Tegula funebralis) rather than
higher-ranked mussel. Expanding this type of reasoning, Thakar (2011) analyzes a

123
Author's personal copy
J Archaeol Res

small sample of shellfish from a midden on Santa Cruz Island, the largest of the
northern Channel Islands. She shows that the quantity of Tivela stultorum (Pismo
clam) decreased but size increased over a very brief period ca. 1100 BP. Because
Tivela establish themselves infrequently and sporadically, Thakar counterintuitively
argues that this change shows intensive harvest of a single colonization/establish-
ment event, with younger, smaller clams exploited first, and then the remaining,
more difficult to procure (because they move to deeper water with age), larger clams
exploited later, perhaps over a century-long period. The gist here is that large,
semisedentary, and sociopolitically complex Chumash groups mapped onto and
intensively exploited Pismo clams whenever and wherever they became established
and available, resulted in more of a short-term intensification than a long-term
diachronic trend (and a potential ‘‘tragedy of the commons’’ that contrasts with the
management of shellfish patches that Cannon and Burchell [2009] identified in the
Pacific Northwest).
On the other end of the marine exploitation spectrum, important contributions
have been made regarding shifts toward the taking of larger-bodied fish and sea
mammals in the late Holocene. In this vein, Erlandson et al. (2009) review extant
data from shell middens on the northern Channel Islands through most of the
Holocene, finding that Santa Barbara Channel populations fished ‘‘up the food
web,’’ in contrast to modern fishery practices, which have tended toward exploiting
smaller-bodied prey as populations of larger species have declined (Reitz 2004).
Erlandson and colleagues show a dominance of shellfish early on, beginning around
8000 BP, but a shift toward preying on pelagic finfish and pinnipeds in the late
Holocene. They attribute this in part to technological innovations such as the tomol
(sewn plank canoe) and more sophisticated fishhooks. Braje (2010), Erlandson et al.
(2008), and Rick (2011) make similar assertions about long-term trends in the
development of Channel Islands marine economies, couching their analyses in the
perspective that expanding human populations drove expanding diet breadth. What
is significant about this expansion is that it was based on catching bigger-bodied and
therefore, by common measures, higher-return fish (Broughton 1997). But pelagic
species were caught in tomols, which represent large technological investments
(Arnold 2007). They also were a critical part of a sophisticated, trans-channel,
beads-for-terrestrial resources exchange system (Arnold 1991, 1992a; Fauvelle
2013), so pelagic fishing was embedded with other economic (as well as probably
social and political) pursuits. The question of fishing returns and efficiency is thus a
complex one given the costs of the technological investments required to intensively
exploit pelagic fish and the degree to which these investments were offset by
embedding fishing with other activities such as trade.
A similar boat-based pattern developed in late Holocene central and northern
California. Here Hildebrandt and Jones (2002) present ethnographic and archae-
ological evidence that aboriginal coastal California populations intensified their
exploitation of marine resources in the late Holocene by preying more on pinnipeds
who hauled out on offshore islands and sea stacks. This hunting practice required
not only canoes that were costly to manufacture but also entailed a particularly risky
hunting practice, which can be a cost in and of itself (Winterhalder et al. 1999). On
central California’s Monterey Bay, Whitaker and Byrd (2012) focus specifically on

123
Author's personal copy
J Archaeol Res

population pressure as a driver of boat-based fishing and hunting by positing that


increasingly large populations and sociocultural complexity led to the adoption of
boat-based diving for red abalone, a relatively deepwater shellfish, after 950 cal BP.
They hypothesize that the labor associated with boat construction and diving make
this a relatively low-return strategy, but the practice was undertaken anyway for the
relatively large meat package and for the shell, which was used for making
ornaments.
Research in central and northern California also reflects more traditional
optimal foraging approaches to hunter-gatherer intensification. This is most
evident in the San Francisco Bay area, where Broughton and colleagues have
worked since the 1990s. In his analysis of avifauna from Emeryville Shellmound
on the east side of the bay, Broughton (2001) uses, as in most of his work, body
size as a proxy for return rate (itself with its own debates—see Bird et al. 2009),
with smaller birds representing higher-cost, lower-return resources. He identifies a
shift to costly-to-procure cormorants over geese and ducks in the late Holocene.
He is equivocal about how to interpret these preliminary findings but suggests they
indicate intensification and predation pressure on high-ranked avifauna, a pattern
he had already documented at the site with regard to fish and terrestrial fauna
(Broughton 1994a, 1997). More recently, Broughton et al. (2007) present
comparable data from five shellmounds on the San Francisco peninsula, across
the bay west of Emeryville. After accounting for the potential for smaller elements
to decompose over time, they use a ‘‘Goose Index’’ (R NISP of anserines [geese
and large waterfowl] divided by R NISP of anatids [all waterfowl]) to show that
anserines declined in abundance between 2000 and 700 BP. After accounting for
environmental changes that might have caused this decline, they convincingly
argue that human predation was the root of this phenomenon. More importantly,
this research perpetuates the idea that declines in large-bodied prey mark predation
pressure that might be linked to increasing human population density and therefore
intensification s.s.
Like the southern California coast, research farther north also points to variables
that confound the linking of changes in prey type and size to declining foraging
efficiency and by dint, population pressure. A critical and oft-overlooked example is
Kennedy’s (2005) analysis of shellfish foraging on the coast north of San Francisco.
He uses prey and patch choice models, experimental archaeological return rate
studies, and d18O and d13C stable isotope analyses from a diachronic sample of shell
middens to argue that shellfish harvest efficiency declined after about 2000 cal BP,
as documented mainly by a shift to costly-to-procure red abalone. He attributes this
to a shift to more sedentary settlement systems predicated in large part on increasing
population densities and more sedentary settlement patterns inland that forced
subsistence shifts on the coast, similar to the pattern Zangrando (2009) found in
Tierra del Fuego (see also Hildebrandt and Jones [2002] and Whitaker and Byrd
[2014] for assessments of how human population density and circumscription
affected patterning in coastal California’s zooarchaeological assemblages). What
was happening inland in terms of macroscale population-resource dynamics that
force intensification is more important than immediate ecological relationships on
the coast.

123
Author's personal copy
J Archaeol Res

At the local scale, Whitaker (2008) uses data from Punta Gorda Rockshelter on
California’s northwest coast to argue that shellfish reproductive behavior and
harvest technique are critical to midden formation processes. Among the two
methods of harvesting mussels, which reproduce more like r-selected plants than
animals, plucking only large-bodied mussels yields greater immediate returns; this
is a function of body size. But stripping entire populations (which results in a higher
frequency of smaller shells in middens) nets a higher caloric yield over the course of
multiple seasons and is thus a better way to maximize returns over the long run (see
Jones [2003] and Jones and Richman [1995] for more information on mussels,
mussel size, plucking versus stripping, and the relationship of these phenomena to
human population size in coastal California). Key here are three ideas: that prey size
may not indicate decreasing foraging efficiency over the long term, that optimal
immediate returns might be suboptimal when averaged over multiple seasons, and
that population-mediated intensification might entail tactics, like stripping, that
maximize long-term rather than immediate returns.
In a related subject, what drives intensification can be mediated by factors other
than simple population increase. Tushingham (2009; see also Tushingham and
Bettinger 2013) makes this clear in her revisionist analysis of the development of
intensive fishing and foraging systems on California’s northwest coast. Using flaked
and ground stone tool, macrofloral, and faunal data controlled for time by
radiocarbon dates, obsidian hydration measurements, and diagnostic point types, she
postulates that settlement and subsistence changed quite late, around 1250 BP,
marked in part by a shift to people living in semipermanent plank houses and eating
more salmon compared to earlier occupations. She argues that the shift to intensive
exploitation of salmon, a front-loaded resource (it has to be processed before
storage), occurred after northwest California groups had earlier intensified their use
of a back-loaded resource, acorn (which is processed after being stored). Acorn,
though lower return than salmon, was less risky, stored better, and helped contribute
to the development of a system of resource ownership that prevented the
entrenchment of elites who might have appropriated labor associated with large-
scale salmon procurement and storage, in a manner akin to processes seen much
earlier on the Pacific Northwest coast (Bettinger 2015). If she is correct, her
conclusions point to what might be interpreted as de-intensification given the
extremely high costs of processing acorn (McCarthy 1993) and the late switch to
higher-return salmon. This elicits a situation where subsistence choices were
predicated on something of a historical ‘‘founder effect’’—northwest coast
California groups intensified their use of acorn by about 5000 BP, much like the
rest of California (Basgall 1987; Bouey 1987; Wohlgemuth 1996). But the norms
associated with this subsistence focus, hinging in large part on viewing acorn as
private property, prevented the wholesale, intensive adoption of what was likely a
higher-return, more efficient subsistence focus—salmon—at least from a strict
optimal foraging perspective.
Finally, in a somewhat anomalous vein, Todt (2007) surveys the ethnographic
distribution of tobacco (Nicotiana attenuata and N. quadrivalvis) using groups
along the Shasta River in the border country between southern Oregon and
California. He quantifies labor invested in managing and obtaining tobaccos, from

123
Author's personal copy
J Archaeol Res

gathering wild plants, to setting fires to facilitate germination, to sowing, pruning,


and even watering the plants. The most intensive behaviors (sowing, pruning, and
watering) were found among the Western Shasta, which Todd attributes to their
obtaining seeds via trade for the stronger variety, N. quadrivalvis, which is not
endemic to Western Shasta territory but is farther downstream, to the west. The
implication here is that the Western Shasta invested more labor in horticultural
management of tobacco to increase returns on the intensity of the experience of its
alkaloids. Todt’s analysis implies that intensification s.s., marked here by more
labor invested to maximize psychotropic return, can be associated with nonsubsis-
tence goals. This relegates explanations of this aspect of intensification to plant and
human physiology, historical factors, and cultural preference.
In sum, recent research on intensification in California is similar to that in the
Pacific Northwest: it is confounded by middle-range problems of identifying
declining foraging efficiency in the zooarchaeological record and tends toward an
optimal foraging perspective that declining foraging efficiency, marked mainly by
reduced prey or package size as a marker of resource depression, indeed marks
intensification. A newer focus on the effects of boat-based fishing and hunting
suggests, however, that innovation and investments in technology significantly
altered marine mammal and pelagic fish return rates. If rates markedly increased, it
is conceivable that such investments resulted in increased net efficiency, meaning
that in circumstances where there were large enough populations to ensure long use-
life (and hence substantial overall returns) for substantial up-front technological
investments (i.e., capital) like boats, overall economic efficiency may actually
increase, freeing up the labor necessary for the rise of specialists in not only
economic but also perhaps social, political, and even ideological spheres. In the next
section, this dichotomy between investments in labor versus capital is shown to be
of critical import in explaining the processes and drivers of increased yield in
hunter-gatherer economies.

Models and theory

As the preceding review shows, there is a frustrating set of unresolved issues


regarding how intensification is identified in the archaeological record and what
actually initiates intensification processes. Even more fundamental disagreements
are seen in the diversity of opinions concerning what qualifies as intensification
relative to return on labor and capital. Is declining efficiency required to qualify as
intensification? Potential to resolve these issues is found, perhaps less in field and
lab-based empirical studies and more in theory and the development of testable
models, under three main rubrics: zooarchaeological quantification, ecological
modeling, and technological investment models.

Zooarchaeological quantification and foraging goal allocation

One of the main advances in hunter-gatherer zooarchaeology and ethnoarchaeology


over the last several decades is the recognition that different foraging goals can

123
Author's personal copy
J Archaeol Res

produce different zooarchaeological signatures. This results in a middle-range


problem as to how to determine the actual behaviors that produced any give
archaeofaunal assemblage. For example, children, senescent individuals, men
seeking prestige, women foraging near camp, and group hunting techniques all
deposit faunal remains in manners that can be inconsistent with the standard and
intuitive perspective that larger package size indicates higher resource rank (Bird
et al. 2009; Bird and Bliege Bird 2000; Hawkes 1991; Hawkes et al. 1989, 1991;
Lupo and Schmitt 2005; Ugan 2005). Children hunt smaller-bodied prey because
they can do so with minimal technological investment and minimal oversight while
still making a contribution to a group’s caloric intake. Childcare providers forage
closer to camp on often smaller-bodied prey as a way of embedding subsistence
with other pursuits. Prestige seekers might counter these behaviors by actively
seeking out larger-bodied prey despite search times that make such pursuits
suboptimal. Further, Stiner and Munro (2002) show that the way small archaeo-
faunas are classed can have a substantial effect on interpretation of zooarchaeo-
logical data. They argue that categorizing taxa in part on how fast they move rather
than by size alone more clearly shows intensification because prey quickness serves
a proxy for greater pursuit costs, which lower overall return rates.
Recognizing that specific predator and prey behaviors might skew the
zooarchaeological record toward erroneous indications of declining foraging
efficiency and that social rather than subsistence goals might affect zooarchaeo-
logical signatures, Broughton et al. (2011) evaluate the correspondence between the
small quantity (n = 10) of published datasets on experimental resource return rates,
ethnographic data on search, pursuit, and processing costs, and the zooarchaeolog-
ical record of the prehistoric eastern Great Basin. They find that there is indeed a
strong, statistically defensible correlation between prey body size and return rate
and argue that AI based on dividing the sum NISP of large fauna by the sum NISP
for large and small fauna provide very robust means for understanding the
contribution of small, low-ranked fauna versus large, high-ranked fauna to
prehistoric diets. Burger et al. (2005) take this logic one step further, developing
a ‘‘prey as patch model’’ derived from Charnov’s (1976) marginal value theorem.
Putting a new spin on Binford’s (1978) element utility indices and using the novel
idea that each individual animal corresponds to a ‘‘patch’’ in Charnov’s theorem,
they show that more processing produces increasingly marginal returns over time
and that the decision to stop processing (i.e., leave the ‘‘patch’’) is predicated, as in
Charnov’s theorem, when ‘‘in patch’’ returns equal average returns of the
environment at large. In this model, greater degrees of prey butchery thus serve
as markers of more intensive strategies and as proxies for resource depression. In
sum, the implications of these models are not surprising: at the grossest scales and
despite the caveats and exceptions contained in any number of ethnoarchaeological
studies, given animals or faunal elements with similar processing costs and body
composition (e.g., vertebrates), net caloric return, generally speaking, has to equate
with body size and the amount of meat and fat available on any given element. The
problem remains, however, as to whether specific behaviors might confound this
generalization and, perhaps more importantly, whether zooarchaeological measures
can be equated with actual resource return rates.

123
Author's personal copy
J Archaeol Res

Ugan and Bright (2001) directly address this latter question. They develop a way
to use AI as initially generated by Bayham (1979) to actual, experimentally derived
return rates using multi-iterative computer simulations to see if more diverse AI
truly represent the declining foraging efficiencies predicted by Boserupian
intensification models. They suggest that the relationship between AI and return
rate is complex, nonlinear, and very difficult to model and conclude that (1) declines
in the abundance of high ranking prey have only modest effects on overall return
rate, and (2) taphonomic bias and how archaeofauna are quantified can have
significant effects on the curves describing the relationship between AI and return
rate (see also Cannon [2013], Codding et al. [2010], and Lupo [2007] for
discussions regarding AI). Evident in their analysis is the fact that both measures
(AI and return rate) generate marginal returns over multiple iterations, which is not
surprising, and that increased labor input always results in diminishing returns over
time. Their main contribution, other than showing that it is much harder than it
looks to legitimately quantify declines in hunting efficiency per the prey choice
model, is that they reiterate the fact that when using this model, larger-bodied prey
remain in the diet even when human populations undergo resource stress because
larger-bodied, higher-return items are always taken when encountered. If the faunal
record shows a shift to smaller-bodied prey at the expense of large-bodied prey, then
something else has to be going on, such as local extirpation.

Formal ecological models

Geared more to identifying cause than the presence or absence of intensification are
formal ecological models, most based on Binford’s (2001) Frames of Reference.
Hamilton et al. (2007), for example, use Binford’s compilation of hunter-gatherer
ethnological data to assess home range size with reference to metabolic demand,
environmental productivity, and social behaviors. They find that home ranges sizes
vary allometrically rather than isometrically, even when accounting for the effects
of temperature, historical relationships, and other factors that might affect such
ranges. The thrust of this study is that hunter-gatherers with relatively high
population densities live in smaller areas than might be expected of other predators
and omnivores with similar metabolic requirements. They do so because they
critically ‘‘are more efficient [italics added] at extracting materials, energy, and
information from the environment and redistributing those resources to individuals
within societies’’ (Hamilton et al. 2007, p. 4768).
Johnson (2013) extends this perspective. Most of this work is a summary of
conclusions made by Binford: that population density and effective temperature
correlate strongly and predict critical aspects of hunter-gatherer lifeway (especially
storage), that population packing requires some form of increased energetic
extraction from the environment beyond terrestrial hunting, and that increased
extraction can take various pathways (e.g., turning to lower-return terrestrial flora or
shifting to aquatic and marine resources, where available), depending on local
environmental characteristics (see also González-Insuasti and Caballero [2007] for
how this might work for terrestrial plants). In doing so, she notes recent advances in
modeling prehistoric population densities using methods like summed radiocarbon

123
Author's personal copy
J Archaeol Res

probability curves as a means of solving the methodological problem of accounting


for the population side of the ecological equation (Bamforth and Grund 2012;
Williams 2012). The problem is she ultimately relies on the intrinsic rates of
population increase questioned long ago by Cowgill (1975) to explain initial
population packing, meaning populations first entering an area increase as a
function of environmental productivity (prior to changes in diet, technology, or
work strategies). In fact, as Kelly et al. (2013) and Muñoz et al. (2010) recently
showed, hunter-gatherer population growth is indeed intrinsically linked to
environmental productivity. Further, Kelly et al. (2013) and Peros et al. (2010)
demonstrate that long-term hunter-gatherer population growth rates in North
America were quite low, ranging between 0.05 and 0.3% per year (see also Hassan
1981; Pennington 2001), though Peros et al. (2010) identifies a marked increase in
North American populations ca. 2000 cal BP. This means that as a variable, pre-
intensified population densities are really a function of (or even a proxy for)
environmental productivity and that growth rates, except in perhaps short-term and
very late Holocene contexts, may have been too low to initiate intensification
processes. This model, robust in a descriptive sense, is thus much less so in the
processual sense, given its intrinsic environmental determinism and failure to
adequately explain or quantify population growth.
Freeman and Anderies’ (2012) foraging effort model (FEM) uses Binford’s (2001)
data to predict how foraging behaviors should vary given changes in resource density
within the contexts of social relationships, technology, and information exchange. The
FEM unifies the population-level models of Binford (2001) and the individual-level,
competition-based model of Bettinger and Baumhoff (1982) into a single coefficient
representing the costs a forager must pay for the right to exploit a resource in the face of
others who also might wish to do so, operating with the objective of a satisficing, risk-
sensitive goal of meeting minimum caloric thresholds rather than optimizing foraging
returns (Morgan 2009). Critically, as values for this coefficient increase, resource
extraction becomes less efficient due to the higher costs of mediating potential conflict,
which increases, as expected, with increased population density and by inference,
circumscription. This, Freeman and Anderies argue, can lead to investment in land
tenure policies, private ownership, and management practices as ways of reducing
socially mediated uncertainty, which increases efficiency of resource extraction in
population-packed, competitive situations (see Holly 2005 for an example of how
niche differentiation between different social groups might arguably provide an
alternative solution to this problem). A key point derived from this model is a
Malthusian one: that as population levels increase and challenge extant social,
technological, and work strategies, they approach an unstable tipping point where
behaviors and social norms either must change toward a new, higher population
density adaptive peak (Mousoudi 2011) or collapse into a lower-yield, lower-
population density alternative. The FEM does indeed model how population level
processes might play out in the context of competition, but it only accounts for
intensification based on plant resources. Adding mobile, terrestrial animals and
aquatic resources to the equation might be beyond the capabilities of the model as
currently developed, but it would be more consistent with Binford’s (2001) approach,
upon which so much of the FEM is based.

123
Author's personal copy
J Archaeol Res

Technological investment

Technological investment models are in many ways akin to what the FEM predicts
with regard to when investments should be made in resource management rules as
ways of increasing resource extraction efficiency. Ugan et al. (2003) take the initial
stab at this, modeling technological investment as a decision variable akin to how
long to spend in a resource patch per Charnov’s (1976) marginal value theorem.
Their model produces functions describing the marginal return rates for developing
new technologies for fishing, with everything from hook and line (low investment,
low return) to large gill nets (high investment, high return) plotted on the x-axis and
benefit, measured in kg of caught fish, plotted on the y. Though the graphs have little
utility other than making it clear that returns are always marginal, the take-home
message, well taken, is that to offset the time and effort put into larger and more
complex technologies, large technological investments require long handling times.
It only pays to construct a large gill net if it is going to be used often and/or if it will
have a long use life. So Ugan and colleague’s model is really about describing
economies of scale—it really does not pay to make large technological investments
unless one wants to capture a lot of resources or plans on using the technology
repeatedly. Another implication is that some resources have only one or two optimal
solutions. These are depicted by very steep marginal rate curves implying that one
either invests in the technology to capture the resource or one does not; it is all or
nothing, and incremental increases in technological investment cannot increase
returns. This at first appears counterintuitive but indeed makes perfect sense; some
prey like pelagic fish can be obtained only with large technological investments in
seafaring technology (e.g., canoes on the California coast) to bring the predator to
the resource.
The problem with Ugan et al.’s (2003) model is that it lumps multiple technological
investments, each with its own marginal rate of return, into a continuum of marginal
resource return rates. This has the potential to overestimate use times required to make
a technology optimal relative to other alternatives because it pools the cost of
preceding technologies with the additional costs of adding a new technology.
Recognizing this, Bettinger et al. (2006) generate a marginal value model that instead
predicts when switching to another technological class is optimal, again borrowing
from Charnov’s (1976) marginal value theorem. In their model, each technology gets
its own, unique cost-benefit curve rather than being placed on the same curve with
different technologies. Switching to a more costly technology (e.g., a bow and arrow
over an atlatl) will occur when use time is great enough to generate the substantially
higher marginal returns made possible by the new technology. The implication here is
similar to that of Ugan et al. (2003): technological investments are predicted only
when long tool use lives result in qualitatively different— more efficient—but also
temporally extended rates of return (Bamforth 1986; Torrance 1989; but see Fitzhugh
2001). This is critical as both models imply that technological investments may
increase resource extraction efficiency but only when an economy of scale is applied.
Of course, the only time such scales apply is when there is a large enough market (in the
case of hunter-gathers, enough mouths to feed) to make such large capital investments
worth their cost.

123
Author's personal copy
J Archaeol Res

I draw four main points from the preceding. First, there is a dearth of recent
return-rate studies. This limits any discussion and quantitative analysis of
investments in labor, capital, and the economic benefits of such investments. We
have only a vague understanding of the range of variability of return rates using
different technologies for even the most basic hunter-gatherer resources. Second,
although declines in the frequency of large fauna in zooarchaeological assemblages
may indeed indicate overpredation as a function of increasing human population
densities, this generalization is also contingent on taphonomy, differential foraging
goals, and prey response to predation (Emlen 1966). For example, given the
resilience of key prey such as deer (Odocoileus hemionus) to predation (Whitaker
2009), it might be expected that deer encounter rates, and thus the presence of such
fauna in archaeofaunal assemblages, would increase as a function of more hunters
present on the landscape. Third, macroscale ecological models based on population
density and resource abundance, though taking advantage of recent improvements in
reconstructing past population densities and growth rates using radiocarbon summed
probability distributions, are more descriptive than processual, falling into the
circular argument of increasing populations leading to increased resource extraction
but also increased resource extraction leading to increased population density. This
ultimately results in deterministic theory where degrees of environmental produc-
tivity are the sole predictors of the main aspects of hunter-gatherer economic
behavior. Lastly, technological investment models show that innovations can result
in increased efficiency, but only in the context of population densities large enough
to generate the economies of scale necessary to pay for such investments. The
implication is that population may mediate both increases and decreases in overall
economic efficiency in the broader context of increased production but also that
increased efficiency based on capital investments provide the wherewithal for a
portion of increasingly large populations to be freed from solely subsistence-related
pursuits.

Discussion

The preceding deliberately broad sample of hunter-gatherer intensification research


over the last decade suggests considerable stasis in terms of resolving many of the
empirical and ecological issues initially entailed by Boserup’s model and its later
applications to hunter-gatherer archaeology. Though some progress has been made
in identifying intensification (mainly with archaeofaunas) and modeling modes of
intensification (s.l.) within the context of population density and social structure,
determining the root cause(s) of increased production are still hampered to a large
degree by the diversity of opinions as to what exactly intensification entails.
Critically, as explained below, this is much more than a semantic debate; it is a
theoretical divide. Resolving this division is consequently of the utmost importance
in moving the discussion forward to explain the development of hunter-gatherer
economies.
Stable isotope studies and settlement data show some promise for identifying
intensification (Hildebrandt and Jones 2002; Jerardino 2012; Johnson 2013; Johnson

123
Author's personal copy
J Archaeol Res

and Hard 2008; Kennedy 2005). The development of models linking zooarchae-
ological quantification with actual return rates is equally significant (Ugan and
Bright 2001), but doing so is hampered by how little experimental data there are on
resource return rates (Broughton et al. 2011). Though not reviewed at length here,
there has been some progress made in the identification of the other side of the
population pressure equation, where past population densities are estimated using a
variety of proxies—diachronic changes in site or feature frequencies (Eerkens 2008;
Eerkens and Rosenthal 2002; Jerardino 2012; Thoms 2009), environmental
productivity (Binford 2001), and especially summed radiocarbon probability
distributions (Williams 2012).
More important, however, is the recognition that foraging goal variability, prey
habitat preference, prey behavioral characteristics, and environmental change can
all affect archaeofaunal assemblages, by far the most common metric of prehistoric
subsistence. These factors can either mimic population-mediated declining foraging
efficiency or, especially in the context of prestige-related foraging goals, occlude the
effects of population pressure and its effect on the zooarchaeological record. What
this means is that subsistence diversification, most often linked to population-
mediated intensification, and subsistence specialization, sometimes seen as an
alternative means to increase yield (Munro and Atici 2009b; Thurston and Fisher
2007a, b), may be caused by circumstances other than burgeoning human
populations. Most zooarchaeologists, however, appear quite cognizant of these
confounding factors and account for many of them, especially environmental
characteristics, environmental change, and taxa behavior in their analyses. In these
contexts, the resource side of the population pressure equation necessitates not only
strict measures of zooarchaeological quantification but also interpretation of these
measures per relevant social and environmental contexts.
Far less headway has been made in explaining intensification; the majority of
research either addresses the topic indirectly or falls into the trap entailed by the
very long running debates over whether environmental productivity, innovation, or
population increase initiate increases in economic output. What came first:
innovation or increased labor leading to increased population densities; intrinsic
rates of population increase that eventually challenged environmental carrying
capacities, necessitating some sort of adaptive response; or social factors such as
changes in the conception of private property or competitive gift giving that led to
efforts to increase yield?
In the Paleolithic, these questions hinge on whether a greater focus on marine
niches gave anatomically modern humans an adaptive advantage over Archaic
Homo (Hockett and Haws 2005) and whether intensification occurred during the UP
or later in the terminal Pleistocene. Similar questions pertain during the TPEH and
the BSR, with recent evidence for Late Natufian population decline during the YD
calling into question the notion that population pressure led to plant and animal
domestication at the dawn of the Holocene (Munro 2004). Research across North
America not only points to how increasing human population densities may have
driven economic intensification (Eerkens 2008; Eerkens and Rosenthal 2002;
Johnson 2013; Thoms 2009; Yu 2006) but also how mediating risk and satisficing-
based decision making might have resulted in similarly intensive economic

123
Author's personal copy
J Archaeol Res

strategies (Morgan 2009; White 2013). Holocene coastal-focused research points


not only to the ways that coastal people exerted predation and harvest pressure on
marine resources (Reitz 2004) but also how long-term management may have
helped sustain larger populations over the long term (Cannon and Burchell 2009). It
also shows how technological investments may have facilitated the exploitation of
larger-bodied, higher-return (but also perhaps more risky) prey (Erlandson et al.
2009). Research especially along the west coast of North America most clearly
exhibits this conundrum, with no resolution to long-running debates over whether
environmental decline, increasing populations, or early shifts to abundant, high-
yield marine taxa and affiliated technologies such as nets, fishhooks, and storage led
to the large, complex hunter-gatherer societies who lived in the region in the late
Holocene (Ames 2003; Arnold 2004; Fitzhugh 2003b).
This compendium of empirical research suggests that there may indeed be many
(i.e., multilinear) pathways to increases in economic output and the development of
‘‘complexity’’ in hunter-gatherer economies, arguably to the point of entailing
historical rather than evolutionary processes (Boone and Smith 1998). But
ecological modeling drawing mainly on Binford (2001) develops defensible,
descriptive evolutionary generalizations as to what predicts which pathway is taken,
whether it is based on increased exploitation of terrestrial fauna, terrestrial plants,
marine and riparian resources, or domesticates such as maize (Johnson and Hard
2008). The contribution of these models is that they show very strong correlations
between environmental characteristics, seasonality, population density, and some
technological behaviors, especially storage, meaning something other than local
historical processes were likely at play in their development. Put simply, higher-
density hunter-gatherer populations living in mid-latitude, seasonal environments
tend to store food and intensify their economic behavior based on what types of
resources are available. But for the most part, we already knew this (Binford 1980,
1990; Brenton 1988; Keeley 1988; Rocek and Bar-Yosef 1998) and are still left with
the conundrum of explaining process absent historical circumstance.
Attempts to insert process into the equation tend to either rely on intrinsic rates of
population increase to explain population packing/pressure or fall back on models
already developed using empirical data from places such as the Pacific Northwest,
where initial population packing, circumscription, technological improvements, and
the like take primacy in explaining intensification processes (Fitzhugh 2003b;
Johnson 2013). So what explains intrinsic rates of population increase? Environ-
mental abundance (King 1990)? Technological or environmental change (Raab and
Larson 1997)? Changes in social norms (Bettinger 1999)? In a very real sense, we
are right back where we started. Surely in specific cases the actual processes leading
to increases in economic production worked in dynamic, feedback-mediated,
stepwise or cyclical manners, with changes in population pressure leading to
changes in technology, work strategies, or social norms and vice versa. But this type
of thinking occludes identification of ultimate cause (Krebs and Davies 1997) and
has the potential of once again falling into the trap of particularistic description
rather than theoretical explanation (Gremillion et al. 2014). Further, it neglects the
fact that in some cases, fundamental changes in hunter-gatherer behavior (and their
associated economies) may have been quite rapid (for example the adoption of the

123
Author's personal copy
J Archaeol Res

C
B
A

Time

Fig. 2 Intensification of same resource or patch given marginal return rates. After Time A, which
represents a hypothetical pre-intensified economy, diminishing returns are indicated by all times to the
right of ‘‘A’’ (the shaded area below the curve). Past this point, the per-unit cost (in time) of each
additional unit of benefit is greater than per-unit cost prior to Time A, indicating declining efficiency

Search + handling
Search + handling (dashed)
Time per unit of energy acquired

Optimal
point
)
lid
Optimal so
point e(
tim
g
lin
nd Se
Ha arc
ht
im
e
e
tim
g
lin
nd
Sea
rch Ha
tim
T1 T2
e

Higher
ranked A B Lower Higher
ranked ranked A B Lower
ranked
Ranked resources Ranked resources

Fig. 3 Intensification of new resource or patch. Per the diet breadth model (MacArthur and Pianka 1966;
see also Bettinger 1991, fig. 4.1), the graph on the left (‘‘T1’’) depicts a strategy where only higher-ranked
resources (Resource A) are taken because they optimize the amount of time per unit of energy acquired.
The graph on the right (‘‘T2’’) depicts the inclusion of lower-ranked resources (‘‘B’’) in the diet, which
increases search time and moves the optimal solution to the right of the graph. Note the increase in time
per unit of energy acquired to include Resource B in T2, an indicator of reduced efficiency

bow and arrow) and that evolutionary trajectories toward different adaptive peaks
(say between agriculture and the intensive foraging systems in aboriginal
California) are arguably resistant to piecemeal change (Bettinger 2009). In short,

123
Author's personal copy
J Archaeol Res

i
B
Increase in

switching to

A
Tech. B at UT1

Switching
point
UT1

Use time Manufacturing time

Fig. 4 Innovation/technological investment. Switching to Technology B (per Bettinger et al. 2006,


fig. 3) results in marked increases in benefit relative to Technology A, but at a cost of substantially
increased manufacturing time (the right side of the x-axis). The intersection of the dotted line (i) with the
x-axis at Point UT1 is the switching point, in use time, where it becomes more efficient to switch to
Technology B (the slope of the dotted line [i] is determined by its tangential intersection with the
marginal return rate curves for both Technologies A and B). At all use times to the left of this point (the
shaded portion of the graph), Technology B is more efficient than Technology A. Long use times thus
offset the increased costs of technological investments because of the substantially greater returns
garnered by employing a more effective technology for a longer period of time

socioeconomic systems work as a package deal, where shifting from one package to
another has potentially deleterious costs and where changes in one part of the
system, for instance to a storage economy, necessitates not only technological and
behavioral shifts but also fundamental reorganization of work strategies, social
norms, and conceptions of private property (Bettinger 2015).
The diversity of opinions, however, as to what exactly causes increased
productivity in hunter-gatherer economies is informed by (and may be resolvable by
addressing) the critical disjuncture in the very definition of what intensification
entails. On the one side is the strict Boserupian perspective that labor investment
drives the engine of economic output, but that this comes at a cost of declining
efficiency. From this viewpoint, continuing to exploit the same set of resources with
more labor results in increasingly marginal returns (Fig. 2). Adding new, lower-
return resources or resource patches to the foraging itinerary also results in reduced
net returns, though by definition, increased gross returns (Fig. 3). From this
perspective, value is generated via increased labor investment, in general accord
with both classical and Marxist economic perspectives (Marx 2010; Smith 2009).
This is intensification in its strictest sense, driven by population-resource
imbalances.
Providing contrast (and increasingly common in the archaeological literature on
hunter-gatherers) is the perspective that intensification entails any means (special-
ization, diversification, innovation, etc.) that increases economic output, including

123
Author's personal copy
J Archaeol Res

Table 1 Modes of increases in productivity, postulated changes in efficiency, and hypothesized contexts
in hunter-gatherer economies
Mode Change in Ecological context Social context
efficiency

Intensification Decrease Reduced environmental Circumscription; resource


(s.s.) productivity; increased privatization; aggrandizing
population density behaviors
(population pressure)
Diversification Decrease Population pressure Circumscription; resource
privatization
Increase Technological innovation; Division of labor
change in work strategies
Innovation Decrease Short use-life; low yields; Theft?
environmental degradation
Increase Economies of scale Larger populations
Specialization Decrease Reduced environmental Aggrandizing behaviors
productivity; increased
population density
(population pressure)
Increase Environmental amelioration; Aggrandizing behaviors
technological innovation;
change in work strategies

those that increase efficiency. This is the schism Butler and Campell (2004) allude
to, seen often in research in the Pacific Northwest (Ames 2003; Matson 1983) but
present or implied in much of the literature reviewed here. As I have shown,
specialization, diversification, and innovation—all means of increasing productiv-
ity—can result from different social, demographic, and ecological contexts and may
not necessarily have to arise only in the context of population pressure. As Bettinger
et al. (2006) and Ugan et al. (2003) convincingly show, however, investments in
technology—i.e., capital investments—can result in qualitative increases in return
and increased efficiency as long as there are the requisite economies of scale
(meaning more people) present to offset the costs of the initial investment (Fig. 4).
The fundamental question then is whether investments in labor or capital drive
increases in economic output in hunter-gatherer societies and what contexts drive
these different types of investments. On the one hand, Ames (2005) is right: both
investments in capital and labor can result in increased productivity. But each also
entails very different processes, each with its own costs, returns, and contexts under
which it might succeed or fail (Table 1). Technological investments only increase
net efficiency when an economy of scale is applied. Do large populations already
have to be in place to provide this context? Diversification may represent means of
mediating risk in light of uncertainty as well as outright intensification s.s. Absent
technological change or changes in work strategies, simply working harder or longer
necessarily entails marginal return rates over time and thus declining efficiency.
What are the economic and perhaps social incentives behind accepting diminishing
returns? In economic and evolutionary contexts, the question then becomes under
what contexts do increased investments in labor versus capital provide an adaptive

123
Author's personal copy
J Archaeol Res

advantage to their practitioners? The problem that has to be dealt with before we
turn attention to the environmental, demographic, or social causes behind increases
in hunter-gatherer economic productivity is that of identifying the means by which
increased yield was accomplished. Conflating fundamentally different behaviors
that result in either increased or decreased efficiency as ‘‘intensification’’ thus
occludes process and the contexts under which different pathways to increased
productivity occur (Thurston and Fisher 2007a, b). The concept of efficiency is thus
critical to the goal of understanding process, meaning specifying exactly what type
of pathway led to increased economic output in specific cases is the essential first
step toward untangling the long-running circular debates over hunter-gatherer
economic evolution.

Conclusion

In the title of this review, I ask if archaeologists studying hunter-gatherers have


succeeded in identifying and explaining intensification and whether this can be
linked to supporting the larger populations typically associated with ‘‘complex’’
hunter-gatherers. Progress has clearly been made, but the answer at this point has to
be a qualified ‘‘not quite.’’ There are many instances, particularly from research in
California and the Great Basin, where evidence for declining foraging efficiency has
been derived from archaeofaunal assemblages and arguably linked to increasing
population densities. But these efforts are confounded (but certainly not discounted)
by several key problems: generating meaningful estimates of past population
densities, our poor understanding of resource return rates, especially when using
different technologies or work strategies, and the effects of foraging goal variability
on the zooarchaeological record. Clearly, more robust means of estimating past
population densities are critical to solving the first problem, perhaps by building on
researchers who employ radiocarbon-summed probability distributions as popula-
tion proxies (Bamforth and Grund 2012; Kelly et al. 2013; Peros et al. 2010;
Williams 2012), but with more robust mathematical models to account for recovery
and taphonomic biases. A substantial challenge but also a potential means to making
this method more robust is increasing sample size by including chronometric
datasets other than radiocarbon (e.g., OSL) into the statistical packages that generate
these distributions. Putting down our trowels for a moment and conducting more
(and more robust) experimental and ethnoarchaeological research would clearly
help solve the latter two problems, especially by addressing how wider arrays of
technologies (and their associated investments) affect resource return rates.
Explaining why increasing economic output occurs in hunter-gatherer economies
entails more fundamental problems. These hinge mainly on the circular ecological,
demographic, and social logic behind such explanations, where increases in
economic output are either causes or effects of population pressure, environmental
characteristics, or changes in social norms and leadership objectives. At this point,
resolving these issues necessitates unpacking the very concept of intensification.
This requires us to be very specific about the different processes and contexts under
which increases in economic output might be expected to occur. If intensification is

123
Author's personal copy
J Archaeol Res

taken to mean any method used to increase economic yield, it has very little
descriptive and almost no explanatory power. If it is employed in its Boserupian
sense, as entailing declining efficiency, this opens up the floor to much more
nuanced thinking about process, where diversification, specialization, and innova-
tion are alternative and perhaps fundamentally different means of increasing output
in evolving hunter-gatherer economies.

Acknowledgments I thank Robert Kelly, Brian Codding, Virginia Butler and three additional
anonymous reviewers for their insightful and extremely valuable commentary on the draft of this article.
Likewise I thank Gary Feinman for his editorial suggestions and guidance through the entire process,
from proposal to press. Any errors or omissions, however, are my own.

References

Allen, J., and O’Connell, J. F. (2008). Getting from Sunda to Sahul. In Clarke, G., Leach, F., and
O’Connor, S. (eds.), Islands of Inquiry: Colonization, Seafaring and the Archaeology of Maritime
Landscapes, ANU E Press, Canberra, Australia, pp. 31–46.
Álvarez-Fernández, E. (2011). Humans and marine resource interaction reappraised: Archaeofauna
remains during the late Pleistocene and Holocene in Cantabrian Spain. Journal of Anthropological
Archaeology 30: 327–343.
Ames, K. M. (1981). The evolution of social ranking on the Northwest Coast of North America. American
Antiquity 46: 789–805.
Ames, K. M. (1985). Hierarchies, stress, and logistical strategies among hunter-gatherers in northwestern
North America. In Price, T. D., and Brown, J. (eds.), Prehistoric Hunter-Gatherers: The Emergence
of Cultural Complexity, Academic Press, Orlando, FL, pp. 155–180.
Ames, K. M. (1994). The Northwest Coast: Complex hunter-gatherers, ecology, and social evolution.
Annual Review of Anthropology 23: 209–229.
Ames, K. M. (2003). The Northwest Coast. Evolutionary Anthropology 12: 19–33.
Ames, K. M. (2005). Intensification of food production on the Northwest Coast and elsewhere. In Duer,
D., and Turner, N. J. (eds.), Keeping It Living: Traditions of Plant Use and Cultivation on the
Northwest Coast of North America, University of British Columbia Press, Vancouver, pp. 67–100.
Ames, K. M., and Maschner, H. D. (1999). Peoples of the Northwest Coast: Their Archaeology and
Prehistory, Thames and Hudson, London.
Arnold, J. E. (1991). Transformation of a regional economy: Sociopolitical evolution and the production
of valuables in southern California. Antiquity 65: 953–962.
Arnold, J. E. (1992a). Complex hunter-gatherer-fishers of prehistoric California: Chiefs, specialists, and
maritime adaptations of the Channel Islands. American Antiquity 57: 60–84.
Arnold, J. E. (1992b). Cultural disruption and the political economy in Channel Islands prehistory. In
Jones, T. (ed.), Essays on the Prehistory of Maritime California, Center for Archaeological Research
at Davis, Davis, CA, pp. 129–144.
Arnold, J. E. (1993). Labor and the rise of complex hunter-gatherers. Journal of Anthropological
Archaeology 12: 75–119.
Arnold, J. E. (1996). The archaeology of complex hunter-gatherers. Journal of Archaeological Method
and Theory 3: 77–126.
Arnold, J. E. (2004). Foundations of Chumash Complexity, Perspectives in California Archaeology, Vol.
7, Cotsen Institute of Archaeology, University of California, Los Angeles.
Arnold, J. E. (2007). Credit where credit is due: The history of the Chumash oceangoing plank canoe.
American Antiquity 72: 169–209.
Arnold, J. E., Colton, R. H., and Pletka, S. (1997). Contexts of cultural change in insular California.
American Antiquity 62: 300–318.
Atici, L. (2009). Specialization & diversification: Animal exploitation strategies in the terminal
Pleistocene, Mediterranean Turkey. Before Farming 1: 1–17.
Aura Tortosa, J. E., Bonilla, V. V., Ripoll, M. P., Valle, R. M., and Calatayud, P. G. (2002a). Big game
and small prey: Paleolithic and Epipaleolithic economy from Valencia (Spain). Journal of
Archaeological Method and Theory 9: 215–268.

123
Author's personal copy
J Archaeol Res

Aura Tortosa, J. E., Jordá Pardo, J. F., Pérez Ripoll, M., Rodrigo Garcı́a, M. J., Badal Garcı́a, E., and
Guillem Calatayud, P. (2002b). The far south: The Pleistocene–Holocene transition in Nerja Cave
(Andalucı́a, Spain). Quaternary International 93–94: 19–30.
Bamforth, D. B. (1986). Technological efficiency and tool curation. American Antiquity 51: 38–50.
Bamforth, D. B., and Grund, B. (2012). Radiocarbon calibration curves, summed probability
distributions, and early Paleoindian population trends in North America. Journal of Archaeological
Science 39: 1768–1774.
Bar-Yosef, O. (1998). The Natufian culture in the Levant, threshold to the origins of agriculture.
Evolutionary Anthropology 6: 159–177.
Bar-Yosef, O., and Belfer-Cohen, A. (1989). The origins of sedentism and farming communities in the
Levant. Journal of World Prehistory 3: 447–498.
Basgall, M. E. (1987). Resource intensification among hunter-gatherers: Acorn economies in prehistoric
California. Research in Economic Anthropology 9: 21–52.
Baumhoff, M. A. (1963). Ecological determinants of aboriginal California populations. University of
California Publications in American Archaeology and Ethnology 49: 155–236.
Bayham, F. (1979). Factors influencing the Archaic pattern of animal exploitation. Kiva 44: 219–235.
Beaton, J. M. (1991). Extensification and intensification in central California prehistory. Antiquity 65:
946–952.
Bender, B. (1978). Gatherer-hunter to farmer: A social perspective. World Archaeology 10: 204–222.
Bender, B. (1981). Gatherer-hunter intensification. In Sheridan, A., and Bailey, G. (eds.), Economic
Archaeology: Towards an Integration of Ecological and Social Approaches, BAR International
Series vol. 96, Archaeopress, Oxford, pp. 149–157.
Bender, B. (1985). Emergent tribal formation in the American mid-continent. American Antiquity 50:
52–62.
Bettinger, R. L. (1991). Hunter-Gatherers, Archaeological and Evolutionary Theory, Plenum Press, New
York.
Bettinger, R. L. (1999). What happened in the Medithermal. In Beck, C. (ed.), Models for the Millennium:
Great Basin Anthropology Today, University of Utah Press, Salt Lake City, pp. 62–74.
Bettinger, R. L. (2001). Holocene hunter-gatherers. In Feinman, G., and Price, T. D. (eds.), Archaeology
at the Millennium, Kluwer/Plenum, New York, pp. 137–195.
Bettinger, R. L. (2009). Macroevolutionary theory and archaeology: Is there a big picture? In Prentiss, A.
M., and Chatters, H. C. (eds.), Macroevolution in Human Prehistory: Evolutionary Theory and
Processual Archaeology, Springer, New York, pp. 275–295.
Bettinger, R. L. (2015). Orderly Anarchy: Sociopolitical Evolution in Aboriginal California, University
of California Press, Berkeley (in press).
Bettinger, R. L., and Baumhoff, M. A. (1982). The Numic spread: Great Basin cultures in competition.
American Antiquity 47: 485–503.
Bettinger, R. L., Malhi, R., and McCarthy, H. (1997). Central place models of acorn and mussel
processing. Journal of Archaeological Science 24: 887–899.
Bettinger, R. L., Winterhalder, B., and McElreath, R. (2006). A simple model of technological
intensification. Journal of Archaeological Science 33: 538–545.
Betts, M. W., and Friesen, T. M. (2004). Quantifying hunter-gatherer intensification: A zooarchaeological
case study from Arctic Canada. Journal of Anthropological Archaeology 23: 357–384.
Bicho, N., and Haws, J. (2008). At the land’s end: Marine resources and the importance of fluctuations in
the coastline in the prehistoric hunter-gatherer economy of Portugal. Quaternary Science Reviews
27: 2166–2175.
Binford, L. R. (1968). Post Pleistocene adaptations. In Binford, S. R., and Binford, L. R. (eds.), New
Perspectives in Archaeology, Aldine, Chicago, pp. 313–341.
Binford, L. R. (1978). Nunamiut Ethnoarchaeology, Academic Press, New York.
Binford, L. R. (1980). Willow smoke and dog’s tails: Hunter-gatherer settlement systems and
archaeological site formation. American Antiquity 45: 4–20.
Binford, L. R. (1990). Mobility, housing, and environment: A comparative study. Journal of
Anthropological Research 46: 119–152.
Binford, L. R. (2001). Constructing Frames of Reference: An Analytical Method for Archaeological
Theory Building Using Hunter-Gatherer and Environmental Data Sets, University of California
Press, Berkeley.
Bird, D. W., Bird, R. B., and Codding, B. (2009). In pursuit of mobile prey: Martu hunting strategies and
archaeofaunal interpretation. American Antiquity 74: 3–29.

123
Author's personal copy
J Archaeol Res

Bird, D. W., and Bliege Bird, R. (2000). The ethnoarchaeology of juvenile foragers: Shellfishing
strategies among Meriam children. Journal of Anthropological Archaeology 19: 461–476.
Bird, D. W., and O’Connell, J. F. (2006). Behavioral ecology and archaeology. Journal of Archaeological
Research 14: 143–188.
Bliege Bird, R. (2007). Fishing and the sexual division of labor among the Meriam. American
Anthropologist 109: 442–451.
Boone, J. L., and Smith, E. A. (1998). Is it evolution yet? A critique of evolutionary archaeology. Current
Anthropology 39: 141–173.
Boserup, E. (1965). The Conditions of Agricultural Growth, Aldine, Chicago.
Boserup, E. (1970). Present and potential food production in developing countries. In Zelinsky, W.,
Kosinki, L. A., and Prothero, R. M. (eds.), Geography of a Crowding World: A Symposium on
Population Pressure Upon Physical and Social Resources in the Developing Lands, Oxford
University Press, New York, pp. 100–114.
Boserup, E. (1981). Population and Technological Change: Study of Long-Term Trends, University of
Chicago Press, Chicago.
Bouey, P. D. (1979). Population pressure and agriculture in Owens Valley. Journal of California and
Great Basin Anthropology 1: 162–170.
Bouey, P. D. (1987). The intensification of hunter-gatherer economies in the southern north coast ranges
of California. Research in Economic Anthropology 9: 53–101.
Braidwood, R. J. (1963). Prehistoric Men, 6th ed., Popular Series, Anthropology No. 37, Chicago
Museum of Natural History, Chicago.
Braje, T. J. (2007). Archaeology, Human Impacts, and Historical Ecology on San Miguel Island,
California, Ph.D. dissertation, Department of Anthropology, University of Oregon, Eugene.
Braje, T. J. (2010). Modern Oceans, Ancient Sites: Archaeology and Marine Conservation on San Miguel
Island, California, University of Utah Press, Salt Lake City.
Braje, T. J., Kennett, D. J., Erlandson, J. M., and Culleton, B. J. (2007). Trans-Holocene subsistence
changes and human impacts on a marine ecosystem on Santa Rosa Island, California. American
Antiquity 72: 735–756.
Brenton, B. P. (1988). The seasonality of storage. In Huss-Ashmore, R., Curry, J. J., and Hitchcock, R. K.
(eds.), Coping with Seasonal Constraints, Masca, University Museum, University of Pennsylvania,
Philadelphia, pp. 45–54.
Broughton, J. M. (1994a). Declines in mammalian foraging efficiency during the late Holocene, San
Francisco Bay, California. Journal of Anthropological Archaeology 13: 371–401.
Broughton, J. M. (1994b). Late Holocene resource intensification in the Sacramento Valley: The
vertebrate evidence. Journal of Archaeological Science 21: 501–514.
Broughton, J. M. (1997). Widening diet breadth, declining foraging efficiency, and prehistoric harvest
pressure: Ichthyofaunal evidence from the Emeryville Shellmound, California. Antiquity 71:
845–862.
Broughton, J. M. (2001). Resource intensification and late Holocene human impacts on Pacific coast bird
populations: Evidence from the Emeryville Shellmound avifauna. In Hunt, T. L., Lipo, C. P., and
Sterling, S. L. (eds.), Posing Questions for a Scientific Archaeology, Greenwood, Westport, CT,
pp. 251–278.
Broughton, J. M., and Bayham, F. E. (2003). Showing off, foraging models, and the ascendance of large-
game hunting in the California Middle Archaic. American Antiquity 68: 783–789.
Broughton, J. M., Cannon, M. D., Bayam, F. E., and Byers, D. A. (2011). Prey body size and ranking in
zooarchaeology: Theory, empirical evidence and applications from the northern Great Basin.
American Antiquity 76: 403–428.
Broughton, J. M., and Grayson, D. K. (1993). Diet breadth, adaptive change, and the White Mountains
faunas. Journal of Archaeological Science 20: 331–336.
Broughton, J. M., Mullins, D., and Ekker, T. (2007). Avian resource depression or intertaxonomic
variation in bone density? A test with San Francisco bay avifaunas. Journal of Archaeological
Science 34: 374–391.
Burger, O., Hamilton, M. J., and Walker, R. (2005). The prey as patch model: Optimal handling of
resources with diminishing returns. Journal of Archaeological Science 32: 1147–1158.
Butler, V. (2000). Resource depression on the Northwest Coast of North America. Antiquity 74: 649–661.
Butler, V., and Campell, S. K. (2004). Resource intensification and resource depression in the Pacific
Northwest of North America: A zooarchaeological review. Journal of World Prehistory 18:
327–405.

123
Author's personal copy
J Archaeol Res

Byers, D. A., and Broughton, J. M. (2004). Holocene environmental change, artiodactyl abundances, and
human hunting strategies in the Great Basin. American Antiquity 69: 235–255.
Byers, D. A., Smith, C. S., and Broughton, J. M. (2005). Holocene artiodactyl population histories and
large game hunting in the Wyoming Basin, USA. Journal of Archaeological Science 32: 125–142.
Caldwell, J. R. (1958). Trend and Tradition in the Prehistory of the Eastern United States, American
Anthropological Association, Menasha, WI.
Campbell, S. K., and Butler, V. L. (2010). Archaeological evidence for resilience of Pacific Northwest
salmon populations and the socioecological system over the last *7,500 years. Ecology and Society
15: 17. http://www.ecologyandsociety.org/vol15/iss1/art17/. Accessed 15 May 2013.
Cannon, A., and Burchell, M. (2009). Clam growth-stage profiles as a measure of harvest intensity and
resource management on the central coast of British Columbia. Journal of Archaeological Science
36: 1050–1060.
Cannon, M. D. (2013). NISP, bone fragmentation, and the measurement of taxonomic abundance. Journal
of Archaeological Method and Theory 20: 397–419.
Carneiro, R. L. (1970). A theory of the origin of the state. Science 169: 733–738.
Charnov, E. L. (1976). Optimal foraging, the marginal value theorem. Theoretical Population Biology 9:
129–136.
Codding, B. F., Bird, D. W., and Bliege Bird, R. (2010). Interpreting abundance indices: Some
zooarchaeological implications of Martu foraging. Journal of Archaeological Science 37:
3200–3210.
Codding, B. F., Bird, R. B., and Bird, D. W. (2011). Provisioning offspring and others: Risk-energy trade-
offs and gender differences in hunter-gatherer foraging strategies. Proceedings of the Royal Society,
B 278: 2502–2509.
Codding, B. F., Burd, D. W., and Jones, T. L. (2012). A land of work: Foraging behavior and ecology. In
Jones, T. L., and Perry, J. E. (eds.), Contemporary Issues in California Archaeology, Left Coast
Press, Walnut Creek, CA, pp. 115–131.
Codding, B. F., and Jones, T. L. (2007). Man the showoff? Or the ascendance of a just-so-story: A
comment on recent applications of costly signaling theory in American archaeology. American
Antiquity 72: 349–357.
Cohen, M. N. (1977). The Food Crisis in Prehistory: Overpopulation and the Origins of Agriculture, Yale
University Press, New Haven, CT.
Cortés-Sánchez, M., Morales-Muñiz, A., Simón-Vallejo, M. D., Bergadà-Zapata, M. M., Delgado-
Huertas, A., López-Garcı́a, P., López-Sáez, J. A., Lozano-Francisco, M. C., Riquelme-Cantal, J. A.,
Roselló-Izquierdo, E., Sánchez-Marco, A., and Vera-Peláez, J. L. (2008). Palaeoenvironmental and
cultural dynamics of the coast of Málaga (Andalusia, Spain) during the upper Pleistocene and early
Holocene. Quaternary Science Reviews 27: 2176–2193.
Cowgill, G. L. (1975). On causes and consequences of ancient and modern population changes. American
Anthropologist 77: 505–525.
Croes, D. R., and Hackenberger, S. (1988). Hoko River archaeological complex: Modeling prehistoric
Northwest Coast economic evolution. In Isaac, B. L. (ed.), Prehistoric Economies of the Pacific
Northwest Coast, JAI Press, Greenwich, CT, pp. 19–85.
Cuthrell, R. Q., Striplen, C., Hylkema, M., and Lightfoot, K. G. (2012). A land of fire: Anthropologenic
burning on the central coast of California. In Jones, T. L., and Perry, J. E. (eds.), Contemporary
Issues in California Archaeology, Left Coast Press, Walnut Creek, CA, pp. 153–174.
Dennell, R. (2009). The Palaeolithic Settlement of Asia, Cambridge University Press, Cambridge.
Earle, T. K. (1980). A model of subsistence change. In Earle, T., and Christenson, A. (eds.), Modeling
Change in Prehistoric Subsistence Economies, Academic Press, New York, pp. 1–30.
Eerkens, J. W. (2003). Sedentism, storage, and the intensification of small seeds: Prehistoric
developments in Owens Valley, California. North American Archaeologist 24: 281–309.
Eerkens, J. W. (2004). Privatization, small-seed intensification, and the origins of pottery in the western
Great Basin. American Antiquity 69: 653–670.
Eerkens, J. W. (2008). Rock-ring features on the shores of Owens Lake and implications for prehistoric
geophyte processing and storage. Proceedings of the Society for California Archaeology 21:
180-184.
Eerkens, J. W., King, J., and Wohgelmuth, E. (2004). The prehistoric development of intensive green-
cone piñon processing in eastern California. Journal of Field Archaeology 29: 18–27.

123
Author's personal copy
J Archaeol Res

Eerkens, J. W., and Rosenthal, J. S. (2002). Transition from geophyte to seed processing: Evidence for
intensification from thermal features near China Lake, northern Mojave Desert. Pacific Coast
Archaeological Society Quarterly 38: 19–36.
Ellis, E. C., Kaplan, J. O., Fuller, D. Q., Vavrus, S., Klein Goldewijk, K., and Verburg, P. H. (2013). Used
planet: A global history. Proceedings of the National Academy of Sciences 110: 7978–7985.
Elston, R. G., Dong, G., and Zhang, D. (2011). Late Pleistocene intensification technologies in northern
China. Quaternary International 242: 401–415.
Emlen, J. M. (1966). The role of time and energy in food preference. The American Naturalist 100:
611–617.
Eren, M. I. (2012). On the Younger Dryas as a causal determinant of prehistoric hunter-gatherer culture
change. In Eren, M. I. (ed.), Hunter-Gatherer Behavior, Human Response During the Younger
Dryas, Left Coast Press, Walnut Creek, CA, pp. 11–24.
Erickson, C. L. (1993). The social organization of prehispanic raised field agriculture in the Lake Titicaca
basin. In Scarborough, V., and Isaac, B. (eds.), Economic Aspects of Water Management in the New
World, JAI Press, Greenwich, CT, pp. 367–424.
Erlandson, J. M. (2001). The archaeology of aquatic adaptations: Paradigms for a new millennium.
Journal of Archaeological Research 9: 287–350.
Erlandson, J. M., Rick, T. C., and Braje, T. J. (2009). Fishing up the food web? 12,000 years of maritime
subsistence and adaptive adjustments on California’s Channel Islands. Pacific Science 63: 711–724.
Erlandson, J. M., Rick, T. C., Braje, T. J., Steinberg, A., and Vellanoweth, R. L. (2008). Human impacts
on ancient shellfish: A 10,000 year record from San Miguel Island, California. Journal of
Archaeological Science 35: 2144–2152.
Fauvelle, M. (2013). Evaluating cross-channel exchange in the Santa Barbara region: Experimental data
on acorn processing and transport. American Antiquity 78: 790–798.
Fitzhugh, B. (2001). Risk and invention in human technological evolution. Journal of Anthropological
Archaeology 20: 125–167.
Fitzhugh, B. (2002). Residential and logistical strategies in the evolution of complex hunter-gatherers on
the Kodiak Archipelago. In Fitzhugh, B., and Habu, J. (eds.), Beyond Foraging and Collecting,
Evolutionary Change in Hunter-Gatherer Settlement Systems, Kluwer Academic/Plenum Publishers,
New York, pp. 257–304.
Fitzhugh, B. (2003a). The evolution of complex hunter-gatherers on the Kodiak Archipelago. In Habu, J.,
Savelle, J. M., Koyama, S., and Hongo, H. (eds.), Hunter-Gatherers of the North Pacific Rim,
National Museum of Ethnology, Osaka, pp. 13–48.
Fitzhugh, B. (2003b). The Evolution of Complex Hunter-Gatherers, Archaeological Evidence from the
North Pacific, Kluwer Academic/Plenum Publishers, New York.
Flannery, K. V. (1969). Origins and ecological effects of early domestication in Iran and the Near East. In
Ucko, P. J., and Dimbleby, G. (eds.), The Domestication and Exploitation of Plants and Animals,
Aldine, Chicago, pp. 73–100.
Freeman, J., and Anderies, J. M. (2012). Intensification, tipping points, and social change in a coupled
forager-resource system. Human Nature 23: 419–446.
Garfinkel, A. P., Marcom, G., and Schiffman, R. A. (2007). Culture crisis and rock art intensification:
Numic ghost dance painting and Coso representational petroglyphs. American Indian Rock Art 33:
83–103.
González-Insuasti, M. S., and Caballero, J. (2007). Managing plant resources: How intensive can it be?
Human Ecology 35: 303–314.
González-Sainz, C., and González-Urquijo, J. E. (2007). El Magdaleniese reciente en la región
Cantábrica. In Fano, M. A. (ed.), Las sociadades del Paleolı́tico Cantábrico, Kobie Anejo 8, Bilbao,
pp. 275–308.
Gould, R. A. (1985). ‘‘Now let’s invent agriculture…’’: A critical review of concepts of complexity
among hunter-gatherers. In Price, T. D., and Brown, J. (eds.), Prehistoric Hunter-Gatherers,
Academic Press, Orlando, FL, pp. 427–435.
Grayson, D. K., and Cannon, M. D. (1999). Human paleoecology and foraging theory in the Great Basin.
In Beck, C. (ed.), Models for the Millennium, University of Utah Press, Salt Lake City, pp. 141–151.
Grayson, D. K., and Delpech, F. (2002). Specialized early Upper Paleolithic hunters in southwestern
France? Journal of Archaeological Science 29: 1439–1449.
Grayson, D. K., Delpech, F., Rigaud, J.-P., and Simek, J. F. (2001). Explaining the development of dietary
dominance by a single ungulate taxon at Grotte XVI, Dordogne, France. Journal of Archaeological
Science 28: 115–125.

123
Author's personal copy
J Archaeol Res

Gremillion, K. J., Barton, L., and Piperno, D. R. (2014). Particularism and the retreat from theory in the
archaeology of agricultural origins. Proceedings of the National Academy of Sciences 111:
6171–6177.
Grimstead, D. N. (2012). Prestige and prejudice: The role of long distance big game hunting as an optimal
foraging decision. American Antiquity 77: 168–178.
Gutiérrez-Zugasti, I., Cuenca-Solana, D., Rasines del Rı́o, P., Muñoz, E., Santamarı́a, S., and Morlote, J.
M. (2013). The role of shellfish in hunter-gatherer societies during the early Upper Palaeolithic: A
view from El Cuco Rockshelter, Northern Spain. Journal of Anthropological Archaeology 32:
242–256.
Hamilton, M. J., Milne, B. T., Walker, R. S., and Brown, J. H. (2007). Nonlinear scaling of space use in
human hunter-gatherers. Proceedings of the National Academy of Sciences 104: 4765–4769.
Hard, R. J., and Katzenberg, M. A. (2011). Stable isotope study of hunter-gatherer-fisher diet, mobility,
and intensification on the Texas Gulf coastal plain. American Antiquity 76: 709–751.
Hardin, G. (1968). The tragedy of the commons. Science 162: 1243–1248.
Hassan, F. E. (1977). The dynamics of agricultural origins in Palestine: A theoretical model. In Reed, C.
A. (ed.), Origin of Agriculture, Mouton, The Hague, pp. 589–609.
Hassan, F. E. (1981). Demographic Archaeology, Academic Press, New York.
Hawkes, K. (1991). Showing off: Tests of another hypothesis about men’s foraging goals. Ethnology and
Sociobiology 11: 29–54.
Hawkes, K., O’Connell, J. F., and Blurton-Jones, N. G. (1989). Hardworking Hadza grandmothers. In
Standen, V., and R. Foley, R. (eds.), Comparative Socioecology of Mammals and Man, Blackwell,
London, pp. 341–366.
Hawkes, K., O’Connell, J. O., and Jones, N. B. (1991). Hunting income patterns among the Hadza: Big
game, common goods, foraging goals, and the evolution of the human diet. Philosophical
Transactions of the Royal Society 334: 243–250.
Haws, J. A., and Hockett, B. S. (2004). Theoretical perspectives on the dietary role of small mammals in
human evolution. In Brugal, J.-P., and Desse, J. (eds.), Petits animaux et societes humaines: du
complements alimentaires aux resources utilitaires, XXIVe Recontres Internationals d’Archeologie
et d’Histire d’Dantibes, Antibes, France, pp. 533–546.
Hayden, B. (1981). Research and development in the stone age: Technological transitions among hunter-
gatherers. Current Anthropology 22: 519–548.
Hayden, B. (1990). Nimrods, piscators, pluckers, and planters: The emergence of food production.
Journal of Anthropological Archaeology 9: 31–69.
Hayden, B. (1995). Pathways to power: Principles for creating socioeconomic inequalities. In Price, T. D.,
and Feinman, G. (eds.), Foundations of Social Inequality, Plenum, New York, pp. 15–85.
Hayden, B. (2000). The Ancient Past of Keatley Creek, Vol. 2: Socioeconomy, Archaeology Press, Simon
Fraser University, Burnaby, British Colombia.
Hayden, B. (2001). Richman, poorman, beggarman, chief: The dynamics of social inequality. In Feinman,
G. M., and Price, T. D. (eds.), Archaeology at the Millennium, A Sourcebook, Kluwer Academic/
Plenum Publishers, New York, pp. 231–272.
Hayden, B. (2005). The emergence of large villages and large residential corporate group structures
among complex hunter-gatherers at Keatley Creek. American Antiquity 70: 169–174.
Hayden, B. (2007). The Pithouses of Keatley Creek, Harcourt Brace, Fort Worth, TX.
Henry, D. O. (1985). Preagricultural sedentism: The Natufian example. In Price, T. D., and Brown, J. A.
(eds.), Prehistoric Hunter-Gatherers: The Emergence of Cultural Complexity, Academic Press,
Orlando, FL, pp. 364–384.
Hildebrandt, W. R., and Jones, T. L. (2002). Depletion of prehistoric pinniped populations along the
California and Oregon coasts: Were humans the cause? In Kay, C. E., and Simmons, R. T. (eds.),
Wilderness and Political Ecology: Aboriginal Influences and the Original State of Nature,
University of Utah Press, Salt Lake City, pp. 72–100.
Hildebrandt, W. R., and McGuire, K. R. (2002). The ascendance of hunting during the California Middle
Archaic: An evolutionary perspective. American Antiquity 67: 231–236.
Hiscock, P. (2008). Archaeology of Ancient Australia, Routledge, London.
Hockett, B., and Haws, J. A. (2003). Nutritional ecology and diachronic trends in Paleolithic diet and
health. Evolutionary Anthropology: Issues, News, and Reviews 12: 211–216.
Hockett, B., and Haws, J. A. (2005). Nutritional ecology and the human demography of Neanderthal
extinction. Quaternary International 137: 21–34.

123
Author's personal copy
J Archaeol Res

Hockett, B. S. (2009). Continuity in animal resource diversity in the late Pleistocene human diet of central
Portugal. Before Farming 2: 1–14.
Holling, C. S. (2001). Understanding the complexity of economic, ecological, and social systems.
Ecosystems 4: 390–405.
Holly, D. H. (2005). The place of ‘‘others’’ in hunter-gatherer intensification. American Anthropologist
107: 207–220.
Hull, K. L. (2012). A land of many people: Population dynamics as context and catalyst. In Jones, T. L.,
and Perry, J. E. (eds.), Contemporary Issues in California Archaeology, Left Coast Press, Walnut
Creek, CA, pp. 73–92.
Jennings, J. D. (1968). Prehistory of North America, McGraw-Hill, New York.
Jerardino, A. (2010). Large shell middens in Lamberts Bay, South Africa: A case of hunter-gatherer
resource intensification. Journal of Archaeological Science 37: 2291–2302.
Jerardino, A. (2012). Large shell middens and hunter-gatherer resource intensification along the west
coast of South Africa: The Elands Bay case study. The Journal of Island and Coastal Archaeology
7: 76–101.
Jochim, M. (1976). Hunter-Gatherer Subsistence and Settlement: A Predictive Model, Academic Press,
New York.
Jochim, M. A. (1988). Optimal foraging and the division of labor. American Anthropologist 90: 130–136.
Johnson, A. L. (2013). Exploring adaptive variation among hunter-gatherers with Binford’s frames of
reference. Journal of Archaeological Research 22: 1–42.
Johnson, A. L., and Hard, R. J. (2008). Exploring Texas archaeology with a model of intensification.
Plains Anthropologist 53: 137–153.
Jones, E. L. (2006). Prey choice, mass collecting, and the wild European rabbit (Oryctolagus cuniculus).
Journal of Anthropological Archaeology 25: 275–289.
Jones, E. L. (2009). Climate change, patch choice, and intensification at Pont d’Ambon (Dordogne,
France) during the Younger Dryas. Quaternary Research 72: 371–376.
Jones, E. L. (2013). Mobility, settlement, and resource patchiness in Upper Paleolithic Iberia. Quaternary
International 318: 46–52.
Jones, T. L. (2003). Prehistoric Human Ecology of the Big Sur Coast, California, Contributions No. 61,
Archaeological Research Facility, University of California, Berkeley.
Jones, T. L., and Richman, J. R. (1995). On mussels: Mytilus californianus as a prehistoric resource.
North American Archaeologist 16: 33–59.
Jordan, T. A. (2003). Ecological and cultural contributions of controlled fire use by native Californians: A
survey of literature. American Indian Culture and Research Journal 27: 77–90.
Keeley, L. H. (1988). Hunter-gatherer economic complexity and ‘‘population pressure’’: A cross-cultural
approach. Journal of Anthropological Archaeology 7: 373–411.
Kelly, R. L. (2013). The Lifeways of Hunter-Gatherers: The Foraging Spectrum, Cambridge University
Press, Cambridge.
Kelly, R. L., Surovell, T. A., Shuman, B. N., and Smith, G. M. (2013). A continuous climatic impact on
Holocene human population in the Rocky Mountains. Proceedings of the National Academy of
Sciences 110: 443–447.
Kennedy, M. A. (2005). An Investigation of Hunter-Gatherer Shellfish Foraging: Archaeological and
Geochemical Evidence from Bodega Bay, California, Ph.D. Dissertation, Department of Anthro-
pology, University of California, Davis.
Kennett, D. J. (2005). The Island Chumash: Behavioral Ecology of a Maritime Society, University of
California Press, Berkeley.
King, C. D. (1990). Evolution of Chumash Society, A Comparative Study of Artifacts Used for Social
System Maintenance in the Santa Barbara Channel Region Before AD 1804, Garland, New York.
Kirch, P. V. (1994). The Wet and the Dry: Irrigation and Agricultural Intensification in Polynesia,
University of Chicago Press, Chicago.
Kirch, P. V. (2005). Archaeology and global change: The Holocene record. Annual Review of
Environment and Resources 30: 409–440.
Klein, R. G. (2008). Out of Africa and the evolution of modern human behavior. Evolutionary
Anthropology 17: 267–281.
Klein, R. G., Avery, G., Cruz-Uribe, K., Halkett, D., Parkington, J. E., Steele, T., Volman, T. P., and
Yates, R. (2004). The Ysterfontein 1 Middle Stone Age site, South Africa, and early human
exploitation of coastal resources. Proceedings of the National Academy of Sciences 101:
5708–5715.

123
Author's personal copy
J Archaeol Res

Klein, R. G., and Steele, T. E. (2013). Archaeological shellfish size and later human evolution in Africa.
Proceedings of the National Academy of Sciences 110: 10910–10915.
Krebs, J. R., and Davies, N. B. (1997). Behavioral Ecology: An Evolutionary Approach, 4th ed.,
Blackwell, Oxford.
Leach, H. M. (1999). Intensification in the Pacific: A critique of the archaeological criteria and their
application. Current Anthropology 40: 311–340.
Lee, R. B. (1968). What hunters do for a living, or, how to make out on scarce resources. In Lee, R. B.,
and Devore, I. (eds.), Man the Hunter, Aldine de Gruyter, New York, pp. 30–48.
Lepofsky, D., and Peacock, S. L. (2004). A question of intensity, exploring the role of plant foods in
Northern Plateau prehistory. In Prentiss, W. C., and Kuijt, I. (eds.), Complex Hunter-Gatherers,
Evolution and Organization of Prehistoric Communities on the Plateau of Northwestern North
America, University of Utah Press, Salt Lake City, pp. 115–139.
Lewis, H. T. (1973). Patterns of Indian Burning in California: Ecology and Ethnohistory, Ballena Press,
Ramona, CA.
Lilley, I. (2000). So near and yet so far: Reflections on archaeology in Australia and Papau New Guinea,
intensification and culture contact. Australian Archaeology 50: 36–44.
Lourandos, H. (1988). Paleopolitics: Resource intensification in aboriginal Australia and Papua New
Guinea. In Ingold, T., Riches, D., and Woodburn, J. (eds.), Hunters and Gatherers I: History,
Evolution and Social Change, Berg, Oxford, pp. 148–160.
Lourandos, H., and Ross, A. (1994). The great ‘intensification debate’: Its history and place in Australian
archaeology. Australian Archaeology 39: 54–63.
Lupo, K. D. (2007). Evolutionary foraging models in zooarchaeological analysis: Recent applications and
future challenges. Journal of Archaeological Research 15: 143–189.
Lupo, K. D., and Schmitt, D. N. (2005). Small prey hunting technology and zooarchaeological measures
of taxonomic diversity and abundance: Ethnoarchaeological evidence from central African forest
foragers. Journal of Anthropological Archaeology 24: 335–353.
MacArthur, R. H., and Pianka, E. R. (1966). An optimal use of a patchy environment. American
Naturalist 100: 603–609.
Malthus, T. R. (1798). An Essay on the Principle of Population as It Affects the Future Improvements of
Society, J. Johnson, in St. Paul’s Church Yard, London.
Manne, T., and Bicho, N. F. (2009). Vale Boi: Rendering new understandings of resource intensification
and diversification in southwestern Iberia. Before Farming 2009: 1–21.
Manne, T., Cascalhiera, J., Evora, M., Marreiros, J., and Bicho, N. F. (2011). Intensive subsistence
practices at Vale Boi, an Upper Paleolithic site in southwestern Portugal. Quaternary International
264: 83–99.
Mannino, M. A., and Thomas, K. D. (2002). Depletion of a resource? The impact of prehistoric human
foraging on intertidal mollusc communities and its significance for human settlement, mobility and
dispersal. World Archaeology 33: 452–474.
Marlowe, F. W. (2005). Hunter-gatherers and human evolution. Evolutionary Anthropology 14: 54–67.
Marx, K. (2010). Das Kaptital, Pacific Publishing Studio, Madison Park (reprint of Das Kapital, Vol. 1,
1868).
Matson, R. G. (1983). Intensification and the development of cultural complexity: The Northwest versus
the Northeast coast. In Nash, R. (ed.), The Evolution of Maritime Cultures on the Northeast Coasts
of North America, Publication No. 11, Department of Archaeology, Simon Frasier University,
Burnaby, BC, pp. 125–148.
Matson, R. G. (1992). The evolution of Northwest Coast subsistence. In Croes, D. R., Hawkins, R. A., and
Isaac, B. L. (eds.), Long-Term Subsistence Change in Prehistoric North America, JAI Press,
Greenwich, CT, pp. 367–428.
McCarthy, H. (1993). A Political Economy of Western Mono Acorn Production, Ph.D. dissertation,
Department of Anthropology, University of California, Davis.
McGuire, K. R., and Hildebrandt, W. R. (2005). Re-thinking Great Basin foragers, prestige hunting and
costly signaling during the Middle Archaic period. American Antiquity 70: 695–712.
McGuire, K. R., Hildebrandt, W. R., and Carpenter, K. L. (2007). Costly signaling and the ascendance of
no-can-do archaeology: A reply to Codding and Jones. American Antiquity 72: 358–365.
Metcalfe, D., and Barlow, K. R. (1992). A model for exploring the optimal trade-off between field
processing and transport. American Anthropologist 94: 340–356.
Morgan, C. (2009). Climate change, uncertainty and prehistoric hunter-gatherer mobility. Journal of
Anthropological Archaeology 28: 382–396.

123
Author's personal copy
J Archaeol Res

Morgan, C., and Bettinger, R. L. (2012). Chapter 16: Great Basin foraging strategies. In Pauketat, T. (ed.), The
Oxford Handbook of North American Archaeology, Oxford University Press, New York, pp. 185–198.
Morgan, C., Fischer, J., and Pomerleau, M. M. (2012). High-altitude intensification and settlement and in
Utah’s Pahvant Range. Journal of California and Great Basin Anthropology. 32: 27–45.
Morrison, A. E., and Hunt, T. L. (2007). Human impacts on the nearshore environment: An
archaeological case study from Kaua‘i, Hawaiian islands. Pacific Science 61: 325–345.
Morrison, K. D. (1994). The intensification of production: Archaeological approaches. Journal of
Archaeological Method and Theory 1: 111–159.
Morrison, K. D., Feinman, G. M., Nicholas, L. M., Ladefoged, T. N., Myrdal-Runebjer, E., Stone, G. D.,
and Wilk, R. (1996). Typological schemes and agricultural change: Beyond Boserup in precolonial
South India [and comments and reply]. Current Anthropology 37: 583–608.
Moss, M. L. (2012). Understanding variability in Northwest Coast faunal assemblages: Beyond economic
intensification and cultural complexity. The Journal of Island and Coastal Archaeology 7: 1–22.
Mousoudi, A. (2011). Cultural Evolution: How Darwinian Theory Can Explain Human Culture and
Synthesize the Social Sciences, University of Chicago Press, Chicago.
Muñoz, S. E., Gajewski, K., and Peros, M. C. (2010). Synchronous environmental and cultural change in
the prehistory of the northeastern United States. Proceedings of the National Academy of Sciences
107: 22008–22013.
Munro, N. D. (2004). Zooarchaeological measures of hunting pressure and occupation intensity in the
Natufian. Current Anthropology 45(Supplement): S5–S33.
Munro, N. D., and Atici, L. (2009a). Epipaleolithic subsistence intensification in the southern Levant: The
faunal evidence. In Hublin, J.-J., and Richards, M. P. (eds.), The Evolution of Hominin Diets:
Integrating Approaches to the Study of Paleolithic Subsistence, Springer, New York, pp. 141–155.
Munro, N. D., and Atici, L. (2009b). Human subsistence change in the late Pleistocene Mediterranean
Basin: The status of research on faunal intensification, diversification and specialisation. Before
Farming 2009: 5–10.
Nagaoka, L. (2002). The effects of resource depression on foraging efficiency, diet breadth, and patch use
in southern New Zealand. Journal of Anthropological Archaeology 21: 419–442.
Nagaoka, L. (2005). Declining foraging efficiency and moa carcass exploitation in southern New Zealand.
Journal of Archaeological Science 32: 1328–1338.
O’Shea, J. (1981). Coping with scarcity: Exchange and social storage. In Sheridan, A., and Bailey, G.
(eds.), Economic Anthropology: Towards an Integration of Ecological and Social Approaches, BAR
International Series 96, Archaeopress, Oxford, pp. 167–183.
O’Connell, J. F., and Allen, J. (2012). The restaurant at the end of the universe: Modeling the colonisation
of Sahul. Australian Archaeology 74: 5–17.
Pennington, R. (2001). Hunter-gatherer demography. In Panter-Brick, C., Layton, R. H., and Rowley-
Conwy, P. (eds.), Hunter-Gatherers: An Interdisciplinary Perspective, Cambridge University Press,
Cambridge, pp. 170–204.
Peros, M. C., Munoz, S. E., Gajewski, K., and Viau, A. E. (2010). Prehistoric demography of North
America inferred from radiocarbon data. Journal of Archaeological Science 37: 656–664.
Perry, J. E. (2004). Resource intensification and environmental variability: Subsistence patterns in Middle
and Late period deposits at CA-SBA-225, Vandenberg Air Force Base, California. Journal of
California and Great Basin Anthropology 24: 81–102.
Piperno, D. R., Weiss, E., Holst, I., and Nadel, D. (2004). Processing of wild cereal grains in the Upper
Palaeolithic revealed by starch grain analysis. Nature 430: 670–673.
Prendergast, M. E., Yuan, J., and Bar-Yosef, O. (2009). Resource intensification in the late Upper
Paleolithic: A view from southern China. Journal of Archaeological Science 36: 1027–1037.
Prentiss, W. C., and Kuijt, I. (2004). Complex Hunter-Gatherers, Evolution and Organization of
Prehistoric Communities on the Plateau of Northwestern North America, University of Utah Press,
Salt Lake City.
Prentiss, W. C., Lenert, M., Foor, T. A., Goodale, N. B., and Schlegel, T. (2003). Calibrated radiocarbon
dating at Keatley Creek: The chronology of occupation at a complex hunter-gatherer village.
American Antiquity 68: 719–735.
Price, T. D., and Brown, J. A. (1985). Aspects of hunter-gatherer complexity. In Price, T. D., and Brown,
J. A. (eds.), Prehistoric Hunter-Gatherers: The Emergence of Cultural Complexity, Academic Press,
Orlando, FL, pp. 3–20.
Raab, L. M., and Larson, D. O. (1997). Medieval climatic anomaly and punctuated cultural evolution in
coastal southern California. American Antiquity 62: 319–336.

123
Author's personal copy
J Archaeol Res

Raber, P. (2010). The intensification of rockshelter use in the Late Woodland/late prehistoric period.
North American Archaeologist 31: 367–403.
Redding, W. W. (1988). A general model of subsistence change: From hunting and gathering to food
production. Journal of Anthropological Archaeology 7: 56–97.
Reitz, E. J. (2004). ‘‘Fishing down the food web’’: A case study from St. Augustine, Florida, USA.
American Antiquity 69: 63–83.
Reitz, E. J., and Wing, E. S. (1999). Zooarchaeology, Cambridge University Press, New York.
Rick, T. C. (2007). The Archaeology and Historical Ecology of Late Holocene San Miguel island,
California, Cotsen Institute of Archaeology, University of California, Los Angeles.
Rick, T. C. (2011). Weathering the storm: Coastal subsistence and ecological resilience on late Holocene
Santa Rosa Island, California. Quaternary International 239: 135–146.
Ricklis, R. A., and Weinstein, R. A. (2005). Sea-level rise and fluctuation on the central Texas coast:
Exploring cultural and ecological correlates. In White, N. M. (ed.), Gulf Coast Archaeology: The
Southeastern United States and Mexico, University Press of Florida, Gainesville, pp. 108–154.
Rocek, T. R., and Bar-Yosef, O. (1998). Seasonality and Sedentism: Archaeological Perspectives from
Old and New World Sites, Bulletin 6, Peabody Museum of Archaeology and Ethnology, Harvard
University, Cambridge, MA.
Sassaman, K. E. (2004). Complex hunter-gatherers in evolution and history: A North American
perspective. Journal of Archaeological Research 12: 227–280.
Schiffer, M. B. (1987). Formation Processes of the Archaeological Record, University of New Mexico
Press, Albuquerque.
Smith, A. (2009). The Wealth of Nations, Thrift Books, Blacksburg Virginia (Reprint of An Inquiry into
the Nature and Causes of the Wealth of Nations, 1776).
Smith, B. D. (1998). The Emergence of Agriculture, Scientific American Library, New York.
Smith, B. D. (2001). Low-level food production. Journal of Archaeological Research 9: 1-43.
Smith, E. A., and Wishnie, M. (2000). Conservation and subsistence in small-scale societies. Annual
Review of Anthropology 29: 493–524.
Smith, M. (2013). The Archaeology of Australia’s Deserts, Cambridge University Press, Cambridge.
Speth, J. D. (2004). Hunting pressure, subsistence intensification, and demographic change in the
Levantine Late Middle Paleolithic. In Goren-Inbar, N., and Speth, J. D. (eds.), Human Paleoecology
in the Levantine Corridor, Oxbow Books, Oxford, pp. 149–166.
Starkovich, B. M. (2009). Dietary changes during the Upper Palaeolithic at Klissoura Cave 1 (Prosymni),
Peloponnese, Greece. Before Farming 3: 1–14.
Steele, T. E. (2012). Late Pleistocene human subsistence in northern Africa: The state of our knowledge
and placement in a continental context. In Hublin, J.-J., and McPherron, S. P. (eds.), Modern
Origins: A North African Perspective, Springer, New York, pp. 107–125.
Steele, T., and Klein, R. G. (2005). Mollusk and tortoise size as proxies for stone age population density
in South Africa: Implications for the evolution of human cultural capacity. Munibe 57: 221–237.
Steele, T. E., and Klein, R. G. (2009). Late Pleistocene subsistence strategies and resource intensification
in Africa. In Hublin, J.-J., and Richards, M. P. (eds.), The Evolution of Hominin Diets: Integrating
Approaches to the Study of Palaeolithic Subsistence, Springer, Dordrecht, pp. 113–126.
Stephens, D. W. (1990). Risk and incomplete information in behavioral ecology. In Cashdan, E. (ed.),
Risk and Uncertainty in Tribal and Peasant Economies, Westview Press, Boulder, CO, pp. 19–46.
Stephens, D. W., and Krebs, J. R. (1986). Foraging Theory, Princeton University Press, Princeton, NJ.
Stiner, M. C. (2001). Thirty years on the ‘‘broad spectrum revolution’’ and Paleolithic demography.
Proceedings of the National Academy of Sciences 98: 6993–6996.
Stiner, M. C., and Munro, N. D. (2002). Approaches to prehistoric diet breadth, demography, and prey
ranking systems in time and space. Journal of Archaeological Method and Theory 9: 181–214.
Stiner, M. C., Munroe, N. D., and Surovell, T. A. (2000). The tortoise and the hare: Small-game use, the
broad spectrum revolution, and Paleolithic demography. Current Anthropology 41: 39–73.
Stutz, A. J., Munro, N. D., and Bar-Oz, G. (2009). Increasing the resolution of the broad spectrum
revolution in the southern Levantine Epipaleolithic (19–12 ka). Journal of Human Evolution 56:
294–306.
Terrell, J. E., Hart, J. P., Barut, S., Cellinese, N., Curet, A., Denham, T., Kusimba, C. M., Latinis, K.,
Oka, R., Palka, J., Pohl, M. E., Pope, K. O., Williams, P. R., Haines, H., and Staller, J. E. (2003).
Domesticated landscapes: The subsistence ecology of plant and animal domestication. Journal of
Archaeological Method and Theory 10: 323–368.

123
Author's personal copy
J Archaeol Res

Thakar, H. B. (2011). Intensification of shellfish exploitation: Evidence of species-specific deviation from


traditional expectations. Journal of Archaeological Science 38: 2596–2605.
Thoms, A. V. (1989). The Northern Roots of Hunter-Gatherer Intensification: Camas and the Pacific
Northwest, Ph.D. dissertation, Department of Anthropology, Washington State University, Pullman.
Thoms, A. V. (2003). Cook-stone technology in North America: Evolutionary changes in domestic fire
structures during the Holocene. In Frere-Sautot, M. (ed.), Le feu domestique et ses structures au
Neolithique et aux Ages des Metaux, Collections Prehistoires No. 9, Editions Monique Mergoil,
Saint-Appollinaire, France, pp. 87–96.
Thoms, A. V. (2008). Ancient savannah roots of the carbohydrate revolution in south-central North
America. Plains Anthropologist 53: 121–136.
Thoms, A. V. (2009). Rocks of ages: Propagation of hot-rock cookery in western North America. Journal
of Archaeological Science 36: 573–591.
Thurston, T. L., and Fisher, C. T. (2007a). Intensification, innovation, and change: New perspectives and
future directions. In Thurston, T. L., and Fisher, C. T. (eds.), Seeking a Richer Harvest: The
Archaeology of Subsistence Intensification, Innovation, and Change, Springer, New York,
pp. 249–259.
Thurston, T. L., and Fisher, C. T. (2007b). Seeking a richer harvest: An introduction to the archaeology of
subsistence intensification, innovation, and change. In Thurston, T. L., and Fisher, C. T. (eds.),
Seeking a Richer Harvest: The Archaeology of Subsistence Intensification, Innovation, and Change,
Springer, New York, pp. 1–22.
Todt, D. L. (2007). Upriver and downriver: A gradient of tobacco intensification along the Klamath River,
California and Oregon. Journal of California and Great Basin Anthropology 27: 1–14.
Torrance, R. (1989). Time, Energy, and Stone tools, Cambridge University Press, New York.
Tushingham, S. (2009). The Development of Intensive Foraging Systems in Northwestern California,
Ph.D. dissertation, Department of Anthropology, University of California, Davis.
Tushingham, S., and Bettinger, R. L. (2013). Why foragers choose acorns before salmon: Storage,
mobility, and risk in aboriginal California. Journal of Anthropological Archaeology 32: 527–537.
Ugan, A. (2005). Does size matter? Body size, mass collecting, and their implications for understanding
prehistoric foraging behavior. American Antiquity 70: 75–90.
Ugan, A., and Bright, J. (2001). Measuring foraging efficiency with archaeological faunas: The
relationship between relative abundance indices and foraging returns. Journal of Archaeological
Science 28: 1309–1321.
Ugan, A., Bright, J., and Rogers, A. (2003). When is technology worth the trouble? Journal of
Archaeological Science 30: 1315–1329.
Ulm, S. (2002). Reassessing marine fishery intensification in southeast Queensland. Queensland
Archaeological Research 13: 79–96.
Walters, I. (2001). The aboriginal Moreton Bay fishery: Two things I still know: A response to Ross and
Coghill. Australian Aboriginal Studies 2: 61–62.
Whitaker, A. (2009). Are deer really susceptible to resource depression? Modeling deer (Odocoileus
hemionus) populations under human predation. California Archaeology 1: 93–108.
Whitaker, A. R. (2008). Incipient aquaculture in prehistoric California? Long-term productivity and
sustainability vs. immediate returns for the harvest of marine invertebrates. Journal of Archaeo-
logical Science 35: 1114–1123.
Whitaker, A. R., and Byrd, B. F. (2012). Boat-based foraging and discontinuous prehistoric red abalone
exploitation along the California coast. Journal of Anthropological Archaeology 31: 196–214.
Whitaker, A. R., and Byrd, B. F. (2014). Social circumscription, territoriality, and the late Holocene
intensification of small-bodied shellfish along the California coast. Journal of Island and Coastal
Archaeology 9: 150–168.
Whitaker, A. R., and Carpenter, K. (2012). Economic foraging at a distance is not a question of if but
when: A response to Grimstead. American Antiquity 77: 160–167.
White, A. A. (2013). Subsistence economics, family size, and the emergence of social complexity in
hunter-gatherer systems in eastern North America. Journal of Anthropological Archaeology 32:
122–163.
Williams, A. N. (2012). The use of summed radiocarbon probability distributions in archaeology: A
review of methods. Journal of Archaeological Science 39: 578–589.
Williams, A. N. (2013). A new population curve for prehistoric Australia. Proceedings of the Royal
Society, B 280: 1–9.

123
Author's personal copy
J Archaeol Res

Winterhalder, B., and Goland, C. (1997). An evolutionary ecology perspective on diet choice, risk, and
plant domestication. In Gremillion, K. J. (ed.), People, Plants, and Landscapes: Studies in
Paleoethnobotany, University of Alabama Press, Tuscaloosa, pp. 123–160.
Winterhalder, B., and Leslie, P. (2002). Risk-sensitive fertility, the variance compensation hypothesis.
Evolution and Human Behavior 23: 59–82.
Winterhalder, B., Lu, F., and Tucker, B. (1999). Risk-sensitive adaptive tactics: Models and evidence
from subsistence studies in biology and anthropology. Journal of Archaeological Research 7:
301–348.
Wittfogel, K. (1957). Oriental Despotism, Yale University Press, New Haven, CT.
Wohlgemuth, E. (1996). Resource intensification in prehistoric central California: Evidence from
archaeobotanical data. Journal of California and Great Basin Anthropology 18: 81–103.
Yu, P.-L. (2006). Pit Cooking and Intensification of Subsistence in the American Southwest and Pacific
Northwest, Ph.D. dissertation, Department of Anthropology, Southern Methodist University, Dallas,
TX.
Zangrando, A. F. J. (2009). Is fishing intensification a direct route to hunter-gatherer complexity? A case
study from the Beagle Channel region (Tierra del Fuego, southern South America). World
Archaeology 41: 589–608.
Zeder, M. A. (2012). The broad spectrum revolution at 40: Resource diversity, intensification, and an
alternative to optimal foraging explanations. Journal of Anthropological Archaeology 31: 241–264.
Zvelebil, M. (1986). Mesolithic societies and the transition to farming: Problems of time, scale and
organization. In Zvelebil, M. (ed.), Hunters in Transition: Mesolithic Societies of Temperate
Eurasia and the Transitions to Farming, Cambridge University Press, Cambridge, pp. 167–188.

Bibliography of recent literature

Bettinger, R. L. (2009). Hunter-Gatherer Foraging: Five Simple Models, Elliot Werner Publications,
New York.
Boyd, R., and Richerson, P. J. (2005). The Origin and Evolution of Cultures, Oxford University Press,
Oxford.
Brian, D. (2006). Harry Lourandos, the ‘great intensification debate’, and the representation of indigenous
pasts. In David, B., Barker, B., and McNiven, I. J. (eds.), Social Archaeology of Australian
Indigenous Societies, Aboriginal Studies Press, Canberra, pp. 107–122.
Broughton, J. M., and Cannon, M. D. (2010). Evolutionary Ecology and Archaeology, Applications to
Problems in Human Evolution and Prehistory, University of Utah Press, Salt Lake City.
Codding, B. F., and Jones, T. L. (2013). Environmental productivity predicts migration, demographic, and
linguistic patterns in prehistoric California. Proceedings of the National Academy of Sciences 110:
14569–14573.
Codding, B. F., Whitaker, A. R., and Bird, D. W. (2014). Global patterns in the exploitation of shellfish.
The Journal of Island and Coastal Archaeology 9: 145–149.
Davidson, D. J., Andrews, J., and Pauly, D. (2014). The effort factor: Evaluating the increasing marginal
impact of resource extraction over time. Global Environmental Change 25: 63–68.
Feinman, G. M. (2013). The emergence of social complexity, why more than population size matters. In
Carballo, D. M. (ed.), Cooperation and Collective Action, Archaeological Perspectives, University
Press of Colorado, Boulder, pp. 35–56.
Fitzhugh, B., and Habu, J. (2002). Beyond Foraging and Collecting, Evolutionary Change in Hunter-
Gatherer Settlement Systems, Kluwer Academic/Plenum Publishers, New York.
Grove, M. (2009). Hunter-gatherer movement patterns: Causes and constraints. Journal of Anthropo-
logical Archaeology 28: 222–233.
Hayden, B. (2001). Fabulous feasts: A prolegomenon to the importance of feasting. In Dietler, M., and
Hayden, B. (eds.), Feasts: Archaeological and Ethnographic Perspectives on Food, Politics, and
Power, Smithsonian Institution Press, Washington, DC, pp. 23–64.
Janetski, J. C. (1997). Fremont hunting and resource intensification in the eastern Great Basin. Journal of
Archaeological Science 24: 1075–1088.
Kennett, D. J., and Winterhalder, B. (2006). Behavioral Ecology and the Transition to Agriculture,
University of California Press, Berkeley.

123
Author's personal copy
J Archaeol Res

Klaus, H. D., Spencer Larsen, C., and Tam, M. E. (2009). Economic intensification and degenerative joint
disease: Life and labor on the postcontact north coast of Peru. American Journal of Physical
Anthropology 139: 204–221.
Lupo, K. D., Fancher, J. M., and Schmitt, D. N. (2013). The taphonomy of resource intensification:
Zooarchaeological implications of resource scarcity among Bofi and Aka forest foragers. Journal of
Archaeological Method and Theory 20: 420–447.
Magee, P. (2007). Beyond the desert and the sown: Settlement intensification in late prehistoric
southeastern Arabia. Bulletin of the American Schools of Oriental Research 347: 83-105.
Marcus, J., and Stanish, C. (eds.) (2006). Agricultural Strategies, Cotsen Institute of Archaeology,
University of California, Los Angeles.
Minnegal, M., and Dwyer, P. D. (2001). Intensification, complexity, and evolution: Insights from the
Strickland-Bosavi region. Asia Pacific Viewpoint 42: 269–285.
Morales, M., Barberena, R., Belardi, J. B., Borrero, L., Cortegoso, V., Durán, V., Guerci, A., Goñi, R.,
Gil, A., Neme, G., Yacobaccio, H., and Zárate, M. (2009). Reviewing human–environment
interactions in arid regions of southern South America during the past 3000 years. Palaeogeog-
raphy, Palaeoclimatology, Palaeoecology 281: 283–295.
Morgan, C. (2012). Modeling modes of hunter-gatherer food storage. American Antiquity 77: 714–736.
Neme, G., Gil, A., Otaola, C., and Giardina, M. (2013). Resource exploitation and human mobility:
Trends in the archaeofaunal and isotopic record from central western Argentina. International
Journal of Osteoarchaeology. doi:10.1002/oa.2359/abstract.
Neme, G., Wolverton, S., and Gil, A. (2012). Modelo ecológico para evaluar la sobreexplotación de
ungulados: Implicancias en los conjuntos zooarqueológicos de guanaco del sur de Mendoza.
Archaeofauna 21: 207–218.
Panter-Brick, C., Layton, R. H., and Rowley-Conwy, P. (2001). Hunter-Gatherers: An Interdisciplinary
Perspective, Cambridge University Press, Cambridge.
Phillips, D. A. (2009). Adoption and intensification of agriculture in the North American Southwest:
Notes toward a quantitative approach. American Antiquity 74: 691–707.
Read, D. W., and Leblanc, S. A. (2003). Population growth, carrying capacity, and conflict. Current
Anthropology 44: 59–85.
Richerson, P. J., and Boyd, R. (1998). Homage to Malthus, Ricardo, and Boserup: Toward a general
theory of population, economic growth, environmental deterioration, wealth, and poverty. Human
Ecology Review 4: 85–90.
Richerson, P. J., Boyd, R., and Bettinger, R. L. (2009). Cultural innovations and demographic change.
Human Biology 81: 211–235.
Schmidtz, D. (2004). Satisficing as a humanly rational strategy. In Byron, M. (ed.), Satisficing and
Maximizing: Moral Theorists on Practical Reason, Cambridge University Press, Cambridge,
pp. 30–58.
Smith, E. A. (2013). Agency and adaptation: New directions in evolutionary anthropology. Annual
Review of Anthropology 42: 103–120.
Surovell, T. A. (2009). Toward a Behavioral Ecology of Lithic Technology: Cases from Paleoindian
Archaeology, University of Arizona Press, Tucson.
Thomas, D. H. (2008). Native American Landscapes of St. Catherines Island, Georgia, Papers, No. 88,
American Museum of Natural History, New York.
Winterhalder, B., Baillargeon, W., Cappelletto, F., Daniel, J., and Prescott, I. R. (1988). The population
ecology of hunter-gatherers and their prey. Journal of Anthropological Archaeology 7: 289–328.
Winterhalder, B., Kennett, D. J., Grote, M. N., and Bartruff, J. (2010). Ideal free settlement of California’s
northern Channel Islands. Journal of Anthropological Archaeology 29: 469–490.
Yacobaccio, H. D. (2003). Procesos de intensificacion de camelidos en Los Andes centro-sur. In
Memorias del Tercer Congreso Mundial sobre Camelidos, Facultad de Agronomia, Universidad
Técnica de Oruro. Potosı́, Bolivia, pp. 211–216.
Yacobaccio, H. D. (2006). Economic intensification and social complexity of South Andean hunters and
gatherers. Boletı́n de Arqueologı́a PUCP 10: 305–320.
Ziman, J. (2003). Technological Innovation as an Evolutionary Process, Cambridge University Press,
Cambridge.

123

You might also like