THE EVOLUTION OF HUMAN NUTRITION

BARRY BOGIN

IN : The Anthropology of Medicine

During a lifetime, a human being will eat thousands of pounds of food. The

body will use this food to grow, to repair damaged tissue, and to maintain

organs such as the brain and heart. Some of these foods will be enjoyable

to eat because they are perceived to look appetizing and taste delicious.

Other foods may not be enjoyable to eat, but will be consumed anyway

because they are "good for the body or the spirit." Biochemically, the body

does not distinguish between foods that are liked or disliked, for the human

body does not use food, rather the body requires the biological nutrients

contained in food. Biology, however, is not the entire story of human nutri-

tion. Cultural variables, such as the type of food eaten, its manner of prep-

aration, and the social context in which it is consumed, often determine the

efficacy of that food in meeting human needs for health and well-being. It

is the purpose of the chapter to explore the evolution of some of the bio-

logical and cultural requirements of human nutrition. Although at times the

biology and culture of nutrition will be treated separately, the major theme

of this chapter is to view human nutrition holistically as a biocultural phe-

nomenon.

The Conception of the People of Corn

It was night, and the gods sat thinking in the darkness. Among them were the Bearer,

Begetter, the Makers, Modelers named Tepeu Gucumatz, the Sovereign Plumed

Serpent. Twice before they had tried to create a human being to be servant to the

gods. One time the humans were made of clay and the other time of wood; but on

both occasions the creatures so formed were stupid, without any intellect and without

spirit. So, they were destroyed. As the dawn approached the gods thought, "Morning

has come for humankind, for the people of the face of the earth." Their great wisdom

was revealed in the clear light; they discovered what was needed for human flesh-

white corn and yellow corn. Four animals brought the food: fox, coyote, parrot. and

crow. The animals showed the way to the citadel named Broken Place, Bitter Water

Place. Here was a paradise filled with white and yellow corn and all the varieties of

fruits and vegetables. including pataxte and cacao. The white and yellow corn were

given to Xmucane. the divine Grandmother of the gods, and she ground the corn

nine times. She washed the ground corn from her hands with water and this mixture

made grease. The corn was used to make human flesh, the water made human blood,

and the grease made human fat. From these staple foods were born the strength

and vigor of the new beings. (From the Popol Vuh, the Maya book of the dawn of

life and the glories of gods and kings [compiled from the translations of Tedlock

1985 and Figueros. 1986].)

 

The domestication of maize, or corn, and other plants occurred in Mesoamerica about 7000 B.P. By 3000 B.P.

maize-based agricultural societies were established and these developed into the state-level, hierarchical so-

cieties of the Olmec and, eventually, the Maya. The central place of maize

as the staple food in Maya society is emphasized in the creation story. People

are corn, in both the literary and literal sense. Today, the living Maya people

of Guatemala depend on maize for 80 percent of their energy intake. It is

likely that the ancient Maya also consumed a large portion of their calories

from maize, or more correctly from maize-based foods. Very little maize is

eaten in Guatemala today. Instead people eat tortillas, tamalitos, tamales, ta-

cos, enchiladas, atoles (a beverage), and many other foods and drinks made

from masa harina. Masa harina is a flour made from maize that has been

dried, ground, and processed by boiling in lime water (Figure 6.1). Some of

the "tortilla chips" sold in American supermarkets may be made from a flour

like mass harinha, but most brands are made from corn meal, which is ground

maize without any processing. The difference is vitally important in terms

of nutrition and health, for without the processing, a maize-based diet leads

to death from pellagra. Later in this chapter the biochemical and nutritional

properties of masa harina and the cause of pellagra are explained in greater

detail.

The Maya, ancient and modern, do not live by tortillas alone. At Broken

Place, Bitter Water Place (a supernatural site located inside a mountain), all

varieties of fruits and vegetables were found and given to people. A visit to

any Maya marketplace today in Guatemala or southern Mexico shows that

dozens of species of fruits, vegetables, and dried mushrooms are sold, along

with fresh and dried fish and meat. Archaeological and ethnographic field-

work substantiates the diversity of foods in the Maya diet over the past 1,000

years or more (Saenz de Tejada 1988). Even pataxte and cacao were given

to people by the gods (Figure 6.2). These are fruits from which cocoa and

 

 

chocolate are made cochocoholics might recite an extra prayer of thanks to

Sovereign Plumed Serpent before retiring tonighr). A chocolate and hot pep-

per beverage was a drink used in Maya religious ritual. and was usually

reserved for the royal family or other people of high status. Thus, food is

used not only to sustain the body, but also to demarcate social position and

as part of religious behavior.

NUTRIENTS VERSUS FOOD

Nutritional biochemists have determined that there are 50 essential nu-

trients required for growth, maintenance, and the repair of the body. Essen-

tial nutrients are those substances that the body needs but cannot

manufacture. These substances are divided into six classes: protein, carbo-

hydrate, fat, vitamins, minerals, and water. Table 6.1 lists the essential nu-

trients in these categories. One way that nutrients are shown to be essential

is via experiments with non-human animals. A young rat, pig, or monkey is

fed a diet that includes all of the known nutrients except the one being

tested. If the animal gets sick, stops growing, loses weight, or dies it usually

means that the missing nutrient is essential for that animal. Such experi-

ments do not prove that the same nutrient is needed for people. Some

controlled experiments were done in the twentieth century with people,

such as prisoners and residents of villages in underdeveloped nations. Since

about 1980 these experiments have been considered unethical. Certain med-

ical conditions deprive people of nutrients, and social, economic, and po-

litical conditions of life also deprive people of food and nutrients. By using

these " experiments of nature," along with past research, it is possible to

prove the necessity of the essential nutrients.

People do not usually intake essential nutrients directly as pure chemicals,

rather we eat food. This was certainly true for all of our animal ancestors

throughout evolutionary history. Human foods come from five of the six

Kingdoms of living organisms: plants, animals, fungi (e.g., mushrooms), pro-

tists (e.g., species of algae referred to as "seaweed"), and eubacteria (e.g.,

bacteria used in fermented foods). These organisms present us with a daz-

zling array of colors, flavors. odors, textures, shapes, and sizes. The sixth

Kingdom, archaebacteria. are not eaten directly, but are essential in the diet

of other species that people do eat. Herbivores, for example, have archae-

bacteria in their guts to digest the plant cellulose.

 

 

Table 6.1 Essential Nutrients

Carboyhdrate

Glucose

Fat or Lipid

Linoleic acid

Linolenic acid

Protein

Amino acids

Leucine

Isoleucine

Lysine

Methionine

Phenylalanine

Threonine

Tryptophan

VaIine

Histidine

Nonessential amino nitrogen

Minerals

Iron

Selenium

Zinc

Calcium

Phosphorus

Sodium

Potassium

Sulfur

Chlorine

Magnesium

Manganese

Copper

Cobalt

Molybdenum

Iodine

Chromium

Vanadium

tin

Nickel

Silicon

Boron

Arsenic

Fluorine

Vitamins

Fat-soluble

A (retinal)

B (cholecalciferol)

E (tocopherol)

K

Water-soluble

Thiamin

Riboflavin

Niacin

Biotin

Folic acid

Vitamin B6, (pyridoxine)

Vitamin B12, (cobalamin)

Pantothenic acid

Vitamin C (ascorbic acid)

Water

Source: Guthrie and Picciano 1995.

 

EATING A BALANCED DIET

How does a person know which foods to eat so that all of the essential

nutrients are consumed in required amounts. ) Children learn what to eat

because they are dependent on their parents, or other older individuals, to

prepare their food. By tasting these foods, and watching older people prepare

them, children acquire patterns of food preferences, including what should

not be eaten, under what social conditions a food should be eaten, and the

ways to prepare foods. Thus people learn what they like, for not all people

eat all the same foods. For instance, some people in the United States eat

chocolate covered ants, but most Americans do not think of insects as food.

 

 

In parts of Africa and South America, however, insects such as ants, termites,

and beetle larva are food, in fact they are considered delicacies. Yanomamo

Indians of southern Venezuela cultivate certain plants in which they know

beetles will lay their eggs. The Yanomamo harvest the beetle larvae and eat

them raw or roasted (Chagnon 1983). From a nutritional point of view insects

are excellent sources of protein, fats, and some minerals. In fact, pound for

pound, grasshoppers have more protein than cattle or hogs, yet this fact is

unlikely to encourage the sale of "grasshopper nuggets" at fast-food outlets

in the United States.

Every group of people has developed a cuisine: that is, an assortment of

foods and a style of cooking that is unique to that culture. Some examples

are Italian cooking, Chinese cooking, and Mexican cooking. Even Americans

have a cuisine, including foods such as corn-on-the-cob and hamburgers.

Despite the differences in specific foods, the cuisine of each human culture

provides all of the essential nutrients. No one knows how cuisines developed

to meet human biochemical requirements. Experiments with non-human

animals and with people indicate that diets, or cuisines, are developed by

learning to avoid foods that produce illness or feelings of malaise and seeking

foods that promote feelings of well-being (Franken 1988:107).

One fascinating aspect of food preferences in different cultures is the way

two or more foods are combined and eaten together to help assure nutritional

adequacy. One example is complimentary protein consumption. Table 6.1

shows that there are nine amino acids (the building blocks of proteins) that

are essential nutrients. There are 11 additional amino acids in nature that

are needed for life but are not essential nutrients. Not all foods contain all

nine essential amino acids, so we must eat several foods to get them all-

the amino acids in some foods complementing those lacking in others. Ce-

real grains, such as wheat and rice, lack some of the amino acids that are

found in beans, peas, milk, and cheeses. Conversely, beans, peas, milk, and

cheeses lack the amino acids found in cereal grains. In the Middle East,

many people eat wheat and cheese in the same dish. In Mexico, beans,

tortillas, and rice are popular, while on the island of Jamaica peas and rice

is the national favorite. In the United States, cereal (grains) and milk are

complementary protein sources popular at the breakfast meal. The biochem-

istry of complementary protein foods has been discovered only recently, yet

the cultural history of this food practice is ancient.

Each culture developed its own cuisine for many reasons. Not all foods

grow in all countries, for instance maize originally comes from Central Amer-

ica and rice originally comes from Asia. But most food preferences cannot

be so easily explained. The isolation of many human cultures, exploration

and contact between cultures, ethnic identity, and social, economic, political,

and religious status are some further explanations. Hindu culture, for ex-

ample, specifies different cuisines for people of different castes. According

to Burghardt's (1990) analysis of Hindu dietary recommendation, not all

castes can tolerate all foods. The intolerance is due to harmful reactions

between the qualities of the food (such as animal meat) and the nature of

the bodies of different caste members. Thus Hindu epistemology does not

identify the universal set of essential nutrients recognized by Western bio-

medical research. Many other unknown factors occurring throughout

thousands of years of human history are also responsible for the development

of culture-specific cuisines.

From the foregoing, two universal observations about human nutrition can

be made: (1) All people have the same basic biological requirements for

nutrients; and (2) Each culture has a unique cuisine that has the potential

to satisfy these nutrient requirements.

In addition some universal features of human food systems have been

complied by Pelto and Pelto (1983):

1. People are extremely omnivorous. eating hundreds of different species of plants.

animals, fungi, bacteria, and even algae.

2. People depend on systems of food transport from the place where foods are found

or acquired to their place of consumption.

3. People make use of systems for food storage that protect the nutritional quality

of foods from the time of their acquisition until the time of their consumption.

That time period may last for months, even in premodern societies.

4. People expend great effort on food preparation. such as cooking, mixing, flavoring,

and detoxifying natural ingredients, and depend on technology to do this prepa-

ration (e.g., the hand-axes and fire used by Homo erectus or the food processors

and microwave ovens of Homo sapiens).

5. People share and exchange food regularly and have cultural rules that order such

sharing and exchanges.

6. People have food taboos; that is, social proscriptions against the consumption of

certain foods based on age, sex, state of health. religious beliefs, and other cul-

turally defined reasons.

One final item must be included in this list of human food behavior.

7. People use foods for non-nutritional purposes, such as for medicine to cure or

cause disease and as offerings in ritual or religious behavior (see the chapter by

Etkin and Ross in this volume). In these contexts food may have some physiologic

function (plants do contain active pharmaceutical compounds), but the foods also

have symbolic meaning for the people using them.

Evidence from fossil and archaeological remains of human ancestors in-

dicates that these universal features of human nutrition and food have been

in existence for at least 35,000 years, and possibly more than 100,000 years.

Yet, until this century, most foods were acquired locally. The most parsi-

monious way to account for these biological and cultural universals relating

to food is to hypothesize that a common evolutionary history for all people

shaped human nutritional requirements, food acquisition and processing sys-

tems, and food behavior. This is a hypothesis that can be verified or rejected

by research.

SOURCES OF KNOWLEDGE

There are several kinds of data that may be considered in the study of

the evolution of human nutrition. Archaeological and paleontological evi-

dence provide the only direct data on what our ancestors ate and what effect

diet may have had on our physical and behavioral evolution. However, stud-

ies of living primates and other mammals, living hunter/gatherer societies,

and crosscultural comparisons of cuisines provide indirect evidence that is

useful in reconstructing human nutritional history.

PRIMATE STUDIES

The living primates include prosimians. New World monkeys, Old World

monkeys, Asian and African apes, and people. Fossil evidence indicates that

all primates evolved from insectivorelike mammals that lived some 7.5 mil-

lion years ago. The geological context of these fossils indicates that the

general habitat was tropical forest. Primate ancestors may have been those

insectivores that moved into the flowering trees of these tropical forests to

exploit insects, and then the flowers, fruits, gums, and nectars of those trees

(Cartmill 1974; Conroy 1990). The flowering plants and trees, called angio-

sperms, appear in the fossil record about 100 million years ago, and their

appearance opened up new habitats and ecological niches that promoted the

coevolution of other species, including the primates. The earliest primates

of the Paleocence period (65-55 million years ago) exploited an insect-eating

niche. Most species had jaws that moved in a scissorlike motion and teeth

with pointed cusps, both features well suited for catching insects, by rapid

mouth "snapping" and piercing their exoskeletons (the "crunchy" covering

of the body) to extract internal tissues and fluids. By the late Paleocence

many species show some dental traits indicating a mixed diet of insects,

fruits, leaves, seeds, or gums. By about 55 million years ago primate fossils

show changes in jaws and teeth toward those of living forms, with jaws

adapted for greater power in biting and chewing. It seems that by that time

most primates were eating fruits, leaves, and seeds as well as insects.

Thus, the general primate dietary pattern is ancient. That pattern is based

on the ability to eat a wide variety of foods in order to meet nutritional

requirements. Primate nutritional requirements are highly varied; the higher

primates, including people, may be the animals with the longest list of es-

sential nutrients. The reason for this may be our tropical origins. Today,

tropical forests are characterized by having a high diversity of species, but a

low density of any given species. There are thousands of species of tropical

trees and at any one site there may be between 50 to 100 different species

per hectare (Oates 1987), but only a few trees of the same species may be

growing on that hectare. In contrast, temperate and high-altitude forests are

often characterized by a few tree species, such as pine-oak forests, but large

numbers of trees of those species. The diversity and density of animal spe-

cies in tropical forests follows the pattern for plant life. There is no reason

to expect that ancient tropical forests were different than modern tropical

forests in terms of species diversity. Ancestral primates capable of eating a

wide variety of foods would have had a veritable smorgasbord of choices,

and judging by their descendants, the living primates, many food types were

consumed.

The large number of essential nutrients required in the human diet, then,

is likely a consequence of the tropical primate diet. With a wide variety of

food resources, especially fruit, foliage, and insects, ancestral primates were

able to obtain many vitamins, minerals, protein, carbohydrates, and fats from

their diet. It is metabolically expensive, in terms of energy consumption, for

an organism to manufacture its own nutrients (a process called autotroph-

ism). Thus, through mutation and selection, those early primates that re-

duced autotrophism and shifted to a dependency on dietary intake to meet

nutrient needs would have gained an energetic advantage, one that could

be put to use, for instance, toward increasing reproduction. All mammals,

for example, require vitamin C for maintenance and repair of body tissue,

but only in some mammals, including all members of the primate order, is

vitamin C (ascorbic acid) an essential nutrient. About 25 million years ago a

mutation occurred in the metabolic pathway that produces vitamin C in

primates ancestral to living monkeys, apes, and people. The glucose (car-

bohydrate energy) needed to convert biochemical precursors to ascorbic acid

was released for use by other body systems (Scrimshaw and Young 1976).

The wide distribution of vitamin C sources in tropical environments and the

ability of primates to utilize these sources assured that this nutrient could

be supplied by the diet alone.

Using published data, Harding (1981) divided naturally occurring tropical

forest foods into eight categories and calculated the dietary frequency of

each category for 131 species of primates from all families, excluding people

(Table 6.2). The dietary frequency is defined as the percentage of those

species surveyed "for which a given food category was listed in the diet"

(p. 206). The data show that variety is the rule, and most species included

seven of the eight food categories in their diets ("grasses and roots" was the

category most often missing). The chimpanzee, our closest living primate

relative, eats foods from all eight categories. It is worth noting at this point,

that on a worldwide basis, living people eat more grasses, such as wheat and

maize, and roots, such as potatoes and manioc, than any other foods listed

in Table 6.2.

Table 6.2

Dietary Frequency and Major Diet Components of 131 Primate Species

Food category Dietary frequency Major component'
Fruit 90 45
Soil plant foods 79 9
Mature leaves 69 15
Invertebrates 65 23
Seeds 41 2

Hunted and scavenged

vertebrates

   
Tree parts 34 0
Grasses and roots 13 34

Source: Harding 1981.

Some selectivity in diet is seen in its major components, with fruits, in-

vertebrates, and mature leaves being the most common items. Meat from

vertebrates, either hunted or scavenged, and tree parts (e.g., bark, cambium)

are not reported as major components for any non-human primate species.

Thus it might be best to characterize primates not as omnivores, but as

selective omnivores. There are several reasons for this selectivity. First, pri-

mates are, with few exceptions, diurnal and highly active. Second, primates

have brains that are about four times larger, relative to body size, than the

brains of other mammals. Third, primates have relatively long gestations

prior to birth and are nursed on demand for a relatively long period after

birth. Each of these traits places a high metabolic demand on an animal to

maintain activity, to supply the brain with energy and oxygen (the human

brain uses 20% of the body's energy and oxygen), and to meet the nutritional

needs of a mother primate and her fetus or infant.

Accordingly, primates must select foods that are dense in essential nutri-

ents. Fruits and invertebrates are such foods; fruits are dense in carbohy-

drates, minerals and vitamins, and invertebrates are rich in fats and proteins

(remember those grasshoppers!). Soft plant foods are mostly water and tree

parts are mostly cellulose or lignin, all of which are low in nutrients. Grasses

and roots are good foods for those species that live in savanna-woodland

habitats where grasses are abundant (e.g., baboons), however most primates

live in the tropical forests. Vertebrate meat and seeds are also nutrient dense,

but require hunting skills or specialized mastication or behavior to make use

of them. Of the 131 species surveyed, only some baboons and chimpanzees

regularly hunt mammalian prey (Strum 1981; Teleki 1981), and only two

monkey species, Cercopithecus neglectus and Colobus satanas, include seeds as

major foods. Chimpanzees, bonobos, and baboons have been seen to use

rocks to break open seeds to eat the contents, but this requires much effort

and time, which takes away from feeding on more easily acquired foods.

The human primate, not included in Harding's survey, is unusual in that

seeds, grasses, roots, and vertebrate meat are major components of both

modern and ancient diets. Seeds, grasses, and roots have their nutrients

protected by cellulose membranes that must be mechanically broken. This

can be done either by mastication or by using technology. People, and our

hominid ancestors dating back to Australopithecus, possess the anatomy (e.g.,

small canines, flattened molars, and enlarged pterygoid muscles-the mus-

cles that move the lower jaw from side to side) that allows for a type of

chewing called rorary grinding, which can break cellulose. People, and our

ancestors of the genus Homo, are also dependent on technology (e.g., tools,

fire) for food processing (Figure 6.1). Technology is also required for hunting

at a level that makes vertebrate meat a regular part of the diet.

A second reason for selectivity is the coevolution of primates and their

foods. Coevolution refers to the interactions of different species of living

organisms that exist in the same community, which result in genetic change

in those organisms over time. Predator-prey relationships are a common ex-

ample of coevolution. Animals can move, and animal prey may run, jump,

or fly away to evade capture. Over time, there will be selection for predators

that are better suited to capture their prey and selection for prey that are

better able to avoid capture. In contrast, plants are stationary but not de-

fenseless. Plants produce a host of noxious or toxic substances (called sec-

ondary compounds), such as tannins and alkaloids, to discourage their

predators. Plants may also evolve edible parts with low nutritional content

(Hladik 1981) or seeds and fruits with coverings too hard to pierce (Kinzey

and Norconk 1990), thus making those parts less attractive as food items to

primates. In a review of the literature on secondary compounds, Glander

(1982) found that the rich appearance of the tropical forest may be decep-

tive, for many primate species avoid a large percentage of potential plant

foods. Glander concludes that the selectivity of primates for plant species

and parts of plants must be viewed as a strategy balancing "the nutrient and

secondary compound content variation in these foods" (p. 1.).

A third reason for selectivity is that primates have a worldwide distribution

as an order, but are localized as genera into dozens of populations restricted

to species-specific habitats. Thus it is not surprising that many primates,

despite their evolutionary heritage of an eclectic food base, have, in practice,

species-specific diets. The tamarins and marmosets of South America, for

example, eat insects, fruits, and foliage, which are food items common to

most primate diets, but also require tree sap for survival. The tree sap is the

major source of calcium in their diet (Sussman and Kinzey 1984). These

primates have clawlike nails used to cling to tree trunks and procumbent

lower incisors used to gouge bark and release sap. No other group of pri-

mates has this set of anatomical specializations for tree sap consumption.

 

HUMAN DIET EVOLUTION

People also have unique requirements and specializations related to nu-

trition and diet. All primates require a relatively high-quality diet, but people

require a higher quality diet than any other species. Leonard and Robertson

(1994) compared the diet of 5 human foraging societies (!Kung, Ache, Hiwi,

Inuit, and Pygmies) to 72 nonhuman primate species and found that diet

quality of the human groups was almost twice that of other primates of the

same body size. The human ability to include seeds, roots, and meat in the

diet increases quality, as these are nutrient-dense foods. Building on the

research of Martin (1983), Leonard and Robertson show that the need for

high diet quality is a consequence of the human brain being several times

larger than expected for a primate their size. Using estimates of brain and

body size for extinct hominids, Leonard and Robertson estimate that hu-

manlike dietary requirements evolved with the appearance of the genus

Homo. But the only way to find out what our ancestors actually ate is to look

at the evidence, which comes from the study of remains of hominids and

their activities.

ARCHAEOLOGY AND FOSSIL STUDIES

Archaeological methods "include the identification of edible materials,

functional analyses of artifacts employed in food preparation, coprolite [fossil

feces] analysis, information on paleohabitat, and analyses of [hominid] skel-

etal material" (Sillen and Kavanagh 1982). Paleontological data are derived

from the kinds and percentages of fossil remains found at a site. Each type

of evidence contributes some knowledge, but each has serious limitations.

The association of hominid fossil remains with the skeletal remains of other

fossil vertebrates may result from geologic forces, such as rivers carrying dead

carcasses to a central location or a volcanic eruption burying simultaneously

a community of animals, rather than hominid food-gathering behavior. Early

speculation by Dart (1957) that the bone accumulations at the South African

cave sites of Australopithecus represented hominid hunting activity are now

considered incorrect. Rather, Brain (1981) argues that the fossil remains,

including the hominids, represent the activity of nonhominid carnivores, es-

pecially leopards, and geological forces. Dr. Brian aptly names his book on

this subject The Hunters or t/e Hunted, and his conclusion is that the early

hominids were the prey of the leopards.

Research conducted during the 1980s produced a 180-degree shift in the

fossil evidence for the evolution of human hunting. In The Descent of Man,

Charles Darwin (1871) proposed that hunting large game provided much of

the selection pressure for human evolution. That view persisted through the

196Os, and the book Man the Hzmfer (Lee and DeVore 1968) represented

majority opinion that uniquely human characteristics, such as bipedalism,

chapter by Washburn and Lancaster in that volume). Implicit in this argu-

ment is the notion that the type of diet consumed by human ancestors

played a significant role in the evolution of human biology and behavior.

This notion is reasonable, but the explicit assumption of carnivory and hunt-

ing became less acceptable as existing evidence was reevaluated and new

evidence discovered. The existing data, based on fossil and archaeological

remains and the study of living hunting and gathering people such as the

!Kung and Australian Aborigines, showed that gathering and processing plant

foods was the main activity of tropical foragers. Moreover, women in living

foraging societies provided most of the calories consumed by these people.

These observations turned "man the hunter" into "woman the gatherer,"

and the hunting hypothesis was attacked for both lack of data and its male-

biased implications (Zihlman 1981).

The new evidence is based on analyses of bone and stone tool material

associated with early hominids. Potts and Shipman (1981) used scanning

electron microscope images of mammalian long bones dating to 1.7 million

years ago to show that cut marks produced by stone tools were incised above

those made by carnivore teeth and the teeth of known scavengers, such as

porcupines. Assuming that the order of markings reflects the order of use

by hunters and scavengers, the hominids were the last to have at the

bones-even after porcupines. Subsequent analysis shows that hominids

may have been collecting bones for their marrow and brain tissue rather than

for any meat still remaining on the surface of the bone (Binford 1987). Mar-

row and brain are high in fat and protein, but few carnivores have the mor-

phology necessary to break open large long bones. Hyenas have the ability

to exploit marrow and are formidable predators and scavengers but are most

active at night (Schaller and Lowther 1969). Hominids are most active during

the day and thus could scavenge for carcasses with less threat from hyenas.

The invention of stone tools, first manufactured by hominids about 2.2 mil-

lion years ago, may have been a dietary adaptation for extracting marrow. At

Olduvai Gorge there are sites where the bones of large game animals (from

gazelles to elephants) are found together with stone tools. The tools are

called scrapers and choppers. Blumenschine and Cavallo (1992) report that

the bones are mostly from limbs and skulls and that these are precisely the

animal parts that only hyenas and tool-wielding hominids can crack open.

Further, they report that one-half hour's work with a chopper can yield

enough calories from the marrow and brain of a carcass the size of a wilde-

beest to meet an adult's daily energy requirements.

Hominids may also have scavenged for larger pieces of meat. Cavallo

(1990) studied the ecology and behavior of leopards in Tanzania. Most car-

nivores, such as lions and hyenas, leave their prey on the ground and con-

sume most of the internal organs and limb meat within a few hours after

 

prey over several days. The kill may even be left unattended for up to ten

hours, for other terrestrial carnivores ignore carcasses hanging in trees. Cav-

al10 believes that human ancestors may have scavenged these arboreal caches

of meat. This speculation is supported by the South African cave evidence

of Brain (1981), which shows that australopithecines and leopards Iived to-

gether and that the hominids were often the prey of the carnivores. Cavallo

argues chat by the time of the appearance of Homo, some hominids may

have reversed the predator-prey relationship. There are modern-day reports

of groups of baboons killing leopards as well as confirmed observations of

chimpanzees scavenging tree-cached leopard kills and taking and eating

leopard cubs (Cavallo 1990), stone tool-wielding hominids may have done

the same on occasion.

Perhaps it was the occasional (or regular?) consumption of leopard that

caused the hypervitaminosis A of the H. erectus individual from the Koobi

Fora formation, located on the eastern shore of Lake Turkana, Kenya. The

skeleton is dated to 1.6 million years B.P. (Walker et al. 1982), and analysis

indicates that it was female and has "striking pathology" in the long bones

of the limbs. These bones have a deposit of abnormal coarse-woven bone,

up to 7 mm chick in places, above the normal skeletal tissue on the outer

surface of the bone. Walker and his colleagues consider many possible causes

for this pathological bone growth and conclude that an overconsumption of

vitamin A (hypervitaminosis A) is the most likely cause. Similar cases of

hypervitaminosis A have occurred in arctic explorers who consumed the liv-

ers of polar bear and seal. The liver stores vitamin A, and the liver of car-

nivores, who are at the top of the food chain, usually contain the greatest

amounts of this vitamin. Walker et al. suggest that the cause of the bone

pathology in this specimen of H. erectus was due to eating the liver of car-

nivorous animals.

Despite rhe evidence for scavenging animal carcasses and, perhaps, prey-

ing on leopards, the bulk of the hominid diet has almost always been from

plants. The stone tools of the early hominids may also have been used to

process plant foods that were difficult to chew, such as seeds. Walker (1981)

and Kay (1985) studied the finer details of early hominid dental structure

and tooth wear using the scanning electron microscope and tooth wear ex-

periments. These researchers propose that the diet of the early hominids,

including Australopithecus and H. habilis, was largely herbivorous, including

softer plant foods (leaves, fruits) as well as the tougher seeds and tubers.

Given all of the evidence now available, perhaps it is safest to say that the

gathering of plants, insects, birds' eggs, and other relatively immobile foods

along with the scavenging of marrow from carnivore kills typified early hom-

inid food behavior.

The early hominid dietary pattern continues through H. erectus times. Bin-

ford (1987) reanalyzed fossil material from Torralba, a H. erectus site in Spain,

and Zhoukoudian, a cave site near Bejing, China, spanning the period from

H. erectus to H. sapiens. During the H. erectus period of occupation (250,000

to 450,000 years B.P.), both sites show evidence of the gathering of plant

foods and scavenging, rather than hunting. The animal bones at these sites

appear to have been processed and consumed on the spot, rather then car-

ried to any sort of "base camp." If this is so, then past theories about the

evolution of human biology and behavior-including bipedalism, large

brains, division of labor, sharing, and intense parental investment in off-

spring-that depended on hunting and "family style dining" at home bases

have to be rejected. Binford (1984) states that convincing evidence for the

regular hunting of big game does not appear in the fossil record until 90,000

years B.P. at the earliest.

H. erectus added fire to its repertoire of technology. Fire, which may have

been used as early as 1.4 million years ago and was certainly controlled by

750,000 years B.P., provided warmth, light, protection, and a new way to

process foods. Where and how cooking was invented is a matter for specu-

lation. Cooking, by roasting or boiling, increases the nutritional benefit of

many vegetable foods by helping to break down the cellulose in those foods,

which is undigestible to people. Fire may be used to open large seeds that

resist even stone tools. Cooking, especially drying or smoking, helps to pre-

serve foods for storage. Fire may also be used to obtain foods, especially

when used to drive game toward a convenient killing site. All of these uses

of fire did not appear simultaneously, and many appear to be the invention

of H. sapiens rather than H. erectus. What is certain is that the controlled use

of fire was a significant addition to hominid technology with profound con-

sequences for nutritional status.

FOSSILIZED FECES

Coprolite analysis might seem to provide unequivocal evidence of dietary

habits, but it too is subject to misinterpretation. First, the coprolite must be

identified as unambiguously being from a hominid. Second, coprolites can

only verify that a particular substance was eaten. That substance may or may

not have been a food item itself, it may have been ingested coincidentally

along with a food, such as a seed or insect clinging to an animal or plant.

Third, only undigestible substances will be found in feces and those sub-

stances must be suitable candidates for fossilization to be preserved in a

coprolite. Thus, coprolite analysis may provide a very biased picture of the

true dietary intake. Even so, considerable information has been obtained

about the diet of prehistoric humans, and limited information about the diet

of hominid species ancestral CO modern people, using coprolite analysis. The

animal affinity of desiccated coprolites can be determined by placing the

specimen in a trisodium phosphate solution for 72 hours. Human coprolites

produces this effect (Bryant and Williams-Dean 197.5). Other characteristics

of human feces are inclusions of charcoal and the presence of undigested

animal parts from a wide variety of species. Charcoal comes from cooking

food over a wood fire. Since people cook their food and other animals do

not, the presence of charcoal in feces is indirect evidence of a unique human

behavior. People also have an eclectic diet compared to most other mam-

mals, so undigested parts from a wide variety of species is another indicator

of the human affinities of a coprolite. More than 1,000 paleoindian coprolites

from the American southwest have been identified and analyzed. One group

of specimens nas collected from Texas sites that date from 800 B.C.E. to 500

C.E., representing the temporary camps of hunting and gathering peoples

(Bryant 1974). By comparing the pollen content of the coprolites with that

of the adjacent soils, it was determined that the people had consumed high

quantities of flowers. Because the physical characteristics of flower pollens

are unique to each species, it was possible to determine that flowers of agave,

sotol, yucca, prickly pear cactus, gilia, and leadtree were popular foods. Also

found were remains of wild onion bulbs, bark, grasshoppers, fish, small rep-

tiles, and snails. Although not the current cuisine of Texas, this diet is typ-

ically human in its diversity of species. The flower pollen even provides a

time frame for the occupation of the sites of spring and early summer.

Coprolites from paleoindian sites in New Mexico, Arizona, and Texas

contain pollen from plants of known pharmacological value, suggesting that

people have a long history of consuming plants as medicines as well as foods

(Reinhard 1989). Willow, an analgesic with essentially the same active in-

gredient as in aspirin, Ephedra, an antihistamine, and creosote, an antidi-

arrheal, are the most concentrated pollens in the samples. Ethnographic

evidence shows that these three species were, and are still, widely used as

medicines by Native Americans (Moerman, 1986, 1989). Willow tea is used

for the treatment of many aches and pains, Ephedra tea is prescribed for the

stuffy noses of the common cold, and creosote is indicated for any type of

loose bowls. Reinhard (1989) states that the analysis of these coprolites

"demonstrates the antiquity of folk remedies and provides circumstantial

evidence of certain disorders suffered by prehistoric peoples" (p. 2).

The oldest verified coprolites of a hominid species are from the H. erectus

site of Terra Amata located on the French Mediterranean. These coprolites

may be as old as 300,000 B.P, and they are heavily mineralized. They have

only a slight reaction to trisodium phosphate rehydration (Bryant and Wil-

liams-Dean 1975). The specimens contain sand grains, charcoal, and mollusk

shell fragments. The sand and shell are expected, since Terra Amata is a

beachfront site, and the charcoal helps establish that foods were cooked

before consumption (maybe evidence for a prehistoric clam bake!).

modern (e.g., modern people have more gracile skeletal features than archaic

forms). Care was taken to control for differences in the amount of calcium

and strontium in the soils of different fossil sites and other confounding

geological variables. It was found that Sr/Ca ratios in bone increased with

time, suggesting more plant food in the diet, but the increase occurred

20.000 vears after the modern human form appears in the fossil record.

Schoeninger concludes that the morphological transition from archaic to

modern H. sapiens was not due to the utilization of new foods, rather it was

due to "alterations in the means of procuring or processing the same kinds

of foods that had been utilized earlier in time" (p. 37). In other words, be-

havioral and cultural changes were more important than diet change per se

in bringing about the biological form of modern humans. This along with

the other examples of trace element and stable isotope analysis clearly shows

the biocultural nature of people and food.

STUDIES OF LIVING HUNTERS AND GATHERERS

Today, 99.9 percent of people derive their food from some form of agri-

culture. However, from the time of the Australopithecus until about 10,000

years ago, a time period that covers 99 percent of human evolution, all hom-

inids lived by foraging-the gathering, scavenging, and, more recently, hunt-

ing of wild foods. Most human physical traits, and perhaps many behavioral

Propensities, evolved during the time that hominids lived as hunters and

gatherers. That biobehavioral evolution includes current human dietary re-

quirements, adaptations for food acquisition and processing, and biocultural

responses to food intake. Studies of the few remaining cultures of hunting

and gathering peoples offer an indirect view of that style of life, now nearly

extinct. These ethnographic and ecological studies complement the infor-

mation derived from paleontological and archaeological sources.

Foragers are a diverse group geographically and culturally, ranging from

the arctic Inuit and Eskimo, to the tropical forest Ache (Paraguay), to the

dry scrub San (Africa) and the desert Australian Aborigines. Yet research

shows some consistencies in behavior and diet. The diversity of food re-

sources utilized is high among gathering and hunting peoples compared with

agriculturalists. The !Kung San of southern Africa, for instance, eat 105 spe-

cies of plants and 144 species of animals (Lee 1984). The Australian North

Queensland Aborigines exploit 240 species of plants and 120 species of an-

imals (Gould 1981). The Ache forage on fewer species, about 90 types of

plants and animals (Hill and Hurtado 1989). Even the Dogrib, residing in

the subarctic of Canada, gather 10 species of plants and 33 species of animals

(Hayden 1981). That is a small food base for hunters and gathers, but still

a large number relative to agriculturalists who, on a worldwide basis, subsist

largely on four species of plants and two species of animals. Of nine species

of staple plant foods, wheat, rice, potatoes, and maize together account for

Table 6.3 Hunters and Gatherers

Group (location) Group size Population density/per 100 mi2 Freq. Of moves
Nootka (Canada) 1500 200 180
Andamanese (Asia) 45 200 60-180
Paliyans (India) 24 200 As needed=45 days?
!Kung San (Africa) 20 41 14-21
Hazda (Africa) 9 40 14 days
Giwi San (Africa) 55 16 21 days
Ache (S. America) 48 8 Daily,143 days
Guayaki (S. America) 16 7 3 days
Western Desert Australia 20 3 7-14 days
Mistassini (Canada) 15 1 180 days

Source: Hayden 1981. Ache data from Hill and Hurtada, 1989.

1,680 of the 2,284 million metric tons consumed (sorghum, sweet potatoes,

barley, millet, and cassava are the other five staples, [Garine 19941). Of the

animal foods, cattle and hogs account for 80 out of every 100 metric tons of

domesticated animal meat. Poultry, lamb, goat, buffalo, and horse make up

the bulk of the remaining 20 metric tons (Bogin 1985).

- A second common feature is that gathered foods (plants, insects, birds'

eggs, turtles, etc.) are the primary subsistence base in most foraging societies.

Lee (1968) compared 58 forager groups and found that the primary subsis-

tence source was gathering for 29, fishing for 18, and hunting for 11. Ten

of the hunting groups and 16 of the fishing groups lived north or south of

the 40-degree parallel. Thus, not only is gathering the most common sub-

sistence pattern, it is correlated with tropical, subtropical, and low-temperate

habitats. Such habitats were the home for all species of hominids until the

middle to late Paleolithic period.

Often the use of many species for subsistence is correlated with the high

diversity, low density, or seasonality of food items in the environment. In

habitats where low density is combined with the wide dispersal of foods,

foragers must be mobile and live in small groups. Thus a small, mobile social

group is a third typical feature of forager societies, but, as shown in Table

6.3, it is not a universal feature. Leaving aside the Nootka, average group

size ranges from 9 to 55 and average densities range from 1 to 200 people

per 100 square miles. Mobility ranges from daily movement from camp to

camp in the case of the Ache to seasonal sedentariness at one camp (e.g., a

winter lodge) in the case of the Mistassini (hunters of the Canadian boreal

forests).

The Ache are unusual in their daily movement, but contemporary Ache

Table 6.4

The Nine Universal Features of Human Nutrition and Food Behavior and

the Sources of Evidence Used to Study Their Evolution

Universal features

Sources of evidence

1. Large number of essential nutrients Primate studies, biomedical research
2. Each culture has a cuisine Archaeology, ethnography
3. Extreme omnivory Primate studies, hunters& gatherers
4. Transport of foods Archaeology, ethnography
5. Storage of foods Archaeology, ethnography
6. Complex technology for acquisition & preparation Archaeology, ethnography
7. Sharing and division of labor Primate studies, hunters& gatherers
8. Food taboos Ethnography
9. Non-nutritional use of potential foods Archaeolog, ethnography

SUMMARY OF EVIDENCE FOR THE EVOLUTION

OF HUMAN NUTRITION

Table 6.4 lists the nine universal features of human nutrition and food

behavior. Also listed in the table are the sources of evidence that allow an

understanding of the origin and function of these universals. The human

place in nature as primates explains our broad requirements of essential

nutrients. Fossil and archaeological evidence accounts for the development

of cuisines and the technology for food acquisition, preparation, and storage.

The study of living hunting and gathering peoples complements and sup-

ports these other sources of evidence. Five features of food and behavior

are typically found in hunting and gathering societies: (1) a high diversity

of food types; (2) greater dependence on gathering over hunting; (3) small,

mobile social groups; (4) dependence on technology for acquiring and proc-

essing foods; (5) and division of labor and sharing. Additionally, forager stud-

ies detail the nature of human food transport for exchange and sharing and

provide some information on the origin of food taboos and nonnutritional

uses of foods.

"CAVEMAN CUISINE"

Using the methods of research described here, archaeological and pale-

ontological evidence, ethnographic studies of living hunting and gathering

people, and the nutritional analysis of wild plant and animal food, Eaton

and Konner (1985) reconstructed the diet of Paleolithic people living during

 

Table 6.5

The Paleolithic Diet of 15,000 Years B.P., the Current American Diet, and

One Set of Dietary Recommendations for the United States

Daily Intake Paleolithic American' Recommended2
Daily intakes M W
Total dietary energy (%)
Protein 34 15 15 12
Carbohydr. 45 48 50 58
Fat 21 34 34 30
Alcohol Trace 4 2 --
P.S. ratio 1.41 0.56 .61 1.00
Cholesterol[mg] 591 395 244 300
Fiber [g] 46 19 13 30-60
Sodium [mg] 690 4659 3002 2000
Calcium [mg] 1580 1075 778 1200
Ascorbic acid [mg] 392 121 87 60
Simple sugars [g] Trace =154 =154 =50

`.Amcrican diet for adults E-29 years old surveyed between 1988 and 1991, published by the

Centers for Disease Control, 1994: hl = men, W = women. Based on a diet of 3,025 cal-

orirs per day for men and 1,957 calories per day for women.

`Recommendations of I1.S. Senate Select Committee and Food and Nutrition Board, National

Academy of Sciences. Values for adults, sexes combined.

`Ratio of polyunsaturated to saturated fats from ail foods.

Soure: Paleolithic diet: Eaton and Konner 1985: American diet: Guthrie and Picciano 1995,

appendix K.

the last glacial period in western Europe, about 15,000 years ago. Garn and Leonard (1989)

point out that this diet was not typical of the majority of

people alive at that time. The majority lived at tropical or subtropical lati-

tudes and consumed more wild grains and less animal meat-recall Schoen-

inger's (1982) analysis of Paleolithic diets from Israel. Garn and Leonard

state that "many of our ancestors ate poorly,. . . and they were often at risk

for vitamin deficiencies, food-borne diseases, and neurotoxins" (1989: 337).

Despite these caveats, the Paleolithic diet reconstructed by Eaton and Kon-

ner offers some useful data, especially when compared with the modern

American diet.

Table 6.5 compares this Paleolithic diet with that of modern Americans

and U.S. government recommendations for a safe and healthy diet. These

glacial people ate more protein and less fat than we do. Eaton and Konner's

analysis of living foragers indicates that the average diet consists of 3.5 per-

cent of calories from meat and 65 percent of calories from vegetable foods.

Although plants contribute protein to the diet, Eaton and Konner estimate

that most of the protein was from animals, including fish, insects, and other

invertebrates. Fat intake was lower in the Paleolithic due to the low content

of fat in wild game. The average carcass fat content of 15 species of wild

herbivore surveyed by Eaton and Konner is 3.9 percent compared to an

average of 25 to 30 percent in domesticated carcasses (cattle, hogs, etc.).

Moreover, compared with domesticated meat, the fat of wild game is about

five times higher in the polyunsaturated form. Along with plant foods rich

in polyunsaturated fats, the paleolithic diet has a high ratio in polyunsatu-

rated to saturated fats. Cholesterol intakes appear to have been higher in

the ancient diet. In the early 1980s. however, the ancient and modern diets

contained similar amounts of cholesterol, and fat intake of the American diet

was about 42 percent of total calories. It seems that Americans have learned

to eat foods with less fat and cholesterol. Unfortunately, Americans are also

eating more total food (i.e., calories) than in 1980. So, despite a decrease in

fat consumption there has been an increase in body weight, mostly due to

fatness, for both average and obese children and adults (Yip and Scanlon

1995).

The modern ethnographic data and the archaeological data indicate that

Paleolithic people would have gathered a wide assortment of wild plant

foods and many species of animals, ensuring variety in both vitamin and

mineral content and in taste and appearance. In contrast, the many people

living in agricultural and industrial societies eat from an extremely narrow

range of food options. Modern people eat more wheat, rice, potatoes, and

maize than the next 26 most often consumed plants combined (when did

you last eat a turnip?). There are many reasons for this. For most of the

world the reasons are associated with economics. Wheat, rice, potatoes, and

maize are grown on a large scale to make profits for national or multinational

agribusiness corporations. The intensive production of these and a few other

crops is very efficient in terms of business practices and profits. These crops

may be produced relatively cheaply, making them more affordable to the

80 percent of the world population who live in less-developed nations and

who are at or near the poverty level. The more affluent 20 percent of the

world's people, living mostly in the more-developed nations of Europe, Ja-

pan, Australia, and North America, can eat a much wider variety of foods.

They can afford to buy more expensive foods that are produced in limited

quantities and often shipped long distances. The Japanese and French, for

example, eat dozens of species of animals, including many mammals (rabbits,

sheep, horses) and ocean species (shell fish, sea urchins, fish) as part of their

regular diet. The modern American diet, on the other hand, is much more

restricted in food choice. A survey of most supermarkets will reveal a limited

variety of animal protein sources (when did you last eat horse, rabbit, or

even fish?). Americans eat more ground beef, usually in the form of ham-

burgers, than any other animal protein food. It is estimated that 12,000 ham-

burgers are consumed each minute in the United States-a rate of 200 per

second (Lieberman 1991). The consumption of ground beef also accounts

for about one-half of the total fat and three-fourths of the saturated fat in

the American diet. Surveys of food consumption generally find that the uni-

formity and limited variety of offerings available at fast-food restaurants de-

picts the current American diet very well.

The Eaton and Konner reconstruction also indicates that our ancestors ate

much more fiber, calcium, and vitamin C, but far less sodium. Our ancestors

ate simple sugars only in natural forms, for instance from fruits, but today

we each eat about 124 pounds of simple sugar (mostly sucrose and corn

syrup) a year. Paleolithic foragers consumed no dairy products, except for

mother's milk during infancy, and little alcohol. Consumption of dairy foods

is a by-product of animal domestication, and thus less than 12,000 years old.

Even today, dairy products are staple foods in only a few societies. Most

adults lack rhe enzyme lactase needed to digest the milk sugar lactose

(Kretchmer 1972). Some societies do eat cheeses or yogurt, for these foods

have their lactase digested by bacteria, but these foods are often too expen-

sive for most of the world's people. The wealth of the developed nations

allows [their populations to consume large quantities of dairy products. In

the [United States, 20 percent of protein, and most calcium, comes from milk

and cheeses. These foods, unfortunately, also contain high amounts of fat

and sodium, which also typify the American diet. Alcohol is, basically, also

a product of domestication-the domestication of grains such as maize and

barley. Some alcoholic beverages (beer, mead, wine) contains some nutri-

ents, but alcohol itself provides no essential nutrients. Even so, alcohol con-

tains energy-seven calories per gram-and contributes a measurable

percent of calories to the American diet.

High intakes of fat, especially saturated fat, simple sugars, sodium, and al-

cohol are linked with health problems, such as obesity, heart disease, and liver

disease. Apparently, Americans are aware of a problem, for example, millions

of dollars are spent annually for weight loss programs. The weight loss pro-

grams usually have little long-term success, indeed the average weight and

fatness of Americans has increased in the last few decades. A narrow diet is

also a risk for nutrient deficiencies, such as the low calcium intake of many

Americans. The response of the people is to purchase nutritional supplements

and special "health foods." The U.S. Food and Drug Administration (note

that food and drugs are lumped together by American culture) estimates that

40 percent of adults regularly take at least one vitamin and mineral product

(Moss et al. 1989). Taking vitamin pills to compensate for a narrow diet was

not the nutritional behavior followed by our ancestors. Consumption of vita-

mins via pills is not bad, or harmful, in and of itself. Our bodies make no dis-

tinction between the vitamins in food and those synthesized by chemists.

Eating food, however, provides the physical and emotional satisfaction of

taste, aroma, and a full stomach, along with other nutritional factors that are

linked with good health, such as fiber. Of course, one can also buy and con-

sume "fiber pills." Indeed, there are pills to reduce high serum cholesterol

levels, induced by all those hamburgers, and to lower blood pressure, induced,

in part, from our sodium overload. Perhaps it is best to view American nutri-

tional habits as part of a biocultural system with its own internal logic. The

system works, in the sense that there are fewer nutritional deficiency diseases

in the United States today than ever before.

Nutritional oversufficiency is the single biggest dietary problem in the

United States. In this regard, it is important to mention one other major

difference between Paleolithic people and contemporary Americans. The

former were required to perform high levels of both aerobic and anaerobic

exercise, while the latter are sedentary ("couch potatoes" in the vernacular).

Foraging people of the past and of today have the physique and cardiovas-

cular conditioning of athletes. No amount of diet change or pill consumption

for Americans and people of the other developed nations will improve health

without a simultaneous increase in physical activity.

 

AGRICULTURE AND THE DECLINE OF HUMAN

NUTRITION AND HEALTH

Some of the nutritional problems of modern world societies are: (1) a

narrow food base, leading to deficiencies for some essential nutrients, (2) an

inadequate supply of energy (i.e., undernutrition of total calories from all

food sources) for about 60 percent of the world's population, especially the

poor in the least-developed countries, and (3) an oversupply of energy, lead-

ing to obesity and related diseases in the developed nations and, increas-

ingly, among the more affluent in the developing nations. The immediate

causes of these problems include a host of social problems, such as poverty

and other economic inequalities, political unrest (such as civil and ethnic

wars), inadequate water management, and unregulated population growth.

Although these are significant proximate causes for the world's current nu-

tritional crisis, there is a more fundamental explanation that had its origin

at the end of the Paleolithic period.

The major culprit of the nutritional dilemma is agriculture. More recently,

industrialization and urbanization have compounded the effects of agricul-

ture on the nutritional status and health of human populations. Agriculture,

industrialization, and urbanization are often stated to be the hallmarks of

"progress" of the human species. Although progressive in a technological

sense, each of these achievements has had negative consequences for human

nutrition and health.

There is much evidence from the developing nations of the world that

the food production systems of rural people correlate strongly to their nu-

tritional status. The classic study of Indonesian peoples by Geertz (1963)

shows that simple horticulturalists have the most abundant variety and

amount of foods, while food shortages and frank malnutrition are most com-

mon in areas of intensive rice agriculture. Whyte (1974) extended these

findings to much of tropical Asia. Whyte's analysis shows that foragers, hor-

ticulturalists, and fishing societies have diversified diets, but often inade-

quate calorie intakes. These societies are better nourished, however, than

peoples practicing mixed agriculture-pastoralism and intensive irrigation ag-

$eulture, especially of rice. The agriculturalists suffer marginal to serious

malnutrition for total calories and many vitamins, minerals, and protein.

The dilemma of modern agricultural societies has deep historical roots.

Studies of archaeological populations show that several indicators of biolog-

jal stress increase with the transition from foraging to horticulture and ag-

riculture (Cohen and Armelagos 1984). These stress indicators include bone

lesions due to anemia (called porotic hyperostosis), deficits in enamel for-

mation in teeth (hypoplasias), loss of bone tissue from the skeleton, bone

lesions due to infectious disease, such as tuberculosis (called periosteal re-

petjon), and reduced skeletal growth in children and adults (Goodman et al.

1988). The Dickson Mounds site of the Illinois River Valley provides a

classic example. From C.E. 950 to 1300 the human population of that area

hanged from mobile foragers to sedentary intensive agriculturalists. During

"this short time period, "the shift in subsistence led to a fourfold increase in

iron deficiency anemia (porotic hyperostosis) and a threefold increase in in-

fectious disease (periosteal reaction). The frequency of individuals with both

iron deficiency and infectious lesions increased from 6% to 40%" (Goodman

et al. 1988~180).

' The incidence of enamel hypoplasias (malformations of the tooth crown

that include pitting, linear furrowing, or complete lack of enamel) also in-

creases from the forager to agricultural period. These dental deficiencies

occur when malnutrition or disease disrupt the secretion of enamel-forming

material. For the permanent teeth that process takes place during infancy

and childhood. Thus enamel hypoplasias leave a permanent record in the

teeth of nutritional or disease stress that people experienced in early life. In

the Dickson Mound skeletal material the prevalence of hypoplasia increases

with time, going from 45 to 80 percent of individuals affected. Furthermore,

the number of hypoplasias correlates to mortality. Individuals with one hy-

poplasia died, on average, five years earlier than people with no hypoplasias.

With two or more hypoplasias age at death was reduced by nine years (Good-

man et al. 1988~180). Since peoples' teeth form in a fixed pattern that is

Virtually the same in all human beings, it is possible to correlate the frequency

of hypoplasias found on different teeth to the age of the individual

n the disease stress occurred. At Dickson Mound that correlation indi-

cates that infants and young children were especially subject to health stress

at the age of weaning- (about two to four years of age). Deficiencies in the

weaning diet, combined with increased exposure to infections and other

diseases at the time of weaning, were very likely the cause of the hypoplasias

Goodman et al. 1988).

With the development of agriculture the Dickson Mound people shifted

from a diverse food base to one dependent on maize. The emphasis on

monoculture reduced the supply of essential nutrients, especially amino ac-

ids and vitamins not found in maize, and this compromised the health of

the people. Compounding these nutritional problems was rapid population

growth. Despite a lowering in the average age at death at the Dickson

mound site, population sizes increased due to a shorter interval between

births (about two years as compared with four years in forager populations).

Larger populations and sedentarism gave rise to the conditions favorable for

the spread of infectious disease, and the poor nutritional state of the people

made them more susceptible to these diseases.

POLITICAL ECONOMY, FOOD, AND HEALTH

There are other archaeological examples of decline in human health with

the spread of agriculture, including sites in Africa, the Middle East, Latin

America, and Asia. But, not all people living in agricultural societies suffered

health problems. The social organization and political economy of each so-

ciety played a major role in the distribution of resources, especially food and

health care. One example comes from nearly three decades of research on

ancient and medieval Nubia (Van Gerven et al. 1995). Nubia is region of

the Nile River valley from southern Egypt to northern Sudan, bounded by

the First Cataract at Aswan to about the Fourth Cataract in modern Sudan.

For the past 5,000 years the people of Nubia lived by the agricultural pro-

duction of "sorghum, millet (locally known as dura), barely, beans, lentils,

peas, dates, and wheat. In addition, a few cattle, sheep, and pigs were kept,

but animal products appear to have been a minor part of the Nubian diet"

(ibid: 469). While the dietary base remained stable for millennia, the political

base of Nubia changed many times. From about 350 B.C.E to C.E. 350, called

the Meroitic period, there was political unification of all Nubia under a cen-

tralized, militaristic state society. The Meroitic state had great wealth, great

urban centers, but also great social stratification. The following Ballana pe-

riod (ca. C.E. 350-550) was politically decentralized, with people living in

smaller but more self-sufficient settlements. Overall, health status, as re-

vealed from skeletal and dental indicators, was better in the Ballana period

than in Meroitic times. People also lived longer during the Ballana period.

By the end of sixth century AD. Nubia was again unified under a series

of Christian kingdoms. The Christian period ended in C.E. 1365, following

the ascension of a Moslem prince to the throne in C.E. 1323. Van Gerven

and colleagues analyzed skeletons from a Christian period cemetery located

in the town of Kulubnarti, near the Dal Cataract. The early Christian period

was highly centralized and socially stratified, and the people of Kulubnarti

"were but a small and contributing satellite to a centralized and distant

authority" (Van Gerren et al. 478). The people contributed taxes in the form

of food surplus and labor. By the late Christian period, the central state

authority was in decline and satellite communities like Kulubnarti were es-

sentially ignored and independent. The populous of Kulubnarti reverted to

subsistence agriculture, local political control, and were free of taxation. The

skeletons from the cemetery show that infant and child mortality was

greatest during the early Christian period. Enamel hypoplasias occurred at

earlier ages and were more frequent. and there was more evidence of anemia

during early Christian times. Late Christian-period people suffered from all

of these indicators of poor health too, but less so. All of this shows, again,

that the political economy of a society interacts with the food base to shape

the pattern of health of the people.

 

CONQUEST, FOOD, AND HEALTH

Colonization of the New World, Africa. and Asia by the Spanish and other

Europeans introduced new plants. animals. foods, and diseases. Europeans

also introduced a new political economy. Generally, the diet and health of

native people suffered. A book edited by Larsen and Milner ( 1994) provides

up-to-date reviews of the biological effects of New World conquest. The

colonizers, however, also suffered from the introduction of new foods to their

diets. Pellagra is a nutritional disease caused by a lack of niacin (vitamin B3,).

The word pellagra is Italian, meaning rough or painful skin, and was used

to describe the disease when it first appeared in the eighteenth century in

chat country. In Spain the same disease also appeared at that time, but it

was called mal del sol ("sun disease"). Pellagra's classic symptoms are the

four D's-dermatitis, diarrhea, depression, and dementia. The early symp-

tom of light-sensitive dermatitis gave the condition its Spanish name. How-

ever, sunlight only aggravated the real cause, a diet based on the

consumption of highly refined maize. Maize was domesticated in the New

World and exported to Europe after contact with native American people.

Maize grew well in Europe and quickly became an abundant and inexpen-

sive food that replaced many traditional grains. This was true especially in

rhe diet of the poor of southern Europe which. by the 17OOs, was predom-

inantly based on maize, molasses (derived from sugarcane, which is of Asian

origin), and salt pork.

Maize is naturally low in the amino acids lysine, tryptophan, and cystine,

and in the vitamin niacin. Molasses and salt pork are also deficient in these

same nutrients. Milling the maize removes the husk and germ, further re-

ducing the niacin content, from 2.4 mg per cup to 1.4 mg per cup. The

minimum daily need for niacin in adults is set at 13 mg per day by the

World Health Organization. For cultural reason, Europeans preferred the

bleached white appearance of the highly milled maize, as it imitated the

more expensive wheat flour consumed by the wealthy. For the poor, who

followed a monotonous diet based on maize flour, pellagra was the result.

Pellagra spread to the United States as Europeans and their diets became

the dominant cultural force. It was confined to cotton producing and cotton

[SOME TEXT IS LEFT OUT DUE TO LENGTH]

 

indoor air of homes was established. In 1978, due to an excess of miscarriages

by women from Love Canal, the state of New York evacuated 235 families.

In 1980 the Federal government of the United States evacuated the re-

maining 800 families.

Prior to the 1980 evacuation, Paigen and colleagues measured the height

and weight of 921 children between the ages of 1.5 and 16.99 years from

424 households of Love Canal. A second control sample of 428 children from

Niagara Falls was also measured. The children of the control sample were

from homes in noncontaminated neighborhoods but similar to the Love

Canal sample in terms of socioeconomic and ethnic background. No differ-

ence in weight was found between the Love Canal and control samples.

However, children born and residing in Love Canal for at least 75 percent

of their lives were significantly shorter than the children from Niagara Falls.

That difference could not be accounted for by statistically controlling the

effect of parental height, socioeconomic status, nutritional status, birth

weight, or history of chronic illness. The authors of the report conclude that

chronic exposure to the toxic industrial wastes is a likely cause of the growth

retardation of Love Canal residents.

One does not have to live on top of a toxic waste dump to suffer the

effects of industrial pollutants. Lead dust in the air, water, and soil of in-

dustrialized countries affects virtually all of the population (Schell 1992).

Toxic levels of lead causes neurological disease in both children and adults.

The lead dust comes from a variety of sources, especially leaded gasolines

and water pipes joined with lead solder. The nutritional impact of lead is

that food grown in areas with high levels of automobile traffic may be con-

taminated with lead from the soil. Even worse, anyone living in an older

home is at risk for lead in drinking water. Even though lead was virtually

removed from gasolines in the 1980s and banned from plumbing solder in

the United States in 1986, the accumulation of lead in water and soils is so

great that all Americans have measurable lead accumulations in their bodies.

DIET AND THE DISEASES OF MODERN LIFE

After the year 1900, the affluent people of industrial areas in Britain and

other Western nations begin to achieve adult heights greater than those of

people in rural areas. The technology of Western nations, including efficient

transport of regional and nonnative foods, refrigeration, nutritional supple-

mentation, treated water, and public sanitation, allowed their people to over-

come some of the nutritional and health deficits of agriculture,

industrialization, and urbanization (Bogin 1988b). Other evidence for the

improving conditions of life in developed nations comes from mortality sta-

tistics. Prior to 1850, deaths from epidemic diseases were a leading cause of

mortality in the cities of Europe and North America. Death rates were so

high that urban populations required massive migration from rural areas just

to maintain constant numbers (McNeil1 1979). Between 1850 and 1900 death

rates in urban and rural areas began to equalize, and since 1900, urban mor-

tality rates, for all ages, have been lower, generally, than rural mortality rates.

The process of modernization in developed counties that resulted in better

physical growth and lower rates of mortality for urban populations is now

taking place in developing nations. Conditions for life in many Third World

cities are still abominable for the poor underclass. However, current trends

in some Third World countries indicate that these nations may follow the

historical path toward modernization that occurred in Western nations (Bogin

1988a).

Improved physical growth and longer life do not mean that modern urban

populations are free of the specter of malnutrition and disease. Rather, as

the threat of undernutrition and infectious disease relaxed, a suite of new

diseases related to the diet and life-style of industrialized/urbanized people

developed to burden modern affluent people. Data from 1990 show that

cardiovascular disease and cancer are the two leading causes of death in the

United States. They are followed by cerebrovascular disease (e.g., stokes),

accidents (mostly motor vehicle and firearm accidents), pulmonary diseases,

and diabetes. The literature on the relationship of diet to heart disease,

cancer, and diabetes is abundant and controversial. Alcohol, a dietary com-

ponent as well as a drug, contributes substantially to accidents and many of

the other leading causes of death.

The increase of cardiovascular disease in developed nations in this century

can be linked to diet in several ways. An intriguing hypothesis to account

for the epidemic of heart disease that occurred after World War II is pro-

posed by Barker (1992). Using both geographical analysis and the health

history of thousands of individual people, Barker and his colleagues show

that babies with some indication of growth retardation (low birth weight,

small but normal size, or small head circumference) have a higher risk of

cardiovascular disease as adults. Growth retardation at birth is very often a

nutritional problem. Either the mother is poorly nourished, or, despite ad-

equate maternal nutrition, not enough nutrients cross the placenta to the

fetus.

Poor nutrition during childhood may add to the risk of heart disease. Fe-

llague-Ariouat and Barker (1993) interviewed women from England and

Wales aged 80 years or older about their food habits when they were 10 to

1.5 years old (spanning the years 1899 to 1924). All of the women grew up

in families of lower economic status. The women who lived in areas that

today have low cardiovascular mortality tended to be rural, "to eat four meals

a day rather than three, to live in households which had gardens, kept hens

or livestock, and to go into domestic service, where diets were generally

good. Those who grew up in areas which have high cardiovascular mortality

tended to eat less red meat, live in houses without gardens, to enter indus-

trial occupations and have higher fertility rates" (p. 15). The high mortality

[BREAK IN TEXT HERE]

Nubian mummy (D. Moerman, personal communication). A carcinogen (can-

cer-causing substance) is usually needed to provoke a cancer. Many carcin-

ogens are industrial products, but some are found naturally in foods. Cancer

rates for modern people change as food preferences change. Takasaki and

colleagues (1987) studied the rates for different types of digestive system

cancer in the Japanese population during the period 1950 to 1983. The re-

searchers argue that between 40 to 60 percent of the incidence of cancer

may be attributable to diet, especially in those cases in which carcinogens

come into direct contact with the gastrointestinal tract. In the earlier years,

the typical Japanese cuisine was based on rice seasoned with highly salted

condiments, some green and yellow vegetables, and very little milk or diary

products. Takasaki et al. report that this type of diet is linked with stomach

cancer, and stomach cancer is the most frequent type of digestive system

cancer in Japan. During the 1960s the postwar industrialization and eco-

nomic recovery of Japan proceeded rapidly. One of the consequences of that

expansion was a shift from the traditional cuisine to one that included sig-

nificantly more dairy products and fat. Between 196.5 and 1983 rice con-

sumption dropped from about 300 gm per person per day to about 200 gm/

person/day. Milk and dairy products increased from about 75 to 150 gm/

person/day. Diets high in fats are associated with cancers of the intestine

and colon. In Japan, mortality rates (age adjusted) per 100,000 population

for stomach cancers dropped from about 37 to 22, while the same rates for

intestinal cancer increased from about 2 to 4.5.

Epidemiological studies such as Takasaki and colleagues' Japanese re-

search show many links between cancer and specific foods or cuisines. The

same holds true for heart disease, diabetes, and the other diseases of modern

life. This research negates the belief that these diseases are the natural

consequence of aging. Rather, these diseases are potential indicators of the

environmental quality of life and the well-being of human populations. The

fact that heart disease, cancers, and diabetes are the major causes of death

in developed nations belies the notion of "progress" that is a central belief

of European and American culture. While it is true that the average age at

death has increased steadily in developed nations in this century, most of

the increase is due to the control of infant and child mortality from infectious

disease. Adults suffer as much, or more, disease than 100 years ago, and may

suffer these diseases from an earlier age.

Food is safer today than 100 years ago in that processing, refrigeration,

and other technologies prevent spoilage and food poisoning. However, the

processing that increases short-term safety also adds salt, sugar or artificial

sweeteners, and, often, fats (both natural and artificial) to our diet that may

lead to long-term health risks. Advances in food technology also permit pro-

ducers and consumers to eat more preferred foods. Preferred foods may be

"good" or "bad" for people depending on the scientific, philosophical, and

moral code of a society. Many Americans, for example, abhor horse meat

 

Table 6.6

Energy Inputs for Processing Common Food Items of Industrial Societies, the Energy Provided by Eating These Same Foods, and the Energy Cost to Produce the Container for These Foods

Processed food Energy input Energy return Container
(kcal/kg) (kcalkg) (kcal)
Instant coffee 18,948 2,645 2,213
Chocolate 18,591 5,104 722
Breakfast cereal
(corn flakes) 15,675 3,877 722
Beet sugar (= 17% sugar
in beets) 5,660 4,000 =400
Cane sugar (= 20% sugar
cane) 3,390 4,000 =400
Fruits and vegetables
(frozen) 1,815 =500 722
Fish (frozen) 1,815 1,058 =400
Baked goods (white bread) 1,485 2,680 559
Meat (hamburger) 1,206 2,714 =400
Ice cream (vamlla) 850 2,015 722
Fruit and vegetables
(canned) 575 =500 2,213
Flour (enriched, sifted) 484 3,643 =400
Milk (3.7% fat) 354 643 2,159

Source: Energy inputs and containers: Harris 1993: energy rccums: Guthrie and Picciano 1995,

appendix K.

both for reasons of food safety and morality-horses are "pets" not "food."

The French, in contrast, insist that horse meat is best for making the dish

steak tartar, prepared from ground meat, which is seasoned with many spices

and flavorings and then eaten uncooked. The French state that horse meat is

safer to eat, as it is "free of parasites" (D. Moerman, personal communication).

The ability to produce virtually unlimited quantities of preferred foods

may be one reason that people in the developed nations, especially the

United States, tend to eat more food than ever before. In addition to the

technological advances in food production, there are social and ideological

reasons for this. In the more developed nations, food production and con-

sumption have become part of the industrial and commercial social structure

of the society (Lieberman 1991). Food is part of "big business," that is, the

economic, social, and political organizations that structure social organization

in industrial nations. Consider the social impact of the fast-food industry,

McDonald's Corporation for example. In the United States, the industriali-

zation of food production has reached the point where far more energy is

expended by the machines that harvest, process, and transport food than is

returned as food calories. Listed in Table 6.6 are several processed foods,

the amount of energy it takes to manufacture them, the energy return from

the foods themselves, and the energy cost of the container used to package

the food. To manufacture one kilogram of instant coffee, for example, re-

quires 18,948 kilocalories. Drinking all that as coffee, slightly more than 529

six-ounce cups, provides 2,645 kilocalories. The metal and plastic container

that the coffee is packaged in costs an additional 2,213 kilocalories to man-

ufacture. Chocolate, breakfast cereals, table sugar from beets and sugar cane,

frozen fruits and vegetables, and frozen fish also require more, or as much,

energy to produce and package than they return as food energy. Hamburg-

ers, the bread they are sandwiched into, and ice cream provide relatively

high-energy returns for the cost to processing these foods. But people do

not live by hamburgers and ice cream alone, not even Americans. Less-

processed foods such as fruits, vegetables, flour and milk provide as much

or more energy than they cost to process. This energy profit is countered

by relatively high-energy losses for manufacturing the container that holds

these foods. The figures in Table 6.6 do not include the costs to grow,

harvest, and transport the foods, nor the cost of operating factories and stores

that process and sell the foods. From an ecological perspective, one that

measures energy flow through a society, industrial food production systems

operate at an energy loss. Every other species of living things would go

extinct under these conditions. People living in industrial societies manage

to survive because industrial and business activities are able to generate

substantial financial profits, which can be used to offset food production

costs. Unfortunately, these financial profits often come at the expense of the

biologically evolved nutrition needs of the people.

Evidence of the commercialism of food and diet is easy to find. Print and

electronic media bombard people with the message that food promotes plea-

sure. This is especially true of foods that are high in fats and sugars. Igor

de Garine (1987) a French anthropologist of food, observes that the style

of food consumption today in wealthy nations reflects more a quest for plea-

sure and increased social status than a desire to fulfill human biological ne-

cessity. Of course this has been the case since at least the time of the ancient

Maya and Romans. Only recently, however, are people able to eat any food,

in any quantity, as often as they wish, if they can afford the price. This may

make people happier, and even more productive, in some sense, but there

are biological consequences of this sensual and socioeconomic pursuit of

satisfaction. There is an increasing body of research that shows that in ad-

dition to the body fat that accumulates from overeating, the chemical by-

products of excessive food digestion are themselves harmful. This includes

the digestion of any food, including low fat foods, sugar-free foods, and any

other so-called health food. These digestive by-products may cause cellular

damage that induces metabolic disease, such as diabetes, and accelerate ag-

ing (Weindruch 1996). Maybe today is a good time to go on that diet that

you have been thinking about.

 

CONCLUSION

This chapter offers one perspective on the evolution of human nutrition.

The mammalian and primate background for human nutrition, the hominid

fossil and archaeological evidence, the behavior and diet of human foraging

societies, and the development of modern foods and life-styles are treated

in some depth. Other aspects of human evolution related to food and nu-

trition are neglected here, such as a detailed discussion of food taboos, ritual

and profane foods, and the effect of seasonal periods of food shortage and

abundance on human biology and behavior. Also neglected are connections

between food production systems and diseases that are not directly caused

by food or diet. Malaria, for example, spread to human populations following

the introduction of agriculture in Africa (Livingstone 1958). Malaria kills

more people. even today, than any other infectious disease, and, conse-

quently, is a potent agent of natural selection and human evolution. There

is evidence that some African societies have developed biocultural systems

to produce and consume food, especially cassava, that reduce the threat of

malaria (Jackson 1990).

The most pressing nutrition problem of the twentieth century is also

barely mentioned. This is the undernutrition and starvation that afflicts

three-fourths of the world's children-nearly two billion people. The toll

that this takes on human health, productivity, and happiness is virtually

unmeasurable. The cause of this suffering lies in the social economic, and

political inequalities between rich and poor; inequalities that the affluent

populations are unwilling to change (Foster 1992; Shields 1995). These and

other topics dealing with human nutritional biochemistry, genetics, history,

ethnology, and psychology relating to food may be pursued in other sources

of reference, including other chapters in this book. The primary message of

all of these sources, and this chapter, is that food is central to human life.

From a biological perspective food is central because of the essential nutri-

ents needed for growth, repair, and maintenance of the body. From a soci-

ocultural perspective food is central because of the behaviors and beliefs

that have evolved around foods and their use. From a medical perspective

food is central because of the consequences of diet and food behavior for

human health. Combining these discrete perspectives into a single holistic

framework leads to the conclusion that human nutrition is a biocultural phe-

nomenon.

 

 

ACKNOWLEDGMENTS

Professor Daniel Moerman and Dr. Alan Goodman read earlier versions of this

chapter, offering many suggestions to improve the presentation and correct errors of

fact. Their help and friendship is much appreciated.

 

 

REFERENCES

Barker. D. J. P., ed. 1992. The Fetal and Infant Origins of Adult Disease. London: British

;Lledical Journal Books.

Binford, L. R. 1984. Fauna/ Remains from the Klasies River Mouth. New York: Academic

Press.

-. 1987. "American Association of Physical Anthropologists Annual Luncheon

Address, April 1986: The Hunting Hypothesis. Archaeological Methods, and

the Past." Yearbook of Physical Anthropology 30: 1-9.

Blakely, M. 1989. "Bone Strontium in Pregnant and Lactating Females from Ar-

chaeological Samples. Amer. Journal of Physical Anthropology 80: 173-85.

Blumenschine, R. J., and J. A. Cavallo. 1992. "Scavenging and Human Evolution."

Scientific American 267( 10):90-96.

Blurton-Jones. N. G. 1993. "The Lives of Hunter-Gather Children: Effects of Pa-

rental Behavior and Parental Reproductive Strategy." In Juvenile Primates.

11. E. Pereira and L. A. Fairbanks, eds. Oxford: Oxford University Press, 309-

26.

Bogin, B. 1985. "The Extinction of Homo Sapiens. *' .Michigan Quarterly Review 24:329-

43.

-. 1988a. "Rural-to-migration." In Biologic Aspects of Human Migration. G. W.

Lasker and C. G. Mascie-Taylor, eds. Cambridge: Cambridge University

Press, 90-l 29.

-. 1988b. Patterns of Human Growth. Cambridge and New York: Cambridge

University Press.

-. 199.5. "Growth and Development: Recent Evolutionary and Biocultural Re-

search." In Biological AnthopoLogy: The State of the Science. IN. T. Boaz and L. D.

Wolfe. eds. Bend, Oregon: International Institute for Human Evolutionary

Research, 49-70.

Brain. C. K. 198 1. The Hunters or the Hunted? A Introduction to African Cave Taphonomy.

Chicago: Chicago University Press.

Bryant, V. M., Jr. 1974. "Prehistoric Diet in Southwest Texas: The Coprolite Evi-

dence." American Antiquity, 39:407-20.

Bryant, V. M., Jr., and G. Williams-Dean. 1975. "The Coprolites of Man." Scientific

American 232100-109.

Burghardt, R. 1990. "The Cultural Context of Diet, Disease and the Body." In Diet

and Disease in Traditional and Developing Societies. G. A. Harrison and J. C. Wa-

terlow, eds. Cambridge: Cambridge University Press, 307-25.

Caraway, C. 1981. The Mayan Design Book. Owing Mills, Maryland: Stemmer House.

Cartmill, M. 1974. "Rethinking Primate Origins." Science 184:436-43.

Cavallo, J. A. 1990. "Cat in the Human Cradle."Natural History (2):52-60.

Chagnon, N. A. 1983. Yanomamo-The Fierce People. New York: Holt, Rinehart and

Winston.

Chandra, R. K. 1990. "McCollum Award Lecture: Nutrition and Immunity: Lessons

from the Past and New Insights into the Future." American Journal of Clinical

Nutrition 53:1087-l 101.

Cohen, M. N. and G. J. Armelagos, eds. 1984. Paleopathologv at the origins of Agnc-

ture. London: Academic Press.

 

 

Eaton. S. B. and M. Konner. 1985. "Paleolithic Nutrition." The New England Journal

of Medicine 312:283-89.

Ezzo. J. A.. C. S., Larsen, and J. H. Burton. 1995. "Elemental Signatures of Human

Diets from the Georgia Bight." American Journal of Physical Anthropology 98:

471-81.

Fellague-Ariouat, J.. and D. J. P. Barker. 1993. "The Diet of Girls and Young Women

at the Beginning of the Century." Nutrition and Health 9: 15-23.

Figueroa. R. 1986. Leyenda de la formacion de los hombres de maiz (summary). In Cocina

Guatemalteca. C. Figueroa, ed. Guatemala City: Editorial Piedra Santa, xiii-

xiv.

Foster. P. 1992. The World Food Problem. Boulder. Colo.: Lynne Rienner.

Fmnken. R. E. 1988. Human Motivation. 2nd ed. Pacific Grove, Calif.: Brooks/Cole.

Garine, I. de. 1994. "The Diet and Nutrition of Human Populations." In T. Ingold

(ed.), Companion Encylopedia qf Anthropogy. London: Routledge, 226-64.

Garn. S. hl., and W. R. Leonard. 1989. "What Did Our Ancestors Eat?" Nutrition

Reviews 47:337-15.

Geertz, C. 1963. Agncultural Involution. Berkeley: University of California Press.

Glander. K. E. 1982. "The Impact of Plant Secondary Compounds on Primate Feed-

ing Behavior." Yearbook of Physical Anthropology 25:1-l8.

Goodall, J. 1983. "Population Dynamics During a 15-Year Period in One Community

of Free-Living Chimpanzees in the Gombe National Park, Tanzania." Ziet-

schrift fur Tierpsychologie 61: l-60.

Goodman, A. H., R. B., Thomas, A. C. Swedlund, and G. J. Armelagos. 1988. "Bio-

cultural Perspectives on Stress in Prehistoric. Historical, and Contemporary

Population Research. Y'earbook of Physical Anthropologv 31:169-202.

Gould, R. A. 1981. "Comparative Ecology of Food-Sharing in Australia and North-

west California." In Omnivorous Primates. R. S. 0. Harding and G. Teleki. eds.

New York: Columbia University Press, -122-54.

Guthrie, H.. and M. F. Picciano. 1995. Human Nutrition. St. Louis Mo: Mosby.

Hladik, C. 1981. "Diet and the Evolution of Feeding Strategies among Forest Pri-

mates." In Omnivorous Primates. R. S. 0. Harding and G. Teleki, eds. New

York: Columbia University Press, 215-54.

Hamilton. E. M. N., E. N. Whitney, and F. S. Sizer, 1988. Nutrition: Concepts and

Controversies. 4th ed. St. Paul, Minn.: West.

Harding, R. S. 0. 1981. "An Order of Omnivores: Nonhuman Primate Diets in the

Wild." In Omnivorous Primates. R. S. 0. Harding and G. Teleki, eds. New

York: Columbia University Press, 191-214.

Harris, M. 1993. Culture, People, Nature. 6th ed. New York: Addison-Wesley

Hayden, B. 1981. "Subsistence and Ecological Adaptations of Modern Hunter/Gath-

erers." In Omnivorous Primates. R. S. 0. Harding and G. Teleki, eds. New

York: Columbia University Press, 34w21.

Hill, K., and A. M. Hurtado. 1989. "Hunter-Gathers of the New World." American

Scentist 77~436-43.

 

Howell, N. 1976. "The Population of the Dobe Area !Kung." In Kalahari Hunter-

Gatheres. R. B. Lee and I. DeVore, eds. Cambridge: Cambridge University

Press, 137-57.

-. 1979. Demography of fhe Dobe !Kung. New York: Academic Press.

Issac, G. L. 1978. "The Food Sharing Behavior of Proto-human Hominids." Scientific

American 238(4):90-108

Jackson, F. L. 1990. "Two Evolutionary Models for the Interactions of Dietary Or-

ganic Cyanogens, Hemoglobins, and Falciparum Malaria. American Journal

of Human Biology 2:521-32.

Katz, S. H., M. L. Heideger, and L. A. Valleroy. 1974. "Traditional Maize Processing

Techniques in the New World. Science 184:765-73.

Kay, R. F. 1985. "Dental Evidence for the Diet of Australopithecus." Annual Review

Of Anthropologv 14:315-41.

Kinzey, IV. G., and M. A. Norconk. 1990. "Hardness as a Basis of Fruit Choice in

Two Sympatric Primates." American Journal of Physical Anthropologv 81:5-15.

Kretchmer, N. 1972. "Lactose and Lactase." In Human Nutrition. IN. Kretchmer and

W. van B. Robertson, eds. San Francisco: W. H. Freeman. 130-38.

Lancaster, J. B. 1985. "Evolutionary Perspectives on Sex Differences in the Higher

Primates." In Gender and tile Life Course. A. S. Rossi. ed. New York: Aldine.

Lancaster, J. B. and C. S. Lancaster. 1983. "Parental Investment: The Hominid Ad-

aptation." In How Humans Adapt. D. J. Ortner, ed. Washington. D.C.: Smith-

sonian Institution Press, 33-65.

Larsen, C. S., and G. R. Milner, eds. 1994. In the Wake of Contact: BiologicalResponses

IO Conquest. New York: Wiley-Liss.

Lee, R. B. 1968. "What Hunters DO for a Living, or, How to Make Out on Scarce

Resources." In Man the Hunter. R. B. Lee and I. DeVore, eds. Cambridge,

Mass.: Harvard University Press.

-. 1984. The Dobe !Kung. New York: Holt, Rinehart and Winston.

Lee R. B., and I. DeVore, eds. 1968. Man the Hunter. Cambridge, Mass.: Harvard

University Press.

Leonard, \V. R. and M. L. Robertson. 1994. "Evolutionary Perspectives on Human

Nutrition: The Influence of Brain and Body Size on Diet and Metabolism."

American Journal of Human Biology 6:77-88.

Lieberman. L. S. 1991. "The Biocultural Consequences of Contemporary and Future

Diets in Developed Countries." Collegium Antropologicum 15:73-85.

Livingstone, F. B. 1958. "Anthropological Implications of Sickle Cell Gene Distri-

bution in West Africa." American Anthropologist 60:533-62.

Martin, R. D. 1983. Human Brain Evolution in an Ecological Context. Fifty-second

James Arthur Lecture. New York: American Museum of Natural History.

McNeill, W. H. 1979. "Historical Patterns of Migration." Current Anthrpology 20:95-

102.

Meindl, R. S. and A. C. Swedlund. 1977. "Secular Trends in Mortality in the Con-

necticut Valley, 1700-1850." Human Biology 49~389-414.

Moerman, D. E. 1986. Medicinal Plants of Native America. 2 vols. Ann Arbor: Univer-

sity of Michigan Museum of Anthropology.

-. 1989. "Poisoned Apples and Honeysuckle: The Medicinal Plants of Native

America." Medical Anthropology Quarterly 3:52-6 1.

Moss, A. J., A. S. Levy, I. Kim, and Y. K Pak. 1989. "Use of Vitamin and Mineral

Supplements in the United States." Advance Data from Vital and Health Sta-

fictiw 17-1. Hyattsville, Md.: National Center for Health Statistics.

Oates, J. F. 1987. "Food Distribution and Foraging Behavior." In Primate Societies.

B. B. Smuts, D. L. Cheney, R. M. Seyfarth, R. W. Wrangham, and T. T.

Struhsaker, eds. Chicago: University of Chicago Press, 197-209.

Paigen, B.. L. R. Goldman, M. M. Magnant, J. H. Highland, and A. T. Steegmann,

Jr. 1987. "Growth of Children Living Near the Hazardous Waste Site, Love

Canal." Human Biology 59:489-508.

Pelto. J. P. and G. H. Pelto. 1983. "Culture, Nutrition, and Health." In The Antho-

pologv of.Medicine. L. Romanucci, D. Moerman, and L. R. Tancredi, eds. New

York: Pmeger, 173-200.

Potts. R. B.. and P. Shipman. 1981. "Cutmarks Made by Stone Tools on Bones from

Olduvai Gorge, Tanzania." *Vature 291:577-80.

Reinhard. K. J. 1989. "Coprolite Evidence of Medicinal Plants." Paper presented at

the annual meetings of the Paleopathology Association, San Diego, California.

Saenz de Tejada, E. 1988. "Analytical Description of the Food Patterns in Mesoam-

erica since Prehistoric Times to the Present, with Special Attention to the

Triada. (Translated from the Spanish title.) Thesis, Universidad de1 Valle de

Guatemala, Guatemala City.

Schaller, G. B. and G. R. Lowther. 1969. "The Relevance of Carnivore Behavior to

the Study of Early Hominids." Southwest Journal of Anthropology 25:307-41.

Schell, L. ,21. 1986. "Community Health Assessment through Physical Anthropology:

Auxological Epidemiology." Human Organization 45:321-27.

-. 1992. "Risk Focusing: An Example of Biocultural Interaction." In Health

and Lifestyle Change: IMASCA Research Papers in Science and Arclraeologv. R. Huss-

Ashmore et al.. eds. 9:137-44.

Schoeninger. M. J. 1979. "Diet and Status at Chalcatzingo: Some Empirical and

Technical Aspects of Strontium Analysis." American Journal of Physical An-

hopology 5 1:295-3 10.

-. 1982. "Diet and the Evolution of Modern Human Form in the Middle

East. American Journal of Physical Anthropology 58:37-52

Scrimshaw, S. S. and V. R. Young. 1976. "The Requirements of Human Nutrition."

In Human Nutrition. N. Kretchmer and W. van B. Robertson, eds. San Fran-

cisco: W.H. Freeman, 156-70.

Shields. D. L. L. 1995. The Color of Hunger. London: Rowman and Littlefield.

Short, R. V. 1976. "The Evolution of Human Reproduction." Proceedings, Royal So-

ciety, Series B 195:3-24.

Sillen A., and M. Kavanagh. 1982. "Strontium and Paleodietary Research: A Re-

view." Yearbook of Physical Anthopology 25:67-90.

Steegmann, A. T., Jr. 1985. "18th Century British Military Stature: Growth Cessa-

tion, Selective Recruiting, Secular Trends, Nutrition at Birth, Cold and Oc-

cupation." Human Biology 57~77-95.

Strum, S. 1981. "Processes and Products of Change: Baboon Predatory Behavior at

Gilgil, Kenya." In Omnivorvus Primates. R. S. 0. Harding and G. Teleki, eds.

New York: Columbia University Press, 255-302.

Sussman R. W. and W. G. Kinzey. 1984. "The Ecological Role of the Callitrichidae:

A Review." American Journal of Physical Anthropology 64:419-49.

Swedlund, A C., R. S. Meindl, and M. I. Gradie. 1980. "Family Reconstitution in

the Connecticut Valley: Progress on Record Linkage and the Mortality Sur-

vey." In Genealogical Demographv. B. Dyke and W. Morril, eds. Sew York:

Academic Press, 139-55.

Takasaki, Y., C. Pieddeloup, and S. Anzai. 1957. "Trends in Food Intake and Di-

gestive Cancer Mortalities in Japan." Human Biology 59:951-57.

Tedlock, D.. trans. 1985. Pop01 v&. New York: Simon and Schuster.

Teleki, G. 1981. "The Omnivorous Diet and Eclectic Feeding Habits of Chimpan-

zees in Gombe National Park, Tanzania." In Omnivorous Primates. R. S. 0.

Harding and G. Teleki, eds. New York: Columbia University Press, 303-43.

Teleki, G.. E. Hunt, and J. H. Pfifferling. 1976. "Demographic Observations (1963-

1973) on the Chimpanzees of the Gombe rational Park. Tanzania." Journal

of Human Evolution 5:559-98.

Tobias. P. 1'. 1975. Anthropometry among Disadvantaged Peoples: Studies in

Southern Africa." In Biosocial Interrelations IN Populational Adaptation E. S

Watts. F. E. Johnston, and G. N. Lasker. eds. The Hague:: Mouton. 287-305.

Ubelaker, D. H.. hl. A. Katzenberg. and L. G. Doyon. 1995. "Status and Diet in

Precontact Highland Ecuador. American Journal of Physical .4nhopology 97:

-103-l 1.

Van Gerven, D. P.. S. G. Sheridan. and W.Y. Adams 1995. "The Health and Nutri-

tion of a Medieval Nubian Population. American Anthropologist 97~168-80.

Walker. A. C. 1981. "Dietary Hypotheses and Human Evolution." Philosophical

Transactions of the Roval Society, B 292:58-64.

Walker, A.. AI. R. Zimmerman. and R. E. F. Leakey. 1982. "A Possible Case of Hy-

pervitaminosis A in Homo Erectus. Nature 296248-50.

Washburn. S. L., and C. H. Lancaster. 1968. "The Evolution of Hunting." In Man

the Hunter. R. B. Lee and I. DeVore, eds. Cambridge, Mass.: Harvard Uni-

versity Press.

Washburn. S. L. and R. Moore. 1980. Ape into Human: A Study of Human Evolution.

2nd ed. Boston: Little, Brown.

Weindruch, R. 1996. "Caloric Restriction and Aging." Scientific American 274 (1):46-

52.

Weiss. K. hi., R. E. Ferrell. and C. L. Hams. 1984. "A New World Syndrome of

Metabolic Diseases with a Genetic and Evolutionary Basis." Yearbook of Phys-

ical Anthropology 2 7: 153-78.

Whyte, R. 0. 1974. Rural Nutrition in MONSOON Asia. Kuala Lumpur: Oxford University

Press.

Yip, R.. and K. Scanlon. 1995. "The Changing Body &lass of US Children: Impli-

cations for Future Growth Curves." Paper presented at the Human Growth

Workshop, Emory University, October 2-3.

Zihlman, A. L. 1981. "Women as Shapers of Human Adaptation." In W'ornen the Gath-

erer. F. Dahlberg, ed. New Haven, Conn.: Yale University Press, 75-120.