11 Archaic Homo
Beth Shook; Ph.D.; Lara Braff; Katie Nelson; Kelsie Aguilera; and M.A.
Amanda Wolcott Paskey, M.A., Cosumnes River College
AnnMarie Beasley Cisneros, M.A., American River College
This chapter is a revision from “Chapter 11: Archaic Homo” by Amanda Wolcott Paskey and AnnMarie Beasley Cisneros. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Learning Objectives
- Identify the main groupings of Archaic Homo sapiens, such as Neanderthals.
- Explain how shifting environmental conditions required flexibility of adaptations, both anatomically and behaviorally.
- Describe the unique anatomical and cultural characteristics of Archaic Homo sapiens, including Neanderthals, in contrast to other hominins.
- Articulate how Middle Pleistocene hominin fossils fit into evolutionary trends including cranial capacity (brain size) development, cultural innovations, and migration patterns.
- Identify the shared traits, regional variations, and local adaptations among Archaic Homo sapiens.
- Detail the increased complexity and debates surrounding the classification of hominins in light of transitional species, species admixture, etc.
Breaking the Stigma of the “Caveman”
What do you think of when you hear the word “caveman”? Perhaps you imagine a character from a film such as The Croods, Tarzan, and Encino Man or from the cartoon The Flintstones. Maybe you picture the tennis-playing, therapy-going hairy Neanderthals from Geico Insurance commercials. Or perhaps you imagine characters from The Far Side or B.C. comics. Whichever you picture, the character in your mind is likely stooped over with a heavy brow, tangled long locks and other body hair, and clothed in animal skins, if anything. They might be holding a club with a confused look on their face, standing at the entrance to a cave or dragging an animal carcass to a fire for their next meal (see Figure 11.1). You might have even signed up to take this course because of what you knew—or expected to learn—about “cavemen.”
These images have long been the stigma and expectation about our ancestors at the transition to modern Homo sapiens. Tracing back to works as early as Carl Linnaeus, scientists once propagated and advanced this imagery, creating a clear picture in the minds of early scholars that informed the general public, even through today, that Archaic Homo sapiens, “cavemen,” were somehow fundamentally different and much less intelligent than we are now. Unfortunately, this view is overly simplistic, misleading, and incorrect. Understanding what Archaic Homo sapiens were actually like requires a much more complex and nuanced picture, one that comes into sharper focus as continuing research uncovers more about the lives of our not-too-distant (and not-too-different) ancestors.
The first characterizations of Archaic Homo sapiens were formed from limited fossil evidence in a time when ethnocentric and species-centric perspectives (anthropocentrism) were more widely accepted and entrenched in both society and science. Today, scientists are working from a more complete fossil record from three continents (Africa, Asia, and Europe), and genetic evidence informs their analyses and conclusions. The existence of Archaic Homo sapiens marks an exciting point in our lineage—a point at which many modern traits had emerged and key refinements were on the horizon. Anatomically, humans today are not that much different from Archaic Homo sapiens.
This chapter will examine how the environment with which Archaic Homo sapiens had to contend shaped their biological and cultural evolution. It will also examine the key anatomical traits that define this group of fossils, focusing on the most well-known of them, including Neanderthals. The chapter will describe cultural innovations that aided their adaptation to the changing environment, as well as their geographic distribution and regional variations. Additionally, exciting new research is examined that suggests even greater nuance and complexity during this time period.
The Changing Environment
While modern climate change is of critical concern today due to its cause (human activity) and pace (unprecedentedly rapid), the existence of global climate change itself is not a recent phenomenon. The focus of this chapter, the Middle Pleistocene (roughly between 780 kya and 125 kya), is the time period in which Archaic Homo sapiens appears in the fossil record—a time that witnessed some of the most drastic climatic changes in human existence. During this time period, there were 15 major and 50 minor glacial events in Europe, alone.
What exactly is glaciation? When scientists talk about glacial events, they are referring to the climate being in an ice age. This means that the ocean levels were much lower than today, because much of the earth’s water was tied up in large glaciers or ice sheets. Additionally, the average temperature would have been much cooler, which would have better supported an Arctic or tundra-adapted plant-and-animal ecosystem in northern latitudes. The most interesting and relevant features of Middle Pleistocene glacial events are the sheer number of them and their repeated bouts: this era alternated between glacial periods and warmer periods, known as interglacials. In other words, the planet wasn’t in an ice age the whole time.
You can see the dramatic and increasing fluctuations in temperature, recorded through foraminifera, in Figure 11.2. The distance between highs and lows demonstrates the severity of temperature shifts. Much as the Richter scale represents more intense earthquakes with more dramatic peaks, so too does this chart, which uses dramatic peaks to demonstrate intense temperature swings.
Glacial periods are defined by Earth’s average temperature being lower. Worldwide, temperatures are reduced, with cold areas becoming even colder. Huge portions of the landscape may have become inaccessible during glacial events due to the formation of glaciers and massive ice sheets. In Europe, the Scandinavian continental glacier covered what is today Ireland, England, Sweden, Norway, Denmark, and some of continental Europe. Plant and animal communities shifted to lower latitudes along the periphery of ice sheets. Additionally, some new land was opened during glacials. Evaporation with little runoff reduced sea levels by as much as almost 150 meters, shifting coastlines outward by in some instances as much as almost 100 kilometers. Additionally, land became exposed that connected what were previously unconnected continents such as Africa into Yemen at the Gulf of Aden.
Glacial periods also affected equatorial regions and other regions that are today thought of as warmer or at least more temperate parts of the globe, including Africa. While these areas were not covered with glaciers, increased global glaciation resulted in lower sea levels and expanded coastlines. Cooler temperatures were accompanied by the drying of the climate, which caused significantly reduced rainfall, increased aridity, and the expansion of deserts. It is an interesting question to consider whether the same plants and animals that lived in these regions prior to the ice ages would be able to survive and thrive in this new climate. Plant and animal communities shifted in response to the changing climate, whenever possible.
Surviving During the Middle Pleistocene
Rather than a single selective force, the Middle Pleistocene was marked by periods of fluctuation, not just cold periods. Interglacials interrupted glaciations, reversing trends in sea level, coastline, temperature, precipitation, and aridity, as well as glacier size and location. Interglacials are marked by increased rainfall and a higher temperature, which causes built-up ice in glaciers to melt. This leads to glacial retreat, which is the shrinking of glaciers and the movement of the glaciers back toward the poles, as we’ve seen in our lifetime. During interglacials, sea levels increase, flooding some previously exposed coastlines and continental connections. In addition, plant and animal communities shift accordingly, often finding more temperate climates to the north and less arid and more humid climates in the tropics (Van Andel and Tzedakis 1996).
Scientists have found that the Olorgesailie region in southern Kenya was at various times in the Middle Pleistocene a deep lake, a drought-dried lakebed with an area criss-crossed by small streams, and a grassland. While various animal species would have moved in and out of the area as the climate shifted, some animal species went extinct, and new, often related, species took up residence. The trend, scientists noted, was that animals with more specialized features went extinct and animals with more generalized features, such as animals we see today, survived in this changing climatic time period. For example, a zebra with specialized teeth for eating grass was ultimately replaced by a zebra that could eat both grass and other types of vegetation. If this small, localized example shows such a dramatic change in terms of the environment and the plant and animal biocommunities, what would have been the impact on humans?
There is no way humans could have escaped the effects of Middle Pleistocene climate change, no matter what region of the world they were living in. As noted earlier, and as evidenced by what was seen in the other biotic communities, humans would have faced changing food sources as previous sources of food may have gone extinct or moved to a different latitude. Depending on where they were living, fresh water may have been limited. Durial glacials, lower sea levels would have given humans more land to live on, while the interglacials would have reduced the available land through the increase in rainfall and associated sea level rise. Dry land connections between the continents would have made movement from one continent to another by foot easier at times than today, although these passageways were not consistently available through the Middle Pleistocene due to the glacial/interglacial cycle. Finally, as evidenced by the Olorgesailie region in Kenya, during the Middle Pleistocene animal species that were overly specialized to one particular type of environment were less likely to survive when compared to their more generalized counterparts. Evidence suggests that this same pattern may have held true for Archaic Homo sapiens, in terms of their ability to survive this dramatic period of climate change.
Defining Characteristics of Archaic Homo sapiens
Archaic Homo sapiens share our species name but are distinguished by the term “Archaic” as a way of recognizing both the long period of time between their appearance and ours, as well as the way in which human traits have continued to evolve over time—making Archaic Homo sapiens look slightly different from us today, despite being considered the same species. Living throughout Africa, and the Middle East during the Middle Pleistocene, Archaic Homo sapiens are considered, in many ways, transitional between Homo erectus and modern Homo sapiens (see Figure 11.3). Archaic Homo sapiens share the defining trait of an increased brain size of at least 1,100 cc and averaging 1,200 cc, although there are significant regional and temporal (time) variations. Because of these variations, scientists disagree on whether these fossils represent a single, variable species or multiple, closely related species (sometimes called Homo antecessor, Homo bodoensis, Homo heidelbergensis, Homo georgicus, Homo neanderthalensis, and Homo rhodesiensis).
An active area of scholarship in the discipline involves reconciling the diversity of species from this time period and establishing the phylogenetic relationships between them. The term “Archaic Homo sapiens” can mean different things to different scholars within the discipline. The intent of this chapter is to provide an understanding of the diversity of this time period and provide data used to make interpretations from among the most likely possibilities. Although we recognize that some anthropologists split many of these fossils into separate species, until the issue is resolved at the discipline level, this chapter will rely on the widely used naming conventions that refer to many fossils from this time period as Archaic Homo sapiens. We will discuss these contemporaneous fossils as a unit, with the exception of a particularly unique population living in Europe and West Asia known as the Neanderthals, which we will examine separately.
Trait |
Homo erectus |
Archaic Homo sapiens (including Neaderthals) |
Anatomically Modern Homo sapiens |
Time |
1.8 mya–200,000 ya |
600,000–40,000 ya |
315,000 ya–today |
Brain size |
900 cc |
1,200 cc (1,500 cc when including Neanderthals) |
1,400 cc |
Skull Shape |
Long and low, angular |
Intermediate |
Short and high, globular |
Forehead |
Absent |
Emerging |
Present |
Nasal Region |
Projecting nasal bones (bridge of the nose), no midfacial prognathism |
Wider nasal aperture and some midfacial prognathism, particularly pronounced among Neanderthals |
Narrower nasal aperture, no midfacial prognathism |
Chin |
Absent |
Absent |
Present |
Other Facial Features |
Large brow ridge and large projecting face |
Intermediate |
Small brow ridge and retracted face |
Other Skull Features |
Nuchal torus, sagittal keel, thick cranial bone |
Projecting occipital bone, often called occipital bun in Neanderthals; intermediate thickness of cranial bone |
Small bump on rear of skull, if anything; thin cranial bone |
Dentition |
Large teeth, especially front teeth |
Slightly smaller teeth; front teeth still large; retromolar gap in Neanderthals |
Smaller teeth |
Postcranial Features |
Robust bones of skeleton |
Robust bones of skeleton |
More gracile bones of skeleton |
When comparing Homo erectus, Archaic Homo sapiens, and anatomically modern Homo sapiens, one can see that Archaic Homo sapiens are intermediate in their physical form. For some features, this follows the trends first seen in Homo erectus with other features having early, less developed forms of traits seen in modern Homo sapiens. For example, Archaic Homo sapiens trended toward less angular and higher skulls than Homo erectus. However, the archaic skulls were not as short and globular and had less developed foreheads compared to anatomically modern Homo sapiens. Archaic Homo sapiens had smaller brow ridges and a less-projecting face than Homo erectus and slightly smaller teeth, although incisors and canines were often about as large as those of Homo erectus. Archaic Homo sapiens also had a wider nasal aperture, or opening for the nose, and a forward-projecting midfacial region, which is later seen more fully developed among Neanderthals and is known as midfacial prognathism. The occipital bone often projected and the cranial bone was of intermediate thickness, somewhat reduced from Homo erectus but not nearly as thin as that of anatomically modern Homo sapiens. The postcrania remained fairly robust. Identifying a set of features that is unique to Archaic Homo sapiens is a challenging task, due to both individual and geographic variation—these developments were not all present to the same degree in all individuals. Neanderthals are the exception, as they had several unique traits that made them notably different from modern Homo sapiens as well as their closely related Archaic cousins.
The one thing that is clear about Archaic Homo sapiens is that, despite general features, there is a lot of regional variation, which is first seen in the different Homo erectus specimens across Asia and Africa. While the general features of Archaic Homo sapiens, identified earlier, are present in the fossils of this time period, there are significant regional differences. The majority of this regional variation lies in the degree to which fossils have features more closely aligned with Homo erectus or with anatomically modern Homo sapiens.
To illustrate this point, we will examine three exemplary specimens, one from each of the three continents on which Archaic Homo sapiens lived. First, in Africa, a specimen from Broken Hill is one of several individuals found in the Kabwe lead mine in Zambia. It had a large brain (1,300 cc) and taller cranium as well as many Homo erectus-like skull features, including massive brow ridges, a large face, and thick cranial bones (Figure 11.4). Second, one partial crania from Dali, China, is representative of Archaic Homo sapiens in Asia, with large and robust features with heavy brow ridges, akin to what is seen in Homo erectus, and a large cranial capacity intermediate between Homo erectus and anatomically modern Homo sapiens (Figure 11.5). Third, an almost-complete skeleton was found in northern Spain at Atapuerca. Atapuerca 5 (Figure 11.6) has thick cranial bone, an enlarged cranial capacity, intermediate cranial height, and a more rounded cranium than seen previously. Additionally, Atapuerca 5 demonstrates features that foreshadow Neanderthals, including increased midfacial prognathism. After examining some of the fossils such as those from Kabwe, Dali, and Atapuerca, the transitional nature of Archaic Homo sapiens is clear: their features place them squarely between Homo erectus and modern Homo sapiens.
Due to the transitional nature of Archaic Homo sapiens, identifying the time period with which they are associated is problematic and complex. Generally, it is agreed that Archaic Homo sapiens lived between 600,000 and 200,000 years ago, with regional variation and overlap between Homo erectus on the early end of the spectrum and modern Homo sapiens and Neanderthals on the latter end. The earliest-known Archaic Homo sapiens fossils tentatively date to about 600,000 years ago in Africa, to around 300,000 years ago in Asia, and to about 350,000 years ago in Europe (and potentially as early as 600,000 years ago). Determining the end point of Archaic Homo sapiens is also problematic since it largely depends upon when the next subspecies of Homo sapiens appears and the classification of highly intermediate specimens. For example, in Africa, the end of Archaic Homo sapiens is met with the appearance of modern Homo sapiens, while in Europe it is the appearance of Neanderthals that is traditionally seen as marking the transition from other Archaic Homo sapiens.
It is important to remember that this time period is represented by many branching relationships and assuming an evolutionary trajectory that follows a single linear path would not be correct. Even still, Archaic Homo sapiens mark an important chapter in the human lineage, connecting more ancestral forms, such as Homo erectus, to modern Homo sapiens. During this period of climatic transition and fluctuation, Archaic Homo sapiens mirror the challenges of their environments. Showing increasing regional variation due to the need for local adaptation, there is no single archetype for this group; the defining characteristic seems to be variability.
Neanderthals
One well-known population of Archaic Homo sapiens are the Neanderthals, named after the site where they were first discovered in the Neander Valley, or “thal” in German, located near Dusseldorf, Germany. Popularly known as the stereotypical “cavemen” examined at the outset of this chapter, recent research is upending long-held beliefs about this group of Archaics. Neanderthal behavior was increasingly complex, far beyond what was exhibited by even other Archaic Homo sapiens discussed throughout this chapter. We implore you to forget the image of the iconic caveman and have an open mind when exploring the fossil evidence of the Neanderthals.
It is important to understand why Neanderthals are separated from other Archaic Homo sapiens. Unlike the rest of Archaic Homo sapiens, Neanderthals are easily defined and identified in many ways. Evidence suggests the time period when Neanderthals lived was between 150,000 and 40,000 years ago. There is a clear geographic boundary of where Neanderthals lived: western Europe, the Middle East, and western Asia. No Neanderthal fossils have ever been discovered outside of this area, including Africa. This is a bit curious, as other Archaics seem to have adapted in Africa and then migrated elsewhere, but Neanderthals’ regional association makes sense in light of the environment to which they were best adapted: namely, extreme cold weather. Additionally, Neanderthals have a unique and distinct cluster of physical characteristics. While a few aspects of Neanderthals are shared among some Archaic Homo sapiens, such as the types of tools, most Neanderthal anatomical and behavioral attributes are unique to them.
Neanderthals lived during some of the coldest times during the last Ice Age and at far northern latitudes. This means Neanderthals were living very close to the glacial edge, rather than in a more temperate region of the globe like some of their Archaic Homo sapiens relatives. While able to survive in arctic conditions, most Neanderthal sites dating to the glacial periods were found farther away from the severe cold, in a steppe tundra-like environment, which would have been more hospitable to Neanderthals, and their food sources, both flora and fauna (Ashton 2002; Nicholson 2017; Richter 2006). Their range likely expanded and contracted along with European glacial events, moving into the Middle East during glacial events when Europe became even cooler, and when the animals they hunted would have moved for the same reason. During interglacials, when Europe warmed a bit, Neanderthals and their prey would have been able to move back into Western Europe. Clearly, the true hallmark of Neanderthals is their adaptation to an unstable environment, shifting between warm and cold, as the climate was in constant flux throughout their existence (Adler et al. 2003; Boettger et al. 2009).
Many of the Neanderthals’ defining physical features are more extreme and robust versions of traits seen in other Archaic Homo sapiens, clustered in this single population. Brain size, namely an enlargement of the cranial capacity, is one such trait. The average Neanderthal brain size is around 1,500 cc, and the range for Neanderthal brains can extend to upwards of 1,700 cc. The majority of the increase in the brain occurs in the occipital region, or the back part of the brain, resulting in a skull that has a large cranial capacity with a distinctly long and low shape that is slightly wider than previous forms at the far back of the skull. Modern humans have a brain size comparable to that of Neanderthals; however, our brain expansion occurred in the frontal region of the brain, not the back, as in Neanderthal brains. This difference is also the main reason why Neanderthals lack the vertical forehead that modern humans possess. They simply did not need an enlarged forehead, because their brain expansion occurred in the rear of their brain. Due to cranial expansion, the back of the Neanderthal skull is less angular (as compared to Homo erectus), but not as rounded as Homo sapiens, producing an elongated shape, akin to a football.
Another feature that continues the trend noted in previous hominins is the enlargement of the nasal region, or the nose. Neanderthal noses are large and have a wide nasal aperture, which is the opening for the nose. While the nose is only made up of two bones, the nasals, the true size of the nose can be determined by looking at other facial features, including the nasal aperture, and the angle of the nasal and maxillary, or facial bones. In Neanderthals, these indicate a large, forward-projecting nose that appears to be pulled forward away from the rest of the face. This feature is further emphasized by the backward-sloping nature of the cheekbones, or the zygomatic arches. The unique shape and size of the Neanderthal nose is often characterized by the term midfacial prognathism—a jutting out of the middle portion of the face, or nose. This is in sharp contrast to the prognathism exhibited by other hominins, who exhibited prognathism, or the jutting out, of their jaws.
The teeth of the Neanderthals follow a similar pattern seen in the Archaic Homo sapiens, which is an overall reduction in size, especially as compared to the extremely large teeth seen in the genus Australopithecus. However, while the teeth continued to reduce, the jaw size did not keep pace, leaving Neanderthals with an oversized jaw for their teeth, and a gap between their final molar and the end of their jaw. This gap is called a retromolar gap.
The projecting occipital bone present in other Archaic Homo sapiens is also more prominent in Neanderthals, extending the trend found in Archaics. Among Neanderthals, this projection of bone is easily identified by its bun shape on the back of the skull and is known as an occipital bun. This projection appears quite similar to a dinner roll in size and shape. Its purpose, if any, remains unknown.
Continuing the Archaic Homo sapiens trend, Neanderthal brow ridges are prominent but somewhat smaller in size than those of Homo erectus and earlier Archaic Homo sapiens. In Neanderthals, the brow ridges are also often slightly less arched than those of other Archaic Homo sapiens.
In addition to extending traits present in Archaic Homo sapiens, Neanderthals possess several distinct traits. Neanderthal infraorbital foramina, the holes in the maxillae or cheek bones through which blood vessels pass, are notably enlarged compared to other hominins. The Neanderthal postcrania are also unique in that they demonstrate increased robusticity in terms of the thickness of bones and body proportions that show a barrel-shaped chest and short, stocky limbs, as well as increased musculature. These body portions are seen across the spectrum of Neanderthals—in men, women, and children.
Traditionally, many of the unique traits that Neanderthals possess were seen as adaptations to the extreme cold, dry environments in which they often lived and which exerted strong selective forces. For example, Bergmann’s and Allen’s Rules dictate that an increased body mass and short, stocky limbs are common in animals that live in cold conditions. Neanderthals were said to have matched the predictions of Bergmann’s and Allen’s Rules perfectly (Churchill 2006). In addition, the Neanderthal skull also exhibits adaptations to the cold. Neanderthals’ large infraorbital foramina allow for larger blood vessels, increasing the volume of blood that is found closest to the skin, which helps to keep the skin warmer. Their enlarged noses resulted in longer nasal passages and mucus membranes that warmed and moistened cold air before it reached the lungs. The Neanderthals’ larger nose has long been thought to have acted as a humidifier, easing physical exertion in their climate, although research on this particular trait continues to be studied and debated (Rae et al. 2011).
New research, however, seems to suggest that these unique skeletal adaptations might not have been strict adaptations to cold weather (Evteev et al. 2017; Pearce et al. 2013). For example, large brow ridges might have served as a way to shade the face from the sun. The increased occipital portion of the brain, some researchers state, was to support a larger visual system present in Neanderthals. This visual system would have given them increased light sensitivity, which would have been useful in higher latitudes that had dark winters. And, while recent modeling of nostril airflow on reconstructed Neanderthal specimens supports the notion that Neanderthals had extensive mucus membranes inside their noses, the data shows that modern Homo sapiens are superior to Neanderthals in our ability to use our noses as a way to warm and cool air. However, Neanderthals were able to snort air through their noses better than we can. Why is this important? When combined with the fact that Neanderthals tended to prefer a more temperate, tundra-like environment, and that other physical traits suggest that Neanderthals had huge bodies that needed massive amounts of calories to sustain them, the picture gets clearer. Massive amounts of energy would have been required to power a Neanderthal body, and anything that might have made them more calorically efficient would have been favored. Efficient breathing, through larger noses into large lungs, meaning deeper breaths, would have been favored. To further save energy expenditure, body sizes might have been sacrificed as well. These same types of adaptations are similar to ones seen in children today who are born in high altitudes, not cold climates. Figure 11.7 provides a summary of these unique features of the Neanderthal.
Distinct Neanderthal Anatomical Features |
|
Brain Size |
1,500 cc average |
Skull Shape |
Long and low |
Brow Ridge Size |
Large |
Nose Size |
Large, with midfacial prognathism |
Dentition |
Reduced, but large jaw size, creating retromolar gap |
Occipital Region |
Enlarged occipital region, occipital bun |
Other Unique Cranial Features |
Large infraorbital foramina |
Postcranial Features |
Short and stocky body, increased musculature, barrel-shaped chest |
A classic example of a Neanderthal with all of the characteristics mentioned above is the nearly complete La Ferrassie 1 Neanderthal, from France. This is a male skeleton, with a brain size of around 1640cc, an extremely large nose and infraorbital foramina, brow ridges that are marked in size, and an overall robust skeleton (Figure 11.8).
Neanderthal Culture: Tool Making and Use
One key Neanderthal adaptation was their cultural innovations, which are an important way that hominins adapt to their environment. As you recall, Homo erectus‘s tools, Acheulean handaxes, represented an increase in complexity over Oldowan tools, allowing more efficient removal of meat and possibly calculated scavenging. In contrast, Neanderthal tools mark a significant innovation in tool-making technique and use with Mousterian tools (named after the Le Moustier site in southwest France). These tools were significantly smaller, thinner, and lighter than Acheulean handaxes and formed a true toolkit. The materials used for Mousterian tools were of higher quality, which allowed for both more precise toolmaking and tool reworking when the tools broke or dulled after frequent reuse. The use of higher-quality materials is also indicative of required forethought and planning to acquire them for tool manufacture. It is noteworthy that the Neanderthals, unlike Homo erectus, saved and reused their tools, rather than making new ones each time a tool was needed.
Mousterian tools are constructed in a very unique manner, utilizing the Levallois technique (Figure 11.9), named after the first finds of tools made with this technique, which were discovered in the Levallois-Perret suburb of Paris, France. The Levallois technique is a multistep process that requires preparing the core, or raw material, in a specific way that will yield flakes that are roughly uniform in dimension. The flakes are then turned into individual tools. The preparation of the core is akin to peeling a potato or carrot with a vegetable peeler—when peeling vegetables, you want to remove the skin in long, regular strokes, so that you are taking off the same amount of the vegetable all the way around. In the same way, the Levallois technique requires removing all edges of the cortex, or outside surface of the raw material, in a circle before removing the lid. The flakes, which will eventually be turned into the individual tools, can then be removed from the core. The potential yield of tools from one core would be many, as seen in Figure 11.10, compared to all previous tool-making processes, in which one core yielded a single tool. This manufacturing process might be considered the ultimate zero-waste tool-making technique (Delpiano et al. 2018).
Neanderthal tools were used for a variety of purposes, including cutting, butchering, woodworking or antler working, and hide working. Additionally, because the Mousterian tools were lighter than previous stone tools, Neanderthals could haft, or attach the tool onto a handle, as the stone would not have been too heavy (Degano et al. 2019). Neanderthals attached small stone blades onto short wood or antler handles to make knives or other small weapons, as well as attached larger blades onto longer shafts to make spears. New research examining tar-covered stones and black lumps at several Neanderthal sites in Europe suggests that Neanderthals may have been making tar by distilling it from birch tree bark, which could have been used to glue the stone tool onto its handle. If Neanderthals were, in fact, manufacturing tar to act as glue, this would predate modern humans in Africa using tree resin or similar adhesives by nearly 100,000 years.
Evidence shows that raw materials used by Neanderthals came from distances as far away as 100 km. This could indicate a variety of things regarding Neanderthal behavior, including a limited trade network with other Neanderthal groups or simply a large area scoured by Neanderthals when collecting raw materials. While research on specific applications continues, it should be clear from this brief discussion that Neanderthal tool manufacturing was much more complex than previous tool-making efforts, requiring technical expertise, patience, and skills beyond toolmaking to carry out.
Neanderthal Culture: Hunting and Diet
With their more sophisticated suite of tools and robust muscular bodies, Neanderthals were better armed for hunting than previous hominins. The animal remains in Neanderthal sites show that, unlike earlier Archaic Homo sapiens, Neanderthals were very effective hunters who were able to kill their own prey, rather than relying on scavenging. Though more refined than the tools of earlier hominins, the Neanderthal spear was not the kind of weapon that would have been thrown; rather, it would have been used in a jabbing fashion (Churchill 1998; Kortlandt 2002). This may have required Neanderthals to hunt in groups rather than individually and made it necessary to approach their prey quite closely (Gaudzinski-Windheuser et al. 2018). Remember, the animals living with Neanderthals were very large-bodied due to their adaptations to cold weather; this would have included species of deer, horses, and bovids (relatives of the cow).
Isotopes from Neanderthal bones show that meat was a significant component of their diet, similar to that seen in carnivores like wolves (Bocherens et al. 1999; Jaouen et al. 2019; Richards et al. 2000). In addition to large prey, their diet included ibex, seals, rabbits, and pigeons. Though red meat was a critical component of the Neanderthal diet, evidence shows that at times they also ate limpets, mussels, and pine nuts. Tartar examined from Neanderthal teeth in Iraq and Belgium reveal that they also ate plant material including wheat, barley, date palms, and tubers, first cooking them to make them palatable (Henry et al. 2010). While Neanderthals’ diet varied according to the specific environment in which they lived, meat comprised up to 80% of their diet (Wiẞin et al. 2015).
Neanderthal Culture: Caring for the Injured and Sick
While the close-range style of hunting used by Neanderthals was effective, it also had some major consequences. Many Neanderthal skeletons have been found with significant injuries, which could have caused paralysis or severely limited their mobility. Many of the injuries are to the head, neck, or upper body. Thomas Berger and Erik Trinkaus (1995) conducted a statistical comparative analysis of Neanderthal injuries compared to those recorded in modern-day workers’ compensation reports and found that the closest match was between Neanderthal injuries and those of rodeo workers. Rodeo professionals have a high rate of head and neck injuries that are similar to the Neanderthals’ injuries. What do Neanderthals and rodeo workers have in common? They were both getting very close to large, strong animals, and at times their encounters went awry.
The extensive injuries sustained by Neanderthals are evident in many fossil remains. Shanidar 1 (Figure 11.11), an adult male found at the Shanidar site in northern Iraq and dating to 45,000 ya, has a lifetime of injuries recorded in his bones (Stewart 1977). Shanidar 1 sustained—and healed from—an injury to the face that would have likely caused blindness. His lower right arm was missing and the right humerus shows severe atrophy, likely due to disuse. This pattern has been interpreted to indicate a substantial injury that required or otherwise resulted in amputation or wasting away of the lower arm. Additionally, Shanidar 1 suffered from bony growths in the inner ear that would have significantly impaired his hearing and severe arthritis in the feet. He also exhibited extensive anterior tooth wear, matching the pattern of wear found among modern populations who use their teeth as a tool. Rather than an anomaly, the type of injuries evident in Shanidar 1 are similar to those found in many other Neanderthal fossils, revealing injuries likely sustained from hunting large mammals as well as demonstrating a long life of physical activity.
The pattern of injuries is as significant as the fact that Shanidar 1 and other injured Neanderthals show evidence of having survived their severe injuries. One of the earliest-known Neanderthal discoveries—the one on whom misinformed analysis shaped the stereotype of the species for nearly a century—is the La Chapelle-aux-Saints Neanderthal (Trinkaus 1985). The La Chapelle Neanderthal had a damaged eye orbit that likely caused blindness and suffered arthritis of the spine (Dawson and Trinkaus 1997). He had also lost most of his teeth, many of which he had lived without for so long that the mandibular and maxillary bones were partially reabsorbed due to lack of use. The La Chapelle Neanderthal was also thought to be at least in his mid-forties at death, an old age for the rough life of the Late Pleistocene—giving rise to his nickname, “the Old Man.” To have survived so long with so many injuries that obviously precluded successful large game hunting, he must have been taken care of by others. Such caretaking behavior is also evident in the survival of other seriously injured Neanderthals, such as Shanidar 1. Long thought to be a hallmark modern human characteristic, taking care of the injured and elderly, for example preparing or pre-chewing food for those without teeth, indicates strong social ties among Neanderthals.
Neanderthal Culture: Ritual Life
Such care practices may have been expressed upon death as well. Nearly complete Neanderthal skeletons are not uncommon in the fossil record, and most are well preserved within apparently deliberate burials that involve deep graves and bodies found in specific, often fetal or flexed positions (Harrold 1980). Discoveries of pollen in a grave at the Shanidar site in the 1960s led scientists to think that perhaps Neanderthals had placed flowering plants in the grave, an indication of ritual ceremony or spirituality so common in modern humans. But more recent investigations have raised some doubt about this conclusion (Pomeroy et al. 2020). The pollen may have been brought in by burrowing rodents. Claims of grave goods or other ornamentation in burials are similarly debated, although possible.
Some tantalizing evidence for symbolism, and debatably, ritual, is the frequent occurrence of natural pigments, such as ochre (red) and manganese dioxide (black) in Neanderthal sites that could have been used for art. However, the actual uses of pigments are unclear, as there is very little evidence of art or paintings in Mousterian sites. One exception may be the recent discovery in Spain of a perforated shell that appears to be painted with an orange pigment, which may be evidence of Neanderthal art and jewelry. However, many pigments also have properties that make them good emulsifiers in adhesive (like for attaching a stone tool to a wooden handle) or useful in tanning hides. So the presence of pigment may or may not be associated with symbolic thought; however, it definitely does show a technological sophistication beyond that of earlier Archaic hominins.
The Lasting Gift of Neanderthals: Tantalizing New Directions for Research
Examining the more recent time period in which Neanderthals lived and the extensive excavations completed across Europe allows for a much more complete archaeological record from this time period. Additionally, the increased cultural complexity such as complex tools and ritual behaviors expressed by Neanderthals left a more detailed record than previous hominins. Intentional burials enhanced preservation of the dead and potentially associated ritual behaviors. Such evidence allows for a more complete and nuanced picture of this species.
Additional analyses are possible on many Neanderthal finds, due to increased preservation of bone, the amount of specimens that have been uncovered, and the recency in which Neanderthals lived. We should be cautious, however, to consider the potential bias of many Neanderthal sites. Overwhelmingly, Neanderthal skeletons are found complete, with injuries or evidence of disease in caves. Does this mean all Neanderthals lived a tough, disease-wrought life? Probably not. It does, however, indicate that the sick were cared for by others, and that they lived in environments that preserved their bodies incredibly well. These additional studies include the examination of dental calculus and even DNA analysis. While limited, samples of Neanderthal DNA have been successfully extracted and analyzed. Research thus far has identified specific genetic markers that show some Neanderthals were light-skinned and probably red-haired with light eyes. Genetic analyses, different from the typical hominin reconstruction done with earlier species, allow scientists to further investigate soft tissue markers of Neanderthals and other more recent hominin species. These studies offer striking conclusions regarding Neanderthal traits, their physical appearance, and their culture, as reflected in these artists’ reconstructions (Figure 11.12).
Dr. Svante Pääbo (Figure 11.13), of the Max Planck Institute for Evolutionary Anthropology, has been at the forefront of much of this new research, largely in the form of genomic studies (The Nobel Prize 2022). Awarded the Nobel Prize for Physiology or Medicine in 2022, Pääbo is known primarily for his work with ancient DNA. He has successfully sequenced mitochondrial DNA (mtDNA) as well as the entire Neanderthal genome from nuclear DNA. His genomic work has led to the realization that Denisovans are genetically distinct from Neanderthals, as well as the recent identification of a Neanderthal father and teenage daughter, which he discovered by looking for unique DNA markers in the fossil record. Additionally, Pääbo’s genomic work has provided researchers with additional lines of evidence regarding the connections between hominin fossils (such as Neanderthals) and modern people, their time of divergence, and current genetic overlap. The work of Pääbo has even formalized a new field of study within anthropology—paleogenomics. To stay up to date with Dr. Svante Pääbo’s work, be sure to follow his lab’s website.
Neanderthal Culture: Communicating through Speech
To successfully live in groups and to foster cultural innovations, Neanderthals would have required at least a basic form of communication in order to function, possibly using a speech-based communication system. The challenge with this line of research is that speech, of course, is not preserved, so indirect evidence must be used to support this conclusion. It is thought that Neanderthals would have possessed some basic speech, as evidenced from a variety of sources, including throat anatomy and genetic evidence (Lieberman 1971). There is only one bone in the human body that could demonstrate if a hominin was able to speak, or produce clear vocalizations like modern humans, and that is the hyoid, a U-shaped bone that is found in the throat and is associated with the ability to precisely control the vocal cords. Very few hyoid bones have been found in the archaeological record; however, a few have been uncovered in Neanderthal burials. The shape of the Neanderthal hyoid is nearly identical to that of modern humans, pointing to the likelihood that they had the same vocal capabilities as modern humans. In addition, geneticists have uncovered a mutation present in both modern humans and Neanderthals—the FOXP2 gene—that is possibly linked to the ability to speak. However, other scientists argue that we cannot make sweeping conclusions that the FOXP2 gene accounts for speech due to small sample size. Finally, scientists have also pointed to the increasingly complex cultural behavior of Neanderthals as a sign that symbolic communication, likely through speech, would have been the only way to pass down the skills needed to make, for example, a Levallois blade or to position a body for intentional burial.
Neanderthal Intelligence
One of the enduring questions about Neanderthals centers on their intelligence, specifically in comparison to modern humans. Brain volume indicates that Neanderthals certainly had a large brain, but it continues to be debated if Neanderthals were of equal intelligence to modern humans. Remember, creatures with larger body sizes tend to have larger brains; however, scaling of the brain is not always associated with greater intelligence (Alex 2018). Brain volume (cranial capacity), cultural complexity, tool use, and compassion toward their kind all point to an increase in intellect among Neanderthals when compared to previous hominins.
Yet, new research is suggesting additional differences between Neanderthal brains and our own. For example, Euluned Pearce and colleagues (2013), from the University of Oxford, noted the frontal lobes of Neanderthals and modern humans are almost identical. However, Neanderthals had a larger visual cortex—the portion of the brain involved in processing visual information. This would have left Neanderthals with less brain tissue for other functions, including those that would have aided them in dealing with large social groupings, one of the differences that has been suggested to exist between Neanderthals and modern humans. Other differences were found when geneticist John Blangero, from the Texas Biomedical Research Institute, compared data from the Neanderthal genome against data from modern study participants. Blangero and his colleagues (Blangero et al. 2014) discovered that some Neanderthal brain components were very different, and smaller, than those in the modern sample. Differences were found in areas associated with the processing of information and controlling emotion and motivation, as well as overall brain connectivity. In short, as Blangero stated, “Neanderthals were certainly cognitively adept,” although their specific abilities may have differed from modern humans’ in key areas (qtd. in Wong 2015). This point has been echoed in other recent genetic studies comparing Neanderthal and anatomically modern human brains (el-Showk 2019).
Finally, scientists are fairly certain that Neanderthal brain development after birth was not the same as that of modern humans. After birth, anatomically modern Homo sapiens babies go through a critical period of brain expansion and cognitive development. It appears that Neanderthal babies’ brains and bodies did not follow the same developmental pattern (Smith et al. 2010; Zollikofer and Ponce de León 2013). Modern humans enjoy an extended period of childhood, which allows children to engage in imaginative play and develop creativity that fosters cognitive skills. Neanderthals had a more limited childhood, with less development of the creative mind that may have affected their species’ success (Nowell 2016).
The exact nature of Neanderthal intelligence remains under investigation, however. Some studies disagree with the idea that Neanderthal intelligence had limitations compared to our own, noting the extensive evidence of Neanderthals having limb asymmetry. Their tools also have wear marks indicating that they were hand-dominant. This is further supported by marks on Neanderthal teeth that demonstrate hand dominance. The Neanderthal “stuff-and-cut method” of eating, noted by David Frayer and colleagues (Frayer et al. 2012), would have seen Neanderthals hold a piece of meat in their teeth, while pulling it taut with one hand, and then using the other hand, their dominant one, to cut the meat off of the larger slab being held in their teeth. When looking at 17 Neanderthals and their tooth wear, only two do not show markings associated with a right-hand dominant individual eating in this manner. Further, it has been established that favoring the right hand is a key marker between modern humans and chimpanzees, and that handedness also relates to language development, in the form of bilateral brain development. That Neanderthals likely were hand-dominant suggests they had an indicator of bilateral brain development and a precondition for human speech.
The Middle Stone Age: Neanderthal Contemporaries in Africa
While Neanderthals made their home on and adapted to the European and Asian continents, evidence of fossil humans in Africa show they were also adapting to their local environments. These populations in Africa exhibit many more similarities to modern humans than Neanderthals, as well as overall evolutionary success. While the African fossil sample size is smaller and more fragmentary than the number of Neanderthal specimens across Europe and Asia, the African sample is interesting in that it represents a longer time period and larger geographical area. This group of fossils—often represented by the name “Middle Stone Age,” or MSA—dates to between 300,000 and 30,000 years ago across the entire continent of Africa. As with Archaic Homo sapiens, there is much variability seen in this African set of fossils. There are also a few key consistent elements: none of them exhibit Neanderthal skeletal features; instead, they demonstrate features that are increasingly consistent with anatomically modern Homo sapiens.
Similarities to Neanderthals and MSA contemporaries in Africa are seen, however, in their behavioral adaptations, including stone tools and other cultural elements. The tools associated with the specimens living in Africa during this time period are, like their physical features, varied. In some parts of Africa, namely Northern Africa, stone tools from this time so closely resemble Neanderthal tools that they are classified as Mousterian. In sub-Saharan Africa, the stone tools associated with these specimens are labeled as MSA. Some scholars argue that these could also be a type of Mousterian tools, but they are still typically subdivided based on geographical location.
Recall that Mousterian tools were much more advanced than their Acheulean predecessors in terms of how the stone tools were manufactured, the quality of the stones used, and the ultimate use of the tools that were made. In addition, recent evidence suggests that MSA tools may also have been heat treated—to improve the quality of the stone tool produced (Stolarczyk and Schmidt 2018). Evidence for heat treating is seen not only through advanced analysis of the tool itself but also through the residue of fires from this time period. Fire residues show a shift over time from small, short fires fueled by grasses (probably intended for cooking) to larger, more intensive fires that required the exploitation of dry wood, exactly the type of fire that would have been needed for heat treating stone tools (Esteban et al. 2018).
Other cultural elements seen with MSA specimens include the use of marine (sea-based) resources for their diet (Parkington 2003), manufacture of bone tools, use of adhesive and compound tools (e.g., hafted tools), shell bead production, engraving, use of pigments (such as ochre), and other more advanced tool-making technology (e.g., microlithics). While many of these cultural elements are also seen to a limited extent among Neanderthals, developments at MSA sites appear more complex. This MSA cultural expansion may have been a response to climate change or an increased use of language, complex communication, and/or symbolic thought. Others have suggested that the MSA cultural expansion was due to the increase of marine resources in their diet, which included more fatty acids that may have aided their cognitive development. Still others have suggested that the increased cultural complexity was due to increased interaction among groups, which spurred competition to innovate. Recent studies suggest that perhaps the best explanation for the marked cultural complexity of MSA cultural artifacts is best explained by the simple fact that they lived in diverse habitats (Kandel et al. 2015). This would have necessitated a unique set of cultural adaptations for each habitat type (for example, specialized marine tools would have been needed along coastal sites but not at inland locations). Simply put, the most useful adaptation of MSA was their flexibility of behavior and adaptability to their local environment. As noted previously in this chapter, flexibility of behavior and physical traits, rather than specialization, seems to be a feature that was favored in hominin evolution at this time.
Where Did They Go? The End of Neanderthals
While MSA specimens were increasingly successful and ultimately transitioned into modern Homo sapiens, Neanderthals disappeared from the fossil record by around 40,000 years ago. What happened to them? We know, based on genetics, that modern humans come largely from the modern people who occupied Africa around 300,000 to 100,000 years ago, at the same time that Neanderthals were living in northern Europe and Asia. As you will learn in Chapter 12, modern humans expanded out of Africa around 150,000 years ago, rapidly entering areas of Europe and Asia inhabited by Neanderthals and other Archaic hominins. Despite intense interest and speculation in fictional works about possible interactions between these two groups, there is very little direct evidence of either peaceful coexistence or aggressive encounters. It is clear, though, that these two closely related hominins shared Europe for thousands of years, and recent DNA evidence suggests that they occasionally interbred (Fu et al. 2015). Geneticists have found traces of Neanderthal DNA (as much as 1% to 4%) in modern humans of European and Asian descent not present in modern humans from sub-Saharan Africa. This is indicative of limited regional interbreeding with Neanderthals.
While some interbreeding likely occurred, as a whole, Neanderthals did not survive. What is the cause for their extinction? This question has fascinated many researchers and several possibilities (theories) have been suggested, including:
- At the time that Neanderthals were disappearing from the fossil record, the climate went through both cooling and warming periods—each of which posed challenges for Neanderthal survival (Defleur and Desclaux 2019; Staubwasser et al. 2018). It has been argued that as temperatures warmed, large-bodied animals, well adapted to cold weather, moved farther north to find colder environments or faced extinction. A shifting resource base could have been problematic for continued Neanderthal existence, especially as additional humans, in the form of modern Homo sapiens, began to appear in Europe and compete for a smaller pool of available resources.
- It has been suggested that the eruption of a European volcano 40,000 years ago could have put a strain on available plant resources (Golovanova et al. 2010). The eruption would have greatly affected local microclimates, reducing the overall temperature enough to alter the growing season.
- Possible differences in cognitive development may have limited Neanderthals in terms of their creative problem solving. As much as they were biologically specialized for their environment, the nature of their intelligence might not have offered them the creative problem-solving skills to innovate ways to adapt their culture when faced with a changing environment (Pearce et al. 2013).
- CRISPR gene-editing technology has been used in studies to evaluate potential differences between human and Neanderthal brains, based on differences in the genetic code. Potential differences include a Neanderthal propensity for mutations related to brain development that could account for more rapid brain development, maturation, synapse misfires, and less-orderly neural processes (Mora-Bermúdez et al. 2022; Trujillo et al. 2021). Fundamental differences in brain function at the cellular level may account for the differential survival rates of Neanderthal and modern human populations.
- There is evidence that suggests reproduction may have posed challenges for Neanderthals. Childbirth was thought to have been at least as difficult for female Neanderthals as anatomically modern Homo sapiens (Weaver and Hublin 2009). Female Neanderthals may have become sexually mature at an older age, even older than modern humans. This delayed maturation could have kept the Neanderthal population size small. A recent study has further suggested that male Neanderthals might have had a genetic marker on the Y chromosome that could have caused incompatibility between the fetus and mother during gestation; this would have had severe consequences for birth rate and survival (Mendez et al. 2016). Even a small but continuous decrease in fertility would have been enough to result in the extinction of Neanderthals (Degioanni et al. 2019).
- As mentioned above, the end of Neanderthal existence overlaps with modern human expansion into northern Europe and Asia. There is no conclusive direct evidence to indicate that Neanderthals and modern humans lived peacefully side by side, nor that they engaged in warfare, but by studying modern societies and the tendencies of modern humans, it has been suggested that modern humans may not have warmly embraced their close but slightly odd-looking cousins when they first encountered them (Churchill et al. 2009). Nevertheless, direct competition with modern humans for the same resources may have contributed to the Neanderthals’ decline (Gilpin et al. 2016); it may also have exposed them to new diseases, brought by modern humans (Houldcroft and Underdown 2016), which further decimated their population. Estimates of energy expenditures suggest Neanderthals had slightly higher caloric needs than modern humans (Venner 2018). When competing for similar resources, the slightly greater efficiency of modern humans might have helped them experience greater success in the face of competition—at a cost to Neanderthals.
As Neanderthal populations were fairly small to begin with (estimated between 5,000 and 70,000 individuals; Bocquet-Appel and Degioanni 2013), one or a combination of these factors could have easily led to their demise. As more research is conducted, we will likely get a better picture of exactly what led to Neanderthal extinction.
Denisovans
While Neanderthals represent one regionally adapted branch of the Archaic Homo sapiens family tree, recent discoveries in Siberia and the Tibetan Plateau surprised paleoanthropologists by revealing yet another population that was contemporary with Archaic Homo sapiens, Neanderthals, and modern Homo sapiens. The genetic analysis of a child’s finger bone (Figure 11.14) and an adult upper third molar (Figure 11.15) from Denisova Cave in the Altai Mountains in Siberia by a team including Svante Pääbo discovered that the mitochondrial and nuclear DNA sequences reflected distinct genetic differences from all known Archaic populations. Dubbed “Denisovans” after the cave in which the bones were found, this population is more closely related to Neanderthals than modern humans, suggesting the two groups shared an ancestor who split from modern humans first, then the Neanderthal-Denisovan line diverged more recently (Reich et al. 2010).
Denisovans share up to 5% of their DNA with modern Melanesians, aboriginal Australians, and Polynesians, and 0.2% of their DNA with other modern Asian populations and Native Americans. Additional studies have suggested one (Vernot et al. 2018) or two (Browning et al. 2018) separate points of time when interbreeding occurred between modern humans and Denisovans.
Genetic analysis reveals that Denisovans (potentially three distinct populations) had adaptations for life at high altitudes that prevented them from developing altitude sickness and hypoxia in extreme environments such as Tibet, where the average annual temperature is close to 0℃ and the altitude is more than a kilometer (about 4,000 feet) above sea level. Through protein analysis of a jawbone, one study (Chen et al. 2019) has placed Denisovans in Tibet as early as 160,000 years ago. Genetic evidence of interbreeding has linked modern Tibetan populations with Denisovans 30,000 to 40,000 years ago, which implies that the unique high-altitude adaptations seen in modern Tibetans may have originated with Denisovans (Huerta-Sanchez et al. 2014).
Other research suggests tantalizing new directions regarding Denisovans. Stone tools similar to those found in Siberia have been uncovered in the Tibetan plateau suggesting a connection between the Denisovan populations in those two areas (Zhang et al. 2018). The molar of a young girl, possibly Denisovan, has been found in Laos and shows strong similarities to specimens from China (Demeter et al. 2022). And DNA sequencing from discoveries in the Denisova Cave have yielded a genome that has been interpreted as the first-generation offspring of a Denisovan father and Neanderthal mother (Slon et al. 2018). While this research is not yet conclusive and is still being interpreted, exciting new possibilities are being revealed. To stay up-to-date with new discoveries, consider following organizations such as the Smithsonian’s Human Origins Program on social media.
How Do These Fit In? Homo naledi and Homo floresiensis
Recently, some fossils have been unearthed that have challenged our understanding of the hominin lineage. The fossils of Homo naledi and Homo floresiensis are significant for several reasons but are mostly known for how they don’t fit the previously held patterns of hominin evolution. While we examine information about these species, we ask you to consider the evidence presented in this chapter and others to draw your own conclusions regarding the significance and placement of these two unusual fossil species in the hominin lineage.
Homo naledi
In 2013 recreational spelunkers uncovered a collection of bones deep in a cave network in Johannesburg, South Africa. The cave system, known as Rising Star, had been well documented by other cavers; however, it appears few people had ever gone as far into the cave as these spelunkers did. Lee Berger, paleoanthropologist at University of Witwatersrand, in Johannesburg, immediately put out a call for what he termed “underground astronauts” to begin recovery and excavation of the fossil materials. Unlike other excavations, Berger and most other paleoanthropologists would not be able to access the elusive site, as it was incredibly difficult to reach, and at some points there was only eight inches of space through which to navigate. The underground astronauts, all petite, slender female anthropologists, were the only ones who were able to access this remarkable site. Armed with small excavation tools and a video camera, which streamed the footage up to the rest of the team at the surface, the team worked together and uncovered a total of 1,550 bones, representing at least 15 individuals, as seen in Figure 11.16. Later, an additional 131 bones, including an almost-complete cranium, were found in a nearby chamber of the cave, representing three more individuals (Figure 11.17). Berger called in a team of specialists to participate in what was dubbed “Paleoanthropology Summer Camp.” Each researcher specialized in a different portion of the hominin skeleton. With various specialists working simultaneously, more rapid analysis was possible of Homo naledi than most fossil discoveries.
While access to the site, approximately 80 m from any known cave entrance or opening, was treacherous for researchers, it must have been difficult for Homo naledi as well. The route included moving through a portion that is just 25 cm wide at some points, known as “Superman’s Crawl.” The only way to get through this section is by crawling on your stomach with one arm by your side and the other raised above your head. Past Superman’s Crawl, a jagged wall known as the Dragon’s Back would have been very difficult to traverse. Below that, a narrow vertical chute would have eventually led down to the area where the fossils were discovered. While geology changes over time and the cave system likely has undergone its fair share, it is not likely that these features arose after Homo naledi lived (Dirks et al. 2017). This has made scientists curious as to how the bones ended up in the bottom of the cave system in the first place. It has been suggested that Homo naledi deposited the bones there, one way or another. If Homo naledi did deposit the bones, either through random disposal or intentional burial, this raises questions regarding their symbolic behavior and other cultural traits, including the use of fire, to access a very dark cave system. Another competing idea is that a few individuals may have entered the cave system to escape a predator and then got stuck. To account for the sheer number of fossils, this would have had to happen multiple times.
The features of Homo naledi are well-documented due to the fairly large sample, which represents individuals of all sexes and a wide range of ages. The skull shape and features are very much like other members of the genus Homo—including a sagittal keel and large brow, like Homo erectus, and a well-developed frontal lobe, similar to modern humans—yet the brain size is significantly smaller than its counterparts, at approximately 500 cc (560 cc for males and 465 cc for females). The teeth also exhibit features of later members of the genus Homo, such as Neanderthals, including a reduction in overall tooth size. Homo naledi also had unique shoulder anatomy and curved fingers, indicating similarities to tree-dwelling primates, which is very different from any other hominin yet found. Perhaps the greatest shock of all is that Homo naledi has been dated to 335,000 to 236,000 years ago, placing it as a contemporary to modern Homo sapiens, despite its very primitive features. An additional specimen of a child, found in 2021, not only shares many of the unique features found in the adult specimen but will also add insight into the growth and development of individuals of this species (Brophy et al. 2021).
Homo floresiensis
In a small cave called Liang Bua, on the island of Flores, in Indonesia, a small collection of fossils were discovered beginning in 2003 (Figure 11.18). The fossil fragments represent as many as nine individuals, including a nearly complete female skeleton. The features of the skull are very similar to that of Homo erectus, including the presence of a sagittal keel, an arching brow ridges and nuchal torus, and the lack of a chin (Figure 11.19). Homo floresiensis, as the new species is called, had a brain size that was remarkably small at 400 cc, and recent genetic studies suggest a common ancestor with modern humans that predates Homo erectus.
The complete female skeleton, who was an adult, was approximately a meter tall and would have weighed just under 30 kg, which is significantly shorter and just a few kilograms more than the average, modern, young elementary-aged child. A reconstructed comparison between an anatomically modern human and Homo floresiensis can be seen in Figure 11.20. The small size of the fossil has earned the species the nickname “the Hobbit.” Many questions have been asked about the stature of this species, as all of the specimens found also show evidence of diminutive stature and small brain size. Some explanations include pathology; however, this seems unlikely as all fossils found thus far demonstrate the same pattern. Another possible explanation lies in a biological phenomena seen in other animal species also found on the island, which date to a similar time period. This phenomenon, called insular dwarfing, is due to limited food resources on an island, which can create a selective pressure for large-bodied species to be selected for smaller size, as an island would not have been able to support their larger-bodied cousins for a long period of time. This phenomenon is the cause of other unique species known to have lived on the island at the same time, including the miniature stegodon, a dwarf elephant species.
There is ongoing research and debate regarding Homo floresiensis’ dates of existence, with some researchers concluding that they lived on Flores until perhaps as recently as 17,000 years ago, although they are more often dated to 100,000 to 60,000 years ago. Stone tools from that time period uncovered at the site are similar to other hominin stone tools found on the island of Flores. Homo floresiensis would have hunted a wide range of animals, including the miniature stegodon, giant rats, and other large rodents. Other animals on the island that could have threatened them include the giant komodo dragon. An interesting note about this island chain is that ancestors of Homo floresiensis would have had to traverse the open ocean in order to get there, as the nearest island is almost 10 km away, and there is little evidence to support that a land bridge connecting mainland Asia or Australia to the island would have been present. This separation from the mainland would also have limited the number of other animals, including predators and human species, that would have had access to the island. Anatomically modern Homo sapiens arrived on the island around 30,000 years ago and, if some researchers’ later dates for Homo floresiensis are correct, both species may have lived on Flores at the same time. The modern population living on the island of Flores today believes that their ancestors came from the Liang Bua cave; however, recent genetic studies have determined they are not related to Homo floresiensis (Tucci et al. 2018).
Conclusion
Research presented in the chapter contributes to why scientists have taken to nicknaming this time period “the muddle in the middle.” We know that the Middle Pleistocene picks up from Homo erectus and ends with the appearance of anatomically modern Homo sapiens. While the start and the end are clear, it’s the middle that is messy. As more research is conducted and more data is collected, rather than clarifying our understanding of the hominin lineage during this time period, it only inspires more questions, particularly about the relationships between hominins during this time period, including the oft-misunderstood Neanderthal. Research is painting a more detailed picture of Neanderthal intelligence and both biological and behavioral adaptations. At the same time, their relationship to other Middle Pleistocene hominins, including Denisovans, as well as modern humans, remains unclear.
Homo naledi and Homo floresiensis are clear outliers when compared to their contemporary hominin species. Each has surprised paleoanthropologists for both their archaic traits in relatively modern times and their unique combination of traits seen in archaic species and modern humans. While these finds have been exciting, they have also completely upended the assumed trajectory of the human lineage, causing scientists to re-examine assumptions about hominin evolution and what it means to be modern. Add this to the developments being made using ancient DNA, other new fossil discoveries, and other innovations in paleoanthropology, and you see that our understanding of Archaic Homo sapiens and others living during this time period is rapidly developing and changing. This is a true testament to the nature of science and the scientific method.
Clearly, hominins of the Middle Pleistocene are distinct from our species today. Yet, understanding the hominins that directly preceded our species and clarifying the evolutionary relationships between us is important to better understanding our own place in nature.
Hominin Species Summaries
Hominin |
Archaic Homo sapiens |
Dates |
600,000–200,000 years ago (although some regional variation) |
Region(s) |
Africa, Europe, and Asia |
Famous discoveries |
Broken Hill (Zambia), Atapuerca (Spain) |
Brain size |
1,200 cc average |
Dentition |
Slightly smaller teeth in back of mouth, larger front teeth |
Cranial features |
Emerging forehead, no chin, projecting occipital region |
Postcranial features |
Robust skeleton |
Culture |
Varied regionally, but some continue to use Acheulean handaxe, others adopt Mousterian tool culture |
Other |
Lots of regional variation in this species |
Species |
Homo naledi |
Dates |
335,000–236,000 years ago |
Region(s) |
South Africa |
Famous discoveries |
Rising Star Cave |
Brain size |
500 cc average |
Dentition |
Reduced tooth size |
Cranial features |
Sagittal keel, large brow, well-developed frontal region |
Postcranial features |
Suspensory shoulder |
Culture |
unknown |
Other |
N/A |
Hominin |
Neanderthals |
Dates |
150,000–40,000 years ago |
Region(s) |
Western Europe, Middle East, and Western Asia only |
Famous discoveries |
Shanidar (Iraq), La Chapelle-aux-Saints (France) |
Brain size |
1500 cc average |
Dentition |
Retromolar gap |
Cranial features |
Large brow ridge, midfacial prognathism, large infraorbital foramina, occipital bun |
Postcranial features |
Robust skeleton with short and stocky body, increased musculature, barrel chest |
Culture |
Mousterian tools often constructed using the Levallois technique |
Other |
N/A |
Species |
Homo floresiensis |
Dates |
100,000–60,000 years ago, perhaps as recently as 17,000 years ago |
Region(s) |
Liang Bua, island of Flores, Indonesia |
Famous discoveries |
“The Hobbit” |
Brain size |
400 cc average |
Dentition |
unknown |
Cranial features |
Sagittal keel, arching brow ridges, nuchal torus, no chin |
Postcranial features |
Very short stature (approximately 3.5 ft.) |
Culture |
Tools similar to other tools found on the island of Flores |
Other |
N/A |
Hominin |
Denisovans |
Dates |
100,000–30,000 years ago |
Region(s) |
Siberia |
Famous discoveries |
Child’s finger bone and adult molar |
Brain size |
unknown |
Dentition |
Large molars (from limited evidence) |
Cranial features |
unknown |
Postcranial features |
unknown |
Culture |
unknown |
Other |
Closely related to Neanderthals (genetically) |
Review Questions
- What physical and cultural features are unique to Archaic Homo sapiens? How are Archaic Homo sapiens different in both physical and cultural characteristics from Homo erectus?
- Describe the specific changes to the brain and skull first seen in Archaic Homo sapiens. Why does the shape of the skull change so dramatically from Homo erectus?
- What role did the shifting environment play in the adaptation of Archaic Homo sapiens, including Neanderthals? Discuss at least one physical feature and one cultural feature that would have assisted these groups in surviving the changing environment.
- What does the regional variation in Archaic Homo sapiens represent in terms of the broader story of our species’ evolution?
- Describe the issues raised by the discoveries of Homo naledi and Homo floresiensis in the understanding of the story of the evolution of Homo sapiens.
Key Terms
Allele: Each of two or more alternative forms of a gene that arise by mutation and are found at the same place on a chromosome.
Anthropocentrism: A way of thinking that assumes humans are the most important species and leads to interpreting the world always through a human lens. Species-centric science and thought.
Cortex: The outside, or rough outer covering, of a rock. Usually the cortex is removed during the process of stone tool creation.
Ethnocentric: Applying negative judgments to other cultures based on comparison to one’s own.
Exogenous DNA: DNA that originates from sources outside of the specimen you are trying to sequence.
Flexed position: Fetal position, in which the legs are drawn up to the middle of the body and the arms are drawn toward the body center. Intentional burials are often found in the flexed body position.
Foraminifera: Microscopic single-celled organisms with a shell that are common in all marine environments. The fossil record of foraminifera extends back well over 500 million years.
Glaciation: A glacial period, or time when a large portion of the world is covered by glaciers and ice sheets.
Globular: Round-shaped, like a globe.
Grave goods: Items included with a body at burial. Items may signify occupation or hobbies, social status, or level of importance in the community, or they may be items believed necessary for the afterlife.
Haft: A handle. Also used as a verb—to attach a handle to an item, such as a stone tool.
Infraorbital foramina: Small holes on the maxilla bone of the face that allows nerves and blood to reach the skin.
Insular dwarfing: A form of dwarfism that occurs when a limited geographic region, such as an island, causes a large-bodied animal to be selected for a smaller body size.
Interglacial: A warmer period between two glacial time periods.
Levallois technique: A distinctive technique of stone tool manufacturing used by Archaic Homo sapiens, including Neanderthals. The technique involves the preparation of a core and striking edges off in a regular fashion around the core. Then a series of similarly sized pieces can be removed, which can then be turned into different tools.
Midfacial prognathism: A forward projection of the nose or the middle facial region. Usually associated with Neanderthals.
Mousterian tools: The stone tool industry of Neanderthals and their contemporaries in Africa and Western Asia. Mousterian tools are known for a diverse set of flake tools, which is different from the large bifacial tools of the Acheulean industry.
Nasal aperture: The opening for the nose visible on a skull. Often pear- or heart-shaped.
Occipital bun: A prominent bulge or projection on the back of the skull, specifically the occipital bone. This is a feature present only on Neanderthal skulls.
Ochre: A natural clay pigment mixed with ferric oxide and clay and sand. Ranges in color from brown to red to orange.
Retracted face: A face that is flatter.
Retromolar gap: A space behind the last molar and the end of the jaw. This is a feature present only on Neanderthals. It also occurs through cultural modification in modern humans who have had their third molars, or wisdom teeth, removed.
About the Authors
Amanda Wolcott Paskey, M.A.
Cosumnes River College, paskeya@crc.losrios.edu
Amanda Wolcott Paskey is an anthropology professor at Cosumnes River College in Sacramento, California. She earned her B.A. and M.A. in anthropology from the University of California, Davis. Her speciality in anthropology is archaeology; however, she was trained in a holistic program and most of her teaching load is in biological anthropology. She is currently working on analyzing a post–gold rush era archaeological site, in Sacramento, with colleagues and students. This project has given her many opportunities to engage in sharing archaeology with a public audience, including local school children and Sacramentans interested in local history.
AnnMarie Beasley Cisneros, M.A.
American River College, beaslea@arc.losrios.edu
AnnMarie Beasley Cisneros is an anthropology professor at American River College in Sacramento, California. Trained as a four-field anthropologist, she earned her B.A. and M.A. in anthropology from California State University, Sacramento. She regularly teaches biological anthropology, among other courses, and is currently engaged in applied anthropology work in community development with historically underserved communities. She most recently has particularly enjoyed facilitating her students’ involvement in projects serving Sacramento’s Latino and immigrant Mexican populations.
For Further Exploration
Anne and Bernard Spitzer Hall of Human Origins—American Museum of Natural History.
“Dawn of Humanity,” PBS documentary, 2015
“DNA Clues to Our Inner Neanderthal,” TED Talk by Svante Pääbo, 2011.
“The Dirt” Podcast, Episode 30, “The Human Family Tree (Shrub? Crabgrass? Tumbleweed?), Part 3: Very Humany Indeed”.
Frank, Rebecca. 2021. “The Genus Homo.” In Explorations: Lab and Activity Manual, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff. Arlington, VA: American Anthropological Association.
Hobbits on Flores, Indonesia – Smithsonian Human Origins.
Lumping or Splitting in the Fossil Record – UC Berkeley Understanding Evolution.
Neandertals and More – Max Planck Institute for Evolutionary Anthropology.
Neanderthals: Body of Evidence – SAPIENS.
Perash, Rose L., and Kristen A. Broehl. 2021. “Hominin Review: Evolutionary Trends.” In Explorations: Lab and Activity Manual, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff. Arlington, VA: American Anthropological Association.
Perkl, Bradley. “Brain, Language, Lithics.” In Explorations: Lab and Activity Manual, edited by. Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff. CC BY-NC. Arlington, VA: American Anthropological Association.
Shanidar 3 – Neanderthal Skeleton – Smithsonian Human Origins.
Species – Smithsonian Human Origins.
Smithsonian Human Origins Program Facebook page (@smithsonian.humanorigins).
Paleoartist Brings Human Evolution to Life – Elisabeth Daynés.
References
Adler, Daniel S., Timothy J. Prindiville, and Nicholas J. Conard. 2003. “Patterns of Spatial Organization and Land Use During the Eemian Interglacial in the Rhineland: New Data from Wallertheim, Germany.” Eurasian Prehistory 1(2): 25–78.
Alex, Bridget. 2018. “Neanderthal Brains: Bigger, Not Necessarily Better.” Discover, September 21, 2018. https://www.discovermagazine.com/planet-earth/neanderthal-brains-bigger-not-necessarily-better.
Ashton, Nick M. 2002. “Absence of Humans in Britain during the Last Interglacial Period (Oxygen Isotope Stage 5e).” Publications du CERP 8: 93–103.
Berger, Lee R., John Hawks, Darryl J. de Ruiter, Steven E. Churchill, Peter Schmid, Lucas K. Delezene, Tracy L. Kivell, et al. 2015. “Homo naledi, a New Species of the Genus Homo from the Dinaledi Chamber, South Africa.” eLife 4:e09560. https://doi.org/10.7554/eLife.09560.
Berger, Thomas D., and Erik Trinkaus. 1995. “Patterns of Trauma among the Neanderthals.” Journal of Archaeological Science 22 (6): 841–852.
Blangero, J., E.E. Quillen, M.A. Almeida, D.R. McKay, J.M. Peralta, S. Williams-Blangero, J.E. Curran, R. Duggirala, D.C. Glahn. 2014. “Genomic Reconstruction of Neandertal Brain Structure and Function.” Paper presented at the American Association of Physical Anthropologists Annual Meeting, Calgary, Alberta, Canada, 12 April 2014.
Bocherens, Hervé, Daniel Billiou, André Mariotti, Marylène Patou-Mathis, Marcel Otte, Dominique Bonjean, and Michel Toussaint. 1999. “Palaeoenvionmental and Palaeodietary Implications of Isotopic Biogeochemistry of Last Interglacial Neanderthal and Mammal Bones in Scladina Cave (Belgium).” Journal of Archaeological Science 26 (6): 599–607.
Bocquet-Appel, Jean-Pierre and Anna Degioanni. 2013. “Neanderthal Demographic Estimates.” Current Anthropology 54 (S8): S202–S213.
Boettger, Tatjana, Elena Yu. Novenko, Andrej A. Velichko, Olga K. Borisova, Konstantin V. Kremenetski, Stefan Knetsch, and Frank W. Junge. 2009. “Instability of Climate and Vegetation Dynamics in Central and Eastern Europe during the Final Stage of the Last Interglacial (Eemian, Mikulino) and Early Glaciation.” Quaternary International 207 (1-2): 137–144. https://doi.org/10.1016/j.quaint.2009.05.006.
Brophy, Juliet K., Marina C. Elliott, Darryl J. De Ruiter, Debra R. Bolter, Steven E. Churchill, Christopher S. Walker, John Hawks, and Lee R. Berger. 2021. “Immature Hominin Craniodental Remains from a New Locality in the Rising Star Cave System, South Africa.” PaleoAnthropology 1: 1–14. https://doi.org/10.48738/2021.iss1.64.
Browning, Sharon R., Brian L. Browning, Ying Zhou, Serena Tucci, and Josh M. Akey. 2018. “Analysis of Human Sequence Data Reveals Two Pulses of Archaic Denisovan Admixture.” Cell 173 (1): 53–61.
Chen, Fahu, Frido Welker, Chuan-Chou Shen, Shara E. Bailey, Inga Bergmann, Simon Davis, Huan Xia, et al. 2019. “A Late Middle Pleistocene Denisovan Mandible from the Tibetan Plateau.” Nature 569: 409–412. https://doi.org/10.1038/s41586-019-1139-x.
Churchill, Steven E. 1998. “Cold Adaptations, Heterochrony, and Neanderthals.” Evolutionary Anthropology 7 (2): 46–60.
Churchill, Steven E. 2006. “Bioenergetic Perspective on Neanderthal Thermoregulatory and Activity Budgets.” In Neanderthals Revisited: New Approaches and Perspectives, edited by K. Havarti and T. Harrison, pp. 113–134. Dordrecht, Germany: Springer.
Churchill, Steven E., Robert G. Franciscus, Hilary A. McKean-Peraza, Julie A. Daniel, and Brittany R. Warren. 2009. “Shanidar 3 Neanderthal Rib Puncture Wound and Paleolithic Weaponry.” Journal of Human Evolution 57 (2): 163–178.
Cook, Rebecca W., Antonio Vazzana, Rita Sorrentino, Stefano Benazzi, Amanda L. Smith, David S. Strait, and Justin A. Ledogar. 2021. “The Cranial Biomechanics and Feeding Performance of Homo floresiensis.” Interface Focus 11 (5). https://doi.org/10.1098/rsfs.2020.0083.
Dawson, James E., and Erik Trinkaus. 1997. “Vertebral Osteoarthritis of La Chapelle-Aux-Saints Neanderthal.” Journal of Archaeological Science 24 (11): 1015-1021. https://doi.org/10.1006/jasc.1996.0179.
Defleur, Alban R., and Emmanuel Desclaux. 2019. “Impact of the Last Interglacial Climate Change on Ecosystems and Neanderthals Behavior at Baume Moula-Guercy, Ardèche, France.” Journal of Archaeological Science 104: 114–124. https://doi.org/10.1016/j.jas.2019.01.002.
Degano, Ilaria, Sylvain Soriano, Paola Villa, Luca Pollarolo, Jeannette J. Lucejko, Zenobia Jacobs, Katerina Douka, Silvana Vitagliano, and Carlo Tozzi. 2019. “Hafting of Middle Paleolithic Tools in Latium (Central Italy): New Data from Fossellone and Sant’Agostino Caves.” PLoS ONE 14(10): 1–29. https://doi.org/10.1371/journal.pone.0213473.
Degioanni Anna, Chritophe Bonenfant, Sandrine Cabut, and Silvana Condemi. 2019. Living on the Edge: Was Demographic Weakness the Cause of Neanderthal Demise? PLoS ONE 14 (5): e0216742.
https://doi.org/10.1371/journal.pone.0216742.
Delpiano, Davide, Kristin Heasley, and Marci Peresani. 2018. “Assessing Neanderthal Land Use and Lithic Raw Material Management in Discoid Technology.” Journal of Anthropological Sciences 96: 89–110. https://doi.org/10.4436/jass.96006.
Demeter, Fabrice, Clément Zanolli, Kira E. Westaway, Renaud Joannes-Boyau, Phillippe Duringer, Mike W. Morley, Frido Welker, et al. 2022. “A Middle Pleistocene Denisovan Molar from the Annamite Chain of Northern Laos.” Nature Communications 13 (1): 2557. https://doi.org/10.7554/eLife.24231.
Dirks, Paul H. G. M., Eric M. Roberts, Hannah Hilbert-Wolf, Jan D. Kramers, John Hawks, Anthony Dosseto, Mathieu Duval, et al. 2017. “The Age of Homo naledi and Associated Sediments in the Rising Star Cave, South Africa.” eLife 6: e24231. https://doi.org/10.7554/eLife.24231.
el-Showk, Sedeer. 2019. “Neanderthal Clues to Brain Evolution in Humans.” Nature 571: S10–S11. https://doi.org/10.1038/d41586-019-02210-6.
Esteban, Irene, Curtis W. Marean, Erich C. Fisher, Panagiotis Karkanas, Dan Carbanes, and Rosa M. Albert. 2018. “Phytoliths as an Indicator of Early Modern Humans’ Plant-Gathering Strategies, Fire, Fuel, and Site-Occupation Intensity During the Middle Stone Age at Pinnacle Point 5-6 (South Coast, South Africa).” PLoS One 13 (6): 1–33.
Evteev, Andrej A., Alla A. Movsesian, and Alexandra N. Grosheva. 2017. “The Association between Mid-facial Morphology and Climate in Northeast Europe Differs from That in North Asia: Implications for Understanding the Morphology of Late Pleistocene Homo sapiens.” Journal of Human Evolution 107: 36–48. https://doi.org/10.1016/j.jhevol.2017.02.008.
Frayer, David W., Marina Lozano, Jose M. Bermudez de Castro, Eudald Carbonell, Juan-Luis Arsuaga, Jakov Radovcic, Ivana Fiore, and Luca Bondioli. 2012. “More Than 500,000 Years of Right-handedness in Europeans.” Laterality 17 (1): 51–69. https://doi.org/10.1080/1357650X.2010.529451.
Fu, Qiaomei, Mateja Hajdinjak, Oana Teodora Moldovan, Silviu Constantin, Swapan Mallick, Pontus Skoglund, Nick Patterson, et al. 2015. “An Early Modern Human from Romania with a Recent Neanderthal Ancestor.” Nature 524 (7564): 216. https://doi.org/10.1038/nature14558.
Gaudzinski-Windheuser, Sabine, Elizabeth S. Noack, Eduard Pop, Constantin Herbst, Johannes Pfleging, Jonas Buchli, Arne Jacob, et al. 2018. “Evidence for Close-Range Hunting by Last Interglacial Neanderthals.” Nature Ecology and Evolution 2: 1087–1092. https://doi.org/10.1038/s41559-018-0596-1.
Gilpin, William., Marcus W. Feldman, and Kenichi Aoki. 2016. “An Ecocultural Model Predicts Neanderthal Extinction through Competition with Modern Humans.” Proceedings of the National Academy of Sciences 113 (8): 2134–2139. https://doi.org/10.1073/pnas.1524861113.
Golovanova, Liubov. Vitaliena, Vladimir B. Doronichev, Naomi E. Cleghorn, Marianna A. Koulkova, Tatiana V. Sapelko, and M. Steven Shackley. 2010. “Significance of Ecological Factors in the Middle to Upper Paleolithic Transition.” Current Anthropology 51 (5): 655–691. https://doi.org.10.1086/656185.
Harrold, Francis B. 1980. “A Comparative Analysis of Eurasian Paleolithic Burials.” World Archaeology 12 (2): 195–211.
Hawks, John, Marina Elliott, Peter Schmid, Steven E. Churchill, Darryl J. de Ruiter, Eric M. Roberts, Hannah Hilbert-Wolf, et al. 2017. “New Fossil Remains of Homo naledi from the Lesedi Chamber, South Africa.” eLife 6: e24232. https://doi.org/10.7554/eLife.24232.
Henry, Amanda G., Alison S. Brooks, and Dolores R. Piperno. 2010. “Microfossils and Calculus Demonstrate Consumption of Plants and Cooked Foods in Neanderthal Diets (Shanidar III, Iraq; Spy I and II, Belgium).” Proceedings of the National Academy of Sciences USA 108 (2): 486–491. https://doi.org/10.1073/pnas.1016868108.
Houldcroft, Charlotte J., and Simon J. Underdown. 2016. “Neanderthal Genomics Suggests a Pleistocene Time Frame for the First Epidemiologic Transition.” American Journal of Physical Anthropology 160 (3): 379–388.
Huerta-Sánchez, Emilia, Xin Jin, Asan, Zhuoma Bianba, Benjamin M. Peter, Nicolas Vinckenbosch, Yu Liang, et al. 2014. “Altitude Adaptation in Tibetans Caused by Introgression of Denisovan-like DNA.” Nature 512 (7513): 194-197.
Jaouen, Klervia, Michael P. Richards, and Adeline Le Cabec. 2019. “Exceptionally High δ15N Values in Collagen Single Amino Acids Confirm Neanderthals as High-Trophic Level Carnivores.” Proceedings of the National Academy of Sciences 116 (11): 4928–4933. https://doi.org/10.1073/pnas.1814087116.
Kandel, Andrew W., Michael Bolus, Knut Bretzke, Angela A. Bruch, Miriam N. Haidle, Christine Hertler, and Michael Märker. 2015. “Increasing Behavioral Flexibility? An Integrative Macro-Scale Approach to Understanding the Middle Stone Age of Southern Africa.” Journal of Archaeological Method and Theory 23: 623–668. https://doi.org/10.1007/s10816-015-9254-y.
Kortlandt, Adriaan. 2002. “Neanderthal Anatomy and the Use of Spears.” Evolutionary Anthropology 11 (5): 183–184.
Kruger, A., P. Randolph-Quinney, and M. Elliott. 2016. “Multimodal Spatial Mapping and Visualization of Dinaledi Chamber and Rising Star Cave.” South African Journal of Science 112 (5–6). https://doi.org/10.17159/sajs.2016/20160032.
Lieberman, Philip, and Edmund S. Crelin. 1971. “On the Speech of Neanderthal Man.” Linguistic Inquiry 2 (2): 203–222.
Mendez, Fernando L., G. David Poznik, Sergi Castellano, and Carlos D. Bustamante. 2016. “The Divergence of Neanderthal and Modern Human Y Chromosomes.” American Journal of Human Genetics 98 (4): 728–734.
Mora-Bermúdez, Felipe, Philipp Kanis, Dominik Macak, Jula Peters, Ronald Naumann, Lei Xing, Mihail Sarov, et al. 2022. “Longer Metaphase and Fewer Chromosome Segregation Errors in Modern Human Than Neanderthal Brain Development.” Science Advances 8 (30). https://doi.org/10.1126/sciadv.abn7702.
Nicholson, Christopher M. 2017. “Eemian Paleoclimate Zones and Neanderthal Landscape Use: A GIS Model of Settlement Patterning during the Last Interglacial.” Quaternary International 438 (B): 144–157. https://doi.org/10.1016/j.quaint.2017.04.023.
Nobel Prize, The. 2022. “Press Release: The Nobel Prize in Physiology or Medicine 2022.” Nobelförsamlingen The Nobel Assembly at Karolinska Institutet. Press release, November 3, 2022. https://www.nobelprize.org/prizes/medicine/2022/press-release/.
Nowell, April. 2016. “Childhood, Play and the Evolution of Cultural Capacity in Neanderthals and Modern Humans.” In The Nature of Culture: Based on an Interdisciplinary Symposium “The Nature of Culture,” Tübingen, Germany, edited by M. N. Haidle, N. J. Conard, and M. Bolus, 87–97. Heidelberg, Germany: Springer. https://doi.org.10.1007/978-94-017-7426-0_9.
Parkington, John. 2003. “Middens and Moderns: Shellfishing and the Middle Stone Age of the Western Cape, South Africa.” South African Journal of Science 99 (5–6): 243–274.
Pearce, Eiluned, Chris Stringer, and R. I. M. Dunbar. 2013. “New Insights into Differences in Brain Organizations between Neanderthals and Anatomically Modern Humans.” Proceedings of the Royal Society B: Biological Sciences 280 (1758): 20130168. https://doi.org/10.1098/rspb.2013.0168.
Pomeroy, Emma, Paul Bennett, Chris O. Hunt, Tim Reynolds, Lucy Farr, Marine Frouin, James Holman, Ross Lane, Charles French, and Graeme Barker. 2020. “New Neanderthal Remains Associated with the ‘Flower Burial’ at Shanidar Cave.” Antiquity 94 (373): 11–26. https://doi.org/10.15184/aqy.2019.207.
Rae, Todd, Thomas Koppe, and Chris B. Stringer. 2011. “The Neanderthal Face Is Not Cold-Adapted.” Journal of Human Evolution 60 (2): 234–239. https://doi.org/10.1016/j.jhevol.2010.10.003.
Reich, David, Richard E. Green, Martin Kircher, Johannes Krause, Nick Patterson, Eric Y. Durand, Bence Viola, et al. 2010. “Genetic History of an Archaic Hominin Group from Denisova Cave in Siberia.” Nature 468: 1053–1060. https://doi.org/10.1038/nature09710.
Richards, Michael P., Paul B. Pettit, Erik Trinkaus, Fred H. Smith, Maja Paunović, and Ivor Karavanić. 2000. “Neanderthal Diet at Vindija and Neanderthal Predation: The Evidence from Stable Isotopes.” Proceedings of the National Academy of Sciences 97 (13): 7663–7666.
https://doi.org/10.1073/pnas.120178997.
Richter, Jürgen, 2006. “Neanderthals in Their Landscape: Neanderthals in Europe.” In Proceedings of the International Conference held in the Gallo-Roman Museum in Tongeren, ERAUL 117, edited by B. Demarsin and M. Otte, 17–32. Köln: Universität zu Köln.
Roksandic, Mirjana, Predrag Radovic, Xiu-Jie Wu, and Christopher J. Bae. 2021. “Resolving the ‘Muddle in the Middle’: The Case for Homo bodoensis.” Evolutionary Anthropology: Issues, News, and Reviews, 31 (1). https://doi.org/10.1002/evan.21929.
Slon, Viviane, Fabrizio Mafessoni, Benjamin Vernot, Cesare de Filippo, Steffi Grote, Bence Viola, Mateja Hajdinjak, et al. 2018. “The Genome of the Offspring of a Neanderthal Mother and a Denisovan Father.” Nature 561 (7721): 113–116.
Smith, Tanya M., Paul Tafforeau, Donald J. Reid, Joanne Pouech, Vincent Lazzari, John P. Zermeno, Debbi Guatelli-Steinberg, et al. 2010. “Dental Evidence for Ontogenetic Differences Between Modern Humans and Neanderthals.” Proceedings for the National Academy of Sciences 107 (49): 20923–20928.
Staubwasser, Michael, Virgil Drăguʂin, Bogdan P. Onac, Sergey Assonov, Vasile Ersek, Dirk L. Hoffman, and Daniel Veres. 2018. “Impact of Climate Change on the Transition of Neanderthals to Modern Humans in Europe.” Proceedings of the National Academy of Sciences 115 (37): 9116–9121. https://doi.org/10.1073/pnas.1808647115.
Stewart, T. D. 1977. “The Neanderthal Skeletal Remains from Shanidar Cave, Iraq: A Summary of Findings to Date.” Proceedings of the American Philosophical Society 121 (2): 121–165.
Stolarczyk, Regine E., and Patrick Schmidt. 2018. “Is Early Silcrete Heat Treatment a New Behavioural Proxy in the Middle Stone Age?” PLoS One 13 (10): 1–21.
Trinkaus, E. 1985. “Pathology and Posture of the La-Chapelle-aux-Saints Neanderthal.” American Journal of Physical Anthropology 67 (1): 19–41.
Trujillo Cleber A., Edward S. Rice, Nathan K. Schaefer, Isaac A. Chaim, Emily C. Wheeler, Assael A. Madrigal, Justin Buchanan, et al. 2021. “Reintroduction of the Archaic Variant of NOVA1 in Cortical Organoids Alters Neurodevelopment.” Science 371 (6530): eaax2537. https://doi.org/10.1126/science.aax2537.
Tucci, Serena, Samuel H. Vohr, Rajiv C. McCoy, Benjamin Vernot, Matthew R. Robinson, Chiara Barbieri, Brad J. Nelson, et al. 2018. “Evolutionary History and Adaptation of a Human Pygmy Population of Flores Island, Indonesia.” Science 361 (6401): 511–516.
Van Andel, T. H., and P. C. Tzedakis. 1996. “Paleolithic Landscapes of Europe and Environs, 150,000–25,000 Years Ago: An Overview.” Quaternary Science Reviews 15 (5–6): 481–500.
Venner, Stephen J. 2018. “A New Estimate for Neanderthal Energy Expenditure.” CUNY Academic Works.
Vernot, Benjamin, Serena Tucci, Janet Kelso, Joshua G. Schraiber, Aaron B. Wolf, Rachel M. Gittelman, Michael Danneman, et al. 2016. “Excavating Neanderthal and Denisovan DNA from the Genomes of Melanesian Individuals.” Science 352 (6282): 235–239.
Weaver, T. D., and J. Hublin. 2009. “Neanderthal Birth Canal Shape and the Evolution of Human Childbirth.” Proceedings of the National Academy of Sciences 106 (20): 8151–8156.
Wiẞing, Christoph, Hélène Rougier, Isabelle Crevecoer, Mietje Germonpré, Yuichi Naito, Patrick Semal, and Hervé Bocherens. 2015. “Isotopic Evidence for Dietary Ecology of Late Neanderthals in Northwestern Europe.” Quaternary International 411 (A): 327–345. https://doi.org/10.1016/j.quaint.2015.09.091.
Wong, Kate. 2015. “Neanderthal Minds.” Scientific American (January): 312(2): 36-43. https://doi.org/10.1038/scientificamerican0215-36.
Zhang, X. L., B. B. Ha, S. J. Wang, Z. J. Chen, J. Y. Ge, H. Long, W. He, et al. 2018. “The Earliest Human Occupation of the High-Altitude Tibetan Plateau 40 Thousand to 30 Thousand Years Ago.” Science 362 (6418): 1049–1051. https://doi.org/10.1126/sciadv.add5582.
Zilhão João, Diego E. Angelucci, Ernestina Badal-García, Francesco d’Errico, Floréal Daniel, Laure Dayet, Katerina Douka, et al. 2010. “Symbolic Use of Marine Shells and Mineral Pigments by Iberian Neandertals.” Proceedings of the National Academy of Sciences 107 (3): 1023–1028. https://doi.org/10.1073/pnas.0914088107.
Zollikofer, Christopher Peter Edwards, and Marcia Silvia Ponce de León. 2013. “Pandora’s Growing Box: Inferring the Evolution and Development of Hominin Brains from Endocasts.” Evolutionary Anthropology 22 (1): 20–33. https://doi.org/10.1002/evan.21333.
Acknowledgments
The authors would like to extend their thanks to Cassandra Gilmore and Anna Goldfield for thoughtful and insightful suggestions on the first edition of this chapter.
Ashley Kendell, Ph.D., California State University, Chico
Alex Perrone, M.A., M.S.N, R.N., P.H.N., Butte Community College
Colleen Milligan, Ph.D., California State University, Chico
This chapter is a revision from "Chapter 15: Bioarchaeology and Forensic Anthropology” by Ashley Kendell, Alex Peronne, and Colleen Milligan. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Content Warning and Disclaimer: This chapter includes images of human remains as well as discussions centered on human skeletal analyses. All images are derived from casts, sketches, nonhuman skeletal material, as well as non-Indigenous skeletal materials curated within the CSU, Chico Human Identification Lab, and the Hartnett-Fulginiti donated skeletal collection.
Learning Objectives
- Define forensic anthropology as a subfield of biological anthropology.
- Describe the seven steps carried out during skeletal analysis.
- Outline the four major components of the biological profile.
- Contrast the four categories of trauma.
- Explain how to identify the different taphonomic agents that alter bone.
- Discuss ethical considerations for forensic anthropology.
Forensic anthropology is a subfield of biological anthropology and an applied area of anthropology. Forensic anthropologists use skeletal analysis to gain information about humans in the present or recent past, then they apply this information within a medicolegal context. This means that forensic anthropologists specifically conduct their analysis on recently deceased individuals (typically within the last 50 years) as part of investigations by law enforcement. Forensic anthropologists can assist law enforcement agencies in several different ways, including aiding in the identification of human remains whether they are complete, fragmentary, burned, scattered, or decomposed. Additionally, forensic anthropologists can help determine what happened to the deceased at or around the time of death as well as what processes acted on the body after death (e.g., whether the remains were scattered by animals, whether they were buried in the ground, or whether they remained on the surface as the soft tissue decomposed).
Many times, because of their expertise in identifying human skeletal remains, forensic anthropologists are called to help with outdoor search-and-recovery efforts, such as locating remains scattered across the surface or carefully excavating and documenting buried remains. In other cases, forensic anthropologists recover remains after natural disasters or accidents, such as fire scenes, and can help identify whether each bone belongs to a human or an animal. Forensic anthropology spans a wide scope of contexts involving the law, including incidences of mass disasters, genocide, and war crimes.
A point that can be somewhat confusing for students is that although the term forensic is included in this subfield of biological anthropology, there are many forensic techniques that are not included in the subfield. Almost exclusively, forensic anthropology deals with skeletal analysis. While this can include the comparison of antemortem (before death) and postmortem (after death) radiographs to identify whether remains belong to a specific person, or using photographic superimposition of the cranium, it does not include analyses beyond the skeleton. For example, blood-spatter analysis, DNA analysis, fingerprints, and material evidence collection do not fall under the scope of forensic anthropology.
So, what can forensic anthropologists glean from bones alone? Forensic anthropologists can address a number of questions about a human individual based on their skeletal remains. Some of those questions are as follows: How old was the person? Was the person biologically male or female? How tall was the person? What happened to the person at or around their time of death? Were they sick? The information from the skeletal analysis can then be matched with missing persons records, medical records, or dental records, aiding law enforcement agencies with identifications and investigations.
Skeletal Analysis
Forensic anthropology relies on skeletal analysis to reveal information about the deceased. The methodology and approaches outlined below are specific to the United States. Forensic anthropological methods differ depending on the country conducting an investigation. In the United States, there are typically seven steps or questions to the process:
- Is it bone?
- Is it human?
- Is it modern or archeological?
- How many individuals are present or what is the minimum number of individuals (MNI)?
- Who is it?
- Is there evidence of trauma before or around the time of death?
- What happened to the remains after death?
Is It Bone?
One of the most important steps in any skeletal analysis starts with determining whether or not material suspected to be bone is in fact bone. Though it goes without saying that a forensic anthropologist would only carry out analysis on bone, this step is not always straightforward. Whole bones are relatively easy to identify, but determining whether or not something is bone becomes more challenging once it becomes fragmentary. As an example, in high heat such as that seen on fire scenes, bone can break into pieces. During a house fire with fatalities, firefighters watered down the burning home. After the fire was extinguished, the sheetrock (used to construct the walls of the home) was drenched and crumbled. The crumbled sheetrock was similar in color and form to burned, fragmented bone, therefore mistakable for human remains (Figure 15.1). Forensic anthropologists on scene were able to separate the bones from the construction material, helping to confirm the presence of bone and hence the presence of individual victims of the fire. In this case, forensic anthropologists were able to recognize the anatomical and layered structure of bone and were able to distinguish it from the uniform and unlayered structure of sheetrock.
As demonstrated by the example above, both the macrostructure (visible with the naked eye) and microstructure (visible with a microscope) of bone are helpful in bone identification. Bones are organs in the body made up of connective tissue. The connective tissue is hardened by a mineral deposition, which is why bone is rigid in comparison to other connective tissues such as cartilage (Tersigni-Tarrant and Langley 2017, 82–83; White and Folkens 2005, 31). In a living body, the mineralized tissue does not make up the only component of bone—there are also blood, bone marrow, cartilage, and other types of tissues. However, in dry bone, two distinct layers of the bone are the most helpful for identification. The outer layer is made up of densely arranged osseous (bone) tissue called compact (cortical) bone. The inner layer is composed of much more loosely organized, porous bone tissue whose appearance resembles that of a sponge, hence the name spongy (trabecular) bone. Knowing that most bone contains both layers helps with the macroscopic identification of bone (Figures 15.2, 15.3). For example, a piece of coconut shell might look a lot like a fragment of a human skull bone. However, closer inspection will demonstrate that coconut shell only has one very dense layer, while bone has both the compact and spongy layers.
The microscopic identification of bone relies on knowledge of osteons, or bone cells (Figure 15.4). Under magnification, bone cells are visible in the outer, compact layer of bone. The bone cells are arranged in a concentric pattern around blood vessels for blood supply. The specific shape of the cells can help differentiate, for example, a small piece of PVC (white plastic) pipe from a human bone fragment (Figure 15.5).
Is It Human?
Once it has been determined that an object is bone, the next logical step is to identify whether the bone belongs to a human or an animal. Forensic anthropologists are faced with this question in everyday practice because human versus nonhuman bone identification is one of the most frequent requests they receive from law enforcement agencies.
There are many different ways to distinguish human versus nonhuman bone. The morphology (the shape/form) of human bone is a good place for students to start. Identifying the 206 bones in the adult human skeleton and each bone’s distinguishing features (muscle attachment sites, openings and grooves for nerves and blood vessels, etc.) is fundamental to skeletal analysis.
Nevertheless, there are many animal bones and human bones that look similar. For example, the declawed skeleton of a bear paw looks a lot like a human hand, pig molars appear similar to human molars, and some smaller animal bones might be mistaken for those of an infant. To add to the confusion, fragmentary bone may be even more difficult to identify as human or nonhuman. However, several major differences between human and nonhuman vertebrate bone help distinguish the two.
Forensic anthropologists pay special attention to the density of the outer, compact layer of bone in both the cranium and in the long bones. Human cranial bone has three distinctive layers. The spongy bone is sandwiched between the outer (ectocranial) and inner (endocranial) compact layers. In most other mammals, the distinction between the spongy and compact layers is not always so definite. Secondly, the compact layer in nonhuman mammal long bones can be much thicker than observed in human bone. Due to the increased density of the compact layer, nonhuman bone tends to be heavier than human bone (Figure 15.6).
The size of a bone can also help determine whether it belongs to a human. Adult human bones are larger than subadult or infant bones. However, another major difference between human adult bones and those of a young individual or infant human can be attributed to development and growth of the epiphyses (ends of the bone). The epiphyses of human subadult bones are not fused to the shaft (Figure 15.7). Therefore, if a bone is small and it is suspected to belong to a human subadult or infant, the epiphyses would not be fused. Many small animal bones appear very similar in form compared to adult human bones, but they are much too small to belong to an adult human. Yet they can be eliminated as subadult or infant bones if the epiphyses are fused to the shaft.
Is It Modern or Archaeological?
Forensic anthropologists work with modern cases that fall within the scope of law enforcement investigations. Accordingly, it is important to determine whether discovered human remains are archaeological or forensic in nature. Human remains that are historic are considered archeaological. The scientific study of human remains from archaeological sites is called bioarchaeology.
Dig Deeper: Bioarchaeology
For readers who are interested in the sister subfield of bioarchaeology, which studies human remains and material culture from the past, please refer to chapter 8 of Bioarchaeology: Interpreting Human Behavior from Skeletal Remains, in TRACES: An Open Invitation to Archaeology (Blatt, Michael, and Bright forthcoming).
A forensic anthropologist should begin their analysis by reviewing the context in which the remains were discovered. This will help them understand a great deal about the remains, including determining whether they are archaeological or forensic in nature as well as considering legal and ethical issues associated with the collection, analysis, and storage of human remains (see “Ethics and Human Rights” section of this chapter for more information).
The “context” refers to the relationship the remains have to the immediate area in which they were found. This includes the specific place where the remains were found, the soil or other organic matter immediately surrounding the remains, and any other objects or artifacts in close proximity to the body. For example, imagine that a set of remains has been located during a house renovation. The remains are discovered below the foundation. Do the remains belong to a murder victim? Or was the house built on top of an ancient burial ground? Observing information from the surroundings can help determine whether the remains are archaeological or modern. How long ago was the foundation of the house erected? Are there artifacts in close proximity to the body, such as clothing or stone tools? These are questions about the surroundings that will help determine the relative age of the remains.
Clues directly from the skeleton may also indicate whether the remains are archaeological or modern. For example, tooth fillings can suggest that the individual was alive recently (Figure 15.8). In fact, filling material has changed over the decades, so the specific type of material used to fix a cavity can be matched with specific time periods. Gold was used in dental work in the past, but more recently composite (a mixture of plastic and fine glass) fillings have become more common.
How Many Individuals Are Present?
What Is MNI?
Another assessment that an anthropologist can perform is the calculation of the number of individuals in a mixed burial assemblage. Because not all burials consist of a single individual, it is important to burial assemblage be able to estimate the number of individuals in a forensic context. Quantification of the number of individuals in a burial assemblage can be done through the application of a number of methods, including the following: the Minimum Number of Individuals (MNI), the Most Likely Number of Individuals (MLNI), and the Lincoln Index (LI). The most commonly used method in biological anthropology, and the focus of this section, is determination of the MNI.
The MNI presents “the minimum estimate for the number of individuals that contributed to the sample” (Adams and Konigsberg 2008, 243). Many methods of calculating MNI were originally developed within the field of zooarchaeology for use on calculating the number of individuals in faunal or animal assemblages (Adams and Konigsberg 2008, 241). What MNI calculations provide is a lowest possible count for the total number of individuals contributing to a skeletal assemblage. Traditional methods of calculating MNI include separating a skeletal assemblage into categories according to the individual bone and the side the bone comes from and then taking the highest count per category and assigning that as the minimum number (Figure 15.9).
Why Calculate MNI?
In a forensic context, the determination of MNI is most applicable in cases of mass graves, commingled burials, and mass fatality incidents. The term commingled is applied to any burial assemblage in which individual skeletons are not separated into separate burials. As an example, the authors of this chapter have observed commingling of remains resulting from mass fatality wildfire events. Commingled remains may also be encountered in events such as a plane or vehicle crash. It is important to remember that in any forensic context, MNI should be referenced and an MNI of one should be substantiated by the fact that there was no repetition of elements associated with the case.
Constructing the Biological Profile
Who Is It?
“Who is it?” is one of the first questions that law enforcement officers ask when they are faced with a set of skeletal remains. To answer this question, forensic anthropologists construct a biological profile (White and Folkens 2005, 405). A biological profile is an individual’s identifying characteristics, or biological information, which include the following: biological sex, age at death, stature, population affinity, skeletal variation, and evidence of trauma and pathology.
Forensic anthropologists typically construct a biological profile to help positively identify a deceased person. The following section will lay out each component of the biological profile and briefly review standard methodology used for each.
Assessing Biological Sex
Assessment of biological sex is often one of the first things considered when establishing a biological profile because several other parts, such as age and stature estimations, rely on an assessment of biological sex to make the calculations more accurate.
Assessment of biological sex focuses on differences in both morphological (form or structure) and metric (measured) traits in individuals. When assessing morphological traits, the skull and the pelvis are the most commonly referenced areas of the skeleton. These differences are related to sexual dimorphism usually varying in the amount of robusticity seen between males and females. Robusticity deals with strength and size; it is frequently used as a term to describe a large size or thickness. In general, males will show a greater degree of robusticity than females. For example, the length and width of the mastoid process, a bony projection located behind the opening for the ear, is typically larger in males. The mastoid process is an attachment point for muscles of the neck, and this bony projection tends to be wider and longer in males. In general, cranial features tend to be more robust in males (Figure 15.10).
When considering the pelvis, the features associated with the ability to give birth help distinguish females from males. During puberty, estrogen causes a widening of the female pelvis to allow for the passage of a baby. Several studies have identified specific features or bony landmarks associated with the widening of the hips, and this section will discuss one such method. The Phenice Method (Phenice 1969) is traditionally the most common reference used to assess morphological characteristics associated with sex. The Phenice Method specifically looks at the presence or absence of (1) a ventral arc, (2) the presence or absence of a subpubic concavity, and (3) the width of the medial aspect of the ischiopubic ramus (Figure 15.11). When present, the ventral arc, a ridge of bone located on the ventral surface of the pubic bone, is indicative of female remains. Likewise the presence of a subpubic concavity and a narrow medial aspect of the ischiopubic ramus is associated with a female sex estimation. Assessments of these features, as well as those of the skull (when both the pelvis and skull are present), are combined for an overall estimation of sex.
Metric analyses are also used in the estimation of sex. Measurements taken from every region of the body can contribute to estimating sex through statistical approaches that assign a predictive value of sex. These approaches can include multiple measurements from several skeletal elements in what is called multivariate (multiple variables) statistics. Other approaches consider a single measurement, such as the diameter of the head of the femur, of a specific element in a univariate (single variable) analysis (Berg 2017, 152–156).
It is important to note that, although forensic anthropologists usually begin assessment of biological profile with biological sex, there is one major instance in which this is not appropriate. The case of two individuals found in California, on July 8, 1979, is one example that demonstrates the effect age has on the estimation of sex. The identities of the two individuals were unknown; therefore, law enforcement sent them to a lab for identification. A skeletal analysis determined that the remains represented one adolescent male and one adolescent female, both younger than 18 years of age. This information did not match with any known missing children at the time.
In 2015, the cold case was reanalyzed, and DNA samples were extracted. The results indicated that the remains were actually those of two girls who went missing in 1978. The girls were 15 years old and 14 years old at the time of death. It is clear that the 1979 results were incorrect, but this mistake also provides the opportunity to discuss the limitations of assessing sex from a subadult skeleton.
Assessing sex from the human skeleton is based on biological and genetic traits associated with females and males. These traits are linked to differences in sexual dimorphism and reproductive characteristics between females and males. The link to reproductive characteristics means that most indicators of biological sex do not fully manifest in prepubescent individuals, making estimations of sex unreliable in younger individuals (SWGANTH 2010b). This was the case in the example of the 14-year-old girl. When examined in 1979, her remains were misidentified as male because she had not yet fully developed female pelvic traits.
Sex vs. Gender
Biological sex is a different concept than gender. While biological anthropologists can estimate sex from the skeleton, estimating an individual’s gender would require a greater context because gender is defined culturally rather than biologically. Take, for example, an individual who identifies as transgender. This individual has a gender identity that is different from their biological sex. The gender identity of any individual depends on factors related to self-identification, situation or context, and cultural factors. While in the U.S. we have historically thought of sex and gender as binary concepts (male or female), many cultures throughout the world recognize several possible gender identities. In this sense, gender is seen as a continuous or fluid variable rather than a fixed one.
Historically, forensic anthropologists have used a binary construct to categorize human skeletal remains as either male or female (with the accompanying categories of probable male, probable female, and indeterminate). In the case of transgender and gender nonconforming individuals, the binary approach to sex assessment may delay or hinder identification efforts (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020; Tallman, Kincer, and Plemons 2021). As such, many forensic anthropologists have begun to address the inherent problems associated with a binary approach to sex identification and to explore ways of assessing social identity and self-identified gender using skeletal remains and forensic context.
For the duration of this section, the term transgender refers to individuals whose gender identity differs from the sex assigned at birth (Schall, Rogers, and Deschamps-Braly 2020:2). Transgender individuals transition from one gender binary to another, such as male-to-female (MTF) or female-to-male (FTM). While many of the gender-affirming procedures available to trans and gender-nonconforming individuals are focused on soft tissue modifications (e.g., breast augmentation, genital reconstruction, hormone therapies, etc.), there are a number of gender-affirmation surgeries that do leave a permanent record on the skeleton. Generally speaking, FTM transgender people are reported to undergo fewer surgical procedures than do MTF transgender people (Buchanan 2014). The discussion below focuses on Facial Feminization Surgery (FFS), which leaves a permanent record on the human skeleton that may be used to help make an identification.
FFS refers to a combination of procedures focused on sexually dimorphic features of the face, with the intent of transforming typically male facial features into more feminine forms. Facial Feminization Surgery procedures were developed by Dr. Douglas Ousterhout, a San Francisco based cranio-maxillofacial surgeon, in the mid-1980s (Schall, Rogers, and Deschamps-Braly 2020:2). FFS can include one or a combination of the following: hairline lowering, forehead reduction and contouring, brow lift, reduction rhinoplasty, cheek enhancement, lift lift, lip filling, chin contouring, jaw contouring, and/or tracheal shave (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020:2). Of the procedures outlined previously, four are known to directly affect the facial skeleton: forehead contouring, rhinoplasty, chin contouring, and jaw contouring (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020:2).
Because FFS procedures have been widely documented in the medical (and more recently the forensic anthropological) literature, there are a number of indicators that a forensic anthropologist can use to make more informed evaluations of gender, including evidence of bone remodeling in sexually dimorphic regions of the skull (e.g., forehead, chin, jawline), as well as the presence of plates, pins, or other surgical hardware that may be evidence of FFS (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020; Tallman, Kincer, and Plemons 2021). Additionally, some forensic anthropologists suggest cautiously integrating contextual information from the scene, such as personal effects, material evidence, and recovery scene information, into their evaluation of an individual’s social identity (Beatrice and Soler 2016; Birkby, Fenton, and Anderson 2008; Soler and Beatrice 2018; Soler et al. 2019; Tallman, Kincer, and Plemons 2021; Winburn, Schoff, and Warren 2016). The ultimate goal of many skeletal analyses is to make a positive identification on a set of unidentified remains.
Assessment of Population Affinity
Population affinity is another component of the biological profile. We use the term population affinity to refer to the variation seen among modern populations—variation that is both genetic and environmentally driven. The word affinity refers to similarities or relationships between individuals. As forensic anthropologists, we compare an unknown individual to multiple reference groups and look for the degree of similarity in observable traits with those groups. As noted previously, population affinity can aid law enforcement in their identification of missing persons or unknown skeletal remains.
Within the field of anthropology, the estimation of population affinity has a contentious history, and early attempts at classification were largely based on the erroneous assumption that an individual’s phenotype (outward appearance) was correlated with their innate intelligence and abilities (see Chapter 13 for a more in-depth discussion of the history of the race concept). The use of the term race is deeply embedded in the social context of the United States. In any other organism/living thing, groups divided according to the biological race concept would be defined as a separate subspecies. The major issue with applying the biological race concept to humans is that there are not enough differences between any two populations to separate on a genetic basis. In other words, biological races do not exist in human populations. However, the concept of race has been perpetuated and upheld by sociocultural constructs of race.
The conundrum for forensic anthropologists is the fact that while races do not exist on a biological level, we still socially recognize and categorize individuals based on their phenotype. Clearly, our phenotype is an important factor in not only how we are viewed by others but also how we identify ourselves. It is also a commonly reported variable. Often labeled as “race,” we are asked to report how we self-identify on school applications, government identification, surveys, census reports, and so forth. It follows then that when a person is reported missing, the information commonly collected by law enforcement and sometimes entered into a missing person’s database includes their age, biological sex, stature, and “race.” Therefore, the more information a forensic anthropologist can provide regarding the individual’s physical characteristics, the more he or she can help to narrow the search.
As an exercise, create a list of all of the women you know who are between the ages of 18 and 24 and approximately 5’ 4” to 5’ 9” tall. You probably have several dozen people on the list. Now, consider how many females you know who are between the ages of 18 and 24, are approximately 5’ 4” to 5’ 9” tall, and are Vietnamese. Your list is going to be significantly shorter. That’s how missing persons searches go as well. The more information you can provide regarding a decedent’s phenotype, the fewer possible matches law enforcement are left to investigate. This is why population affinity has historically been included as a part of the biological profile.
(Put at the beginning of this serction)In an effort to combat the erroneous assumptions tied to the race concept, forensic anthropologists have attempted to reframe this component of the biological profile. The term race is no longer used in casework and teaching. Historically, the word ancestry is and was deemed a more appropriate way to describe an individual’s phenotype. However, in more recent years, forensic anthropologists have begun using the term population affinity, recognizing that we are basing our analysis on the similarities we see based on the reference samples we have available (Winburn and Algee-Hewitt 2021). An important note here is that it is possible to hinder identifications and harm individuals when tools like estimations of population affinity are misapplied, misinterpreted, or misused. For this reason, the field of forensic anthropology has ongoing conversations about the appropriateness of this analysis in the biological profile (Bethard and DiGangi 2020; Stull et al. 2021).
Traditionally, population affinity was accomplished through a visual inspection of morphological variants of the skull (morphoscopics). These methods focused on elements of the facial skeleton, including the nose, eyes, and cheek bones. However, in an effort to reduce subjectivity, nonmetric cranial traits are now assessed within a statistical framework to help anthropologists better interpret their distribution among living populations (Hefner and Linde 2018). Based on the observable traits, a macromorphoscopic analysis will allow the practitioner to create a statistical prediction of geographic origin. In essence, forensic anthropologists are using human variation in the estimation of geographic origin, by referencing documented frequencies of nonmetric skeletal indicators or macromorphoscopic traits.
Population affinity is also assessed through metric analyses. The computer program Fordisc is an anthropological tool used to estimate different components of the biological profile, including ancestry, sex, and stature. When using Fordisc, skeletal measurements are input into the computer software, and the program employs multivariate statistical classification methods, including discriminant function analysis, to generate a statistical prediction for the geographic origin of unknown remains based on the comparison of the unknown to the reference samples in the software program. Fordisc also calculates the likelihood of the prediction being correct, as well as how typical the metric data is for the assigned group.
Estimating Age-at-Death
Estimating age-at-death from the skeleton relies on the measurement of two basic physiological processes: (1) growth and development and (2) degeneration (or aging). From fetal development on, our bones and teeth grow and change at a predictable rate. This provides for relatively accurate age estimates. After our bones and teeth cease to grow and develop, they begin to undergo structural changes, or degeneration, associated with aging. This does not happen at such predictable rates and, therefore, results in less accurate or larger age-range estimations.
During growth and development stages, two primary methods used for estimations of age of subadults (those under the age of 18) are epiphyseal union and dental development. Epiphyseal union (or epiphyseal fusion) refers to the appearance and closure of the epiphyseal plates between the primary centers of growth in a bone and the subsequent centers of growth (see Figure 15.7). Prior to complete union, the cartilaginous area between the primary and secondary centers of growth is also referred to as the growth plates (Schaefer, Black, and Scheuer 2009). Different areas of the skeleton have documented differences in the appearance and closure of epiphyses, making this a reliable method for aging subadult remains (SWGANTH 2013).
As an example of its utility in the identification process, epiphyseal development was used to identify two subadult victims of a fatal fire in Flint, Michigan, in February 2010. The remains represented two young girls, ages three and four. Due to the intensity of the fire, the subadult victims were differentiated from each other through the appearance of the patella, the kneecap. The patella is a bone that develops within the tendon of the quadriceps muscle at the knee joint. The patella begins to form around three to four years of age (Cunningham, Scheuer, and Black 2016, 407–409). In the example above, radiographs of the knees showed the presence of a patella in the four-year-old girl and the absence of a clearly discernible patella in the three-year-old.
Dental development begins during fetal stages of growth and continues until the complete formation and eruption of the adult third molars (if present). The first set of teeth to appear are called deciduous or baby teeth. Individuals develop a total of 20 deciduous teeth, including incisors, canines, and molars. These are generally replaced by adult dentition as an individual grows (Figure 15.12). A total of 32 teeth are represented in the adult dental arcade, including incisors, canines, premolars, and molars. When dental development is used for age estimations, researchers use both tooth-formation patterns and eruption schedules as determining evidence. For example, the crown of the tooth forms first followed by the formation of the tooth root. During development, an individual can exhibit a partially formed crown or a complete crown with a partially formed root. The teeth generally begin the eruption process once the crown of the tooth is complete. The developmental stages of dentition are one of the most reliable and consistent aging methods for subadults (Langley, Gooding, and Tersigni-Tarrant 2017, 176–177).
Degenerative changes in the skeleton typically begin after 18 years of age, with more prominent changes developing after an individual reaches middle adulthood (commonly defined as after 35 years of age in osteology). These changes are most easily seen around joint surfaces of the pelvis, the cranial vault, and the ribs. In this chapter, we focus on the pubic symphysis surfaces of the pelvis and the sternal ends of the ribs, which show metamorphic changes from young adulthood to older adulthood. The pubic symphysis is a joint that unites the left and right halves of the pelvis. The surface of the pubic symphysis changes during adulthood, beginning as a surface with pronounced ridges (called billowing) and flattening with a more distinct rim to the pubic symphysis as an individual ages. As with all metamorphic age changes, older adults tend to develop lipping around the joint surfaces as well as a breakdown of the joint surfaces. The most commonly used method for aging adult skeletons from the pubic symphysis is the Suchey-Brooks method (Brooks and Suchey 1990; Katz and Suchey 1986). This method divides the changes seen with the pubic symphysis into six phases based on macroscopic age-related changes to the surface. Figure 15.13 provides a visual of the degenerative changes that typically occur on the pubic symphysis.
The sternal end of the ribs, the anterior end of the rib that connects via cartilage to the sternum, is also used in age estimations of adults. This method, first developed by M. Y. İşcan and colleagues, considers both the change in shape of the sternal end as well as the quality of the bone (İşcan, Loth, and Wright 1984; İşcan, Loth, and Wright 1985). The sternal end first develops a billowing appearance in young adulthood. The bone typically develops a wider and deeper cupped end as an individual ages. Older adults tend to exhibit bony extensions of the sternal end rim as attaching cartilage ossifies. Figure 15.14 provides a visual of the degenerative changes that typically occur in sternal rib ends.
Estimating Stature
Stature, or height, is one of the most prominently recorded components of the biological profile. Our height is recorded from infancy through adulthood. Doctor’s appointments, driver's license applications, and sports rosters all typically involve a measure of stature for an individual. As such, it is also a component of the biological profile nearly every individual will have on record. Bioarchaeologists and forensic anthropologists use stature estimation methods to provide a range within which an individual’s biological height would fall. Biological height is a person’s true anatomical height. However, the range created through these estimations is often compared to reported stature, which is typically self-reported and based on an approximation of an individual’s true height (Ousley 1995).
In June 2015, two men were shot and killed in Granite Bay, California, in a double homicide. Investigators were able to locate surveillance camera footage from a gas station where the two victims were spotted in a car with another individual believed to be the perpetrator in the case. The suspect, sitting behind the victims in the car, hung his right arm out of the window as the car drove away. The search for the perpetrator was eventually narrowed down to two suspects. One suspect was 5’ 8” while the other suspect was 6’ 4”, representing almost a foot difference in height reported stature between the two. Forensic anthropologists were given the dimensions of the car (for proportionality of the arm) and were asked to calculate the stature of the suspect in the car from measurements of the suspect’s forearm hanging from the window. Approximate lengths of the bones of the forearm were established from the video footage and used to create a predicted stature range. Stature estimations from skeletal remains typically look at the correlation between the measurements of any individual bone and the overall measurement of body height. In the case above, the length of the right forearm pointed to the taller of the two suspects who was subsequently arrested for the homicide.
Certain bones, such as the long bones of the leg, contribute more to our overall height than others and can be used with mathematical equations known as regression equations. Regression methods examine the relationship between variables such as height and bone length and use the correlation between the variables to create a prediction interval (or range) for estimated stature. This method for calculating stature is the most commonly used method (SWGANTH 2012). Figure 15.15 shows the measurement of the bicondylar length of the femur for stature estimations.
Identification Using Individualizing Characteristics
One of the most frequently requested analyses within the forensic anthropology laboratory is assistance with the identification of unidentified remains. While all components of a biological profile, as discussed above, can assist law enforcement officers and medical examiners to narrow down the list of potential identifications, a biological profile will not lead to a positive identification. The term positive identification refers to a scientifically validated method of identifying previously unidentified remains. Presumptive identifications, however, are not scientifically validated; rather, they are based on circumstances or scene context. For example, if a decedent is found in a locked home with no evidence of forced entry but the body is no longer visually identifiable, it may be presumed that the remains belong to the homeowner. Hence, a presumptive identification.
The medicolegal system ultimately requires that a positive identification be made in such circumstances, and a presumptive identification is often a good way to narrow down the pool of possibilities. Biological profile information also assists with making a presumptive identification based on an individual’s phenotype in life (e.g., what they looked like). As an example, a forensic anthropologist may establish the following components of a biological profile: white male, between the ages of 35 and 50, approximately 5’ 7” to 5’ 11.” While this seems like a rather specific description of an individual, you can imagine that this description fits dozens, if not hundreds, of people in an urban area. Therefore, law enforcement can use the biological profile information to narrow their pool of possible identifications to include only white males who fit the age and height outlined above. Once a possible match is found, the decedent can be identified using a method of positive identification.
Positive identifications are based on what we refer to as individualizing traits or characteristics, which are traits that are unique at the individual level. For example, brown hair is not an individualizing trait as brown is the most common hair color in the U.S. But, a specific pattern of dental restorations or surgical implants can be individualizing, because it is unlikely that you will have an exact match on either of these traits when comparing two individuals.
A number of positive methods are available to forensic anthropologists, and for the remainder of this section we will discuss the following methods: comparative medical and dental radiography and identification of surgical implants.
Comparative medical and dental radiography is used to find consistency of traits when comparing antemortem records (medical and dental records taken during life) with images taken postmortem (after death). Comparative medical radiography focuses primarily on features associated with the skeletal system, including trabecular pattern (internal structure of bone that is honeycomb in appearance), bone shape or cortical density (compact outer layer of bone), and evidence of past trauma, skeletal pathology, or skeletal anomalies. Other individualizing traits include the shape of various bones or their features, such as the frontal sinuses (Figure 15.16).
Comparative dental radiography focuses on the number, shape, location, and orientation of dentition and dental restorations in antemortem and postmortem images. While there is not a minimum number of matching traits that need to be identified for an identification to be made, the antemortem and postmortem records should have enough skeletal or dental consistencies to conclude that the records did in fact come from the same individual (SWGANTH 2010a). Consideration should also be given to population-level frequencies of specific skeletal and dental traits. If a trait is particularly common within a given population, it may not be a good trait to utilize for positive identification.
Surgical implants or devices can also be used for identification purposes (Figure 15.17). These implements are sometimes recovered with human remains. One of the ways forensic anthropologists can use surgical implants to assist in decedent identification is by providing a thorough analysis of the implant and noting any identifying information such as serial numbers, manufacturer symbols, and so forth. This information can then sometimes be tracked directly to the manufacturer or the place of surgical intervention, which may be used to identify unknown remains (SWGANTH 2010a).
Special Topic: Trans Doe Task Force
The Trans Doe Task Force (TDTF) is a Trans-led nonprofit organization that investigates cases involving LGBTQ+ missing and murdered persons. The organization specifically focuses on transgender and gender-variant cases, providing connections between law enforcement agencies, medical examiner offices, forensic anthropologists, and forensic genetic genealogists to increase the chances of identification. Additionally, the TDTF curates a data repository of missing, murdered, and unclaimed LGBTQ+ individuals, and they continuously try innovative approaches to identify these individuals, whose lived gender identity may not match their biological sex.
For more information visit transdoetaskforce.org
Trauma Analysis
Types of Trauma
Within the field of anthropology, trauma is defined as an injury to living tissue caused by an extrinsic force or mechanism (Lovell 1997:139). Forensic anthropologists can assist a forensic pathologist by providing an interpretation of the course of events that led to skeletal trauma. Typically, traumatic injury to bone is classified into one of four categories, defined by the trauma mechanism. A trauma mechanism refers to the force that produced the skeletal modification and can be classified as (1) sharp force, (2) blunt force, (3) projectile, or (4) thermal (burning). Each type of trauma, and the characteristic pattern(s) associated with that particular categorization, will be discussed below.
First, let’s consider sharp-force trauma, which is caused by a tool that is edged, pointed, or beveled—for example, a knife, saw, or machete (SWGANTH 2011). The patterns of injury resulting from sharp-force trauma include linear incisions created by a sharp, straight edge; punctures; and chop marks (Figure 15.18; SWGANTH 2011). When observed under a microscope, an anthropologist can often determine what kind of tool created the bone trauma. For example, a power saw cut will be discernible from a manual saw cut.
Second, blunt-force trauma is defined as “a relatively low-velocity impact over a relatively large surface area” (Galloway 1999, 5). Blunt-force injuries can result from impacts from clubs, sticks, fists, and so forth. Blunt-force impacts typically leave an injury at the point of impact but can also lead to bending and deformation in other regions of the bone. Depressions, fractures, and deformation at and around the site of impact are all characteristics of blunt-force trauma (Figure 15.19). As with sharp-force trauma, an anthropologist attempts to interpret blunt-force injuries, providing information pertaining to the type of tool used, the direction of impact, the sequence of impacts, if more than one, and the amount of force applied.
Third, projectile trauma refers to high-velocity trauma, typically affecting a small surface area (Galloway 1999, 6). Projectile trauma results from fast-moving objects such as bullets or shrapnel. It is typically characterized by penetrating defects or embedded materials (Figure 15.20). When interpreting injuries resulting from projectile trauma, an anthropologist can often offer information pertaining to the type of weapon used (e.g., rifle vs. handgun), relative size of the bullet (but not the caliber of the bullet), the direction the projectile was traveling, and the sequence of injuries if there are multiple present.
Finally, thermal trauma is a bone alteration that results from bone exposure to extreme heat. Thermal trauma can result in cases of house or car fires, intentional disposal of a body in cases of homicidal violence, plane crashes, and so on. Thermal trauma is most often characterized by color changes to bone, ranging from yellow to black (charred) or white (calcined). Other bone alterations characteristic of thermal trauma include delamination (flaking or layering due to bone failure), shrinkage, fractures, and heat-specific burn patterning. When interpreting injuries resulting from thermal damage, an anthropologist can differentiate between thermal fractures and fractures that occurred before heat exposure, thereby contributing to the interpretation of burn patterning (e.g., was the individual bound or in a flexed position prior to the fire?).
While there are characteristic patterns associated with the four categories of bone trauma, it is also important to note that these bone alterations do not always occur independently of different trauma types. An individual’s skeleton may present with multiple different types of trauma, such as a projectile wound and thermal trauma. Therefore, it is important that the anthropologist recognize the different types of trauma and interpret them appropriately.
Timing of Injury
Another important component of any anthropological trauma analysis is the determination of the timing of injury (e.g., when did the injury occur). Timing of injury is traditionally split into one of three categories: antemortem (before death), perimortem (at or around the time of death), and postmortem (after death). This classification system differs slightly from the classification system used by the pathologist because it specifically references the qualities of bone tissue and bone response to external forces. Therefore, the perimortem interval (at or around the time of death) means that the bone is still fresh and has what is referred to as a green bone response, which can extend past death by several weeks or even months. For example, in cold or freezing temperatures a body can be preserved for extended periods of time, increasing the perimortem interval, while in desert climates decomposition is accelerated, thereby significantly decreasing the postmortem interval (Galloway 1999, 12). Antemortem injuries (occurring well before death and not related to the death incident) are typically characterized by some level of healing, in the form of a fracture callus or unification of fracture margins. Finally, postmortem injuries (occurring after death, while bone is no longer fresh) are characterized by jagged fracture margins, resulting from a loss of moisture content during the decomposition process (Galloway 1999, 16). In general, all bone traumas should be classified according to the timing of injury, if possible. This information will help the medical examiner or pathologist better understand the circumstances surrounding the decedent’s death, as well as events occurring during life and after the final disposition of the body.
The Role of the Forensic Anthropologist in Trauma Analysis
Within the medicolegal system, forensic anthropologists are often called upon by the medical examiner, forensic pathologist, or coroner to assist with an interpretation of trauma. The forensic anthropologist’s main focus in any trauma analysis is the underlying skeletal system—as well as, sometimes, cartilage. Analysis and interpretation of soft tissue injuries fall within the purview of the medical examiner or pathologist. It is also important to note that the main role of the forensic anthropologist is to provide information pertaining to skeletal injury to assist the medical examiner/pathologist in their final interpretation of injury. Forensic anthropologists do not hypothesize as to the cause of death of an individual. Instead, a forensic anthropologist’s report should include a description of the injury (e.g., trauma mechanism, number of injuries, location, timing of injury); documentation of the injury, which may be utilized in court testimony (e.g., photographs, radiographs, measurements); and, if applicable, a statement as to the condition of the body and state of decomposition, which may be useful for understanding the depositional context (e.g., how long has the body been exposed to the elements; was it moved or in its original location; are any of the alterations to bone due to environmental or faunal exposure instead of intentional human modification).
Taphonomy
What Happened to the Remains After Death?
The majority of the skeletal analysis process revolves around the identity of the deceased individual. However, there is one last, very important question that forensic anthropologists should ask: What happened to the remains after death? Generally speaking, processes that alter the bone after death are referred to as taphonomic changes (refer to Chapter 7 for a discussion regarding taphonomy and the fossil record).
The term taphonomy was originally used to refer to the processes through which organic remains mineralize, also known as fossilization. Within the context of biological anthropology, the term taphonomy is better defined as the study of what happens to human remains after death (Komar and Buikstra 2008). Initial factors affecting a body after death include processes such as decomposition and scavenging by animals. However, taphonomic processes encompass much more than the initial period after death. For example, plant root growth can leach minerals from bone, leaving a distinctive mark. Sunlight can bleach human remains, leaving exposed areas whiter than those that remained buried. Water can wear the surface of the bone until it becomes smooth.
Some taphonomic processes can help a forensic anthropologist estimate the relative amount of time that human remains have been exposed to the elements. For example, root growth through a bone would certainly indicate a body was buried for more than a few days. Forensic anthropologists must be very careful when attempting to estimate time since death based on taphonomic processes because environmental conditions can greatly influence the rate at which taphonomic processes progress. For example, in cold environments, tissue may decay slower than in warm, moist environments.
Forensic anthropologists must contend with taphonomic processes that affect the preservation of bones. For example, high acidity in the soil can break down human bone to the point of crumbling. In addition, when noting trauma, they must be very careful not to confuse postmortem (after death) bone damage with trauma.
Taphonomic Process |
Definition |
Rodent Gnawing |
When rodents, such as rats and mice, chew on bone, they leave sets of parallel grooves. The shallow grooves are etched by the rodent’s incisors. |
Carnivore Damage
|
Carnivores may leave destructive dental marks on bone. The tooth marks may be visible as pit marks or punctures from the canines, as well as extensive gnawing or chewing of the ends of the bones to retrieve marrow. |
Burned Bone |
Fire causes observable damage to bone. Temperature and the amount of time bone is heated affect the appearance of the bone. Very high temperatures can crack bone and result in white coloration. Color gradients are visible in between high and lower temperatures, with lower temperatures resulting in black coloration from charring. Cracking can also reveal information about the directionality of the burn. |
Root Etching |
Plant roots can etch the outer surface of bone, leaving grooves where the roots attached as they leached nutrients. During this process, the plant’s roots secrete acid that breaks down the surface of the bone.
|
Weathering
|
Many different environmental conditions affect bone. River transport can smooth the surface of the bone due to water abrasion. Sunlight can bleach the exposed surface of bone. Dry and wet environments or the mixture of both types of environments can cause cracking and exfoliation of the surface. Burial in different types of soil can cause discoloration, and exposure can cause degreasing. |
Cut Marks |
Humans may alter bone by cutting, scraping, or sawing it directly or in the process of removing tissue. The groove pattern—that is, the depth and width of the cuts—can help identify the tool used in the cutting process. |
Ethics and Human Rights
Working with human remains requires a great deal of consideration and respect for the dead. Forensic anthropologists have to think about the ethics of our use of human remains for scientific purposes. How do we conduct casework in the most respectable manner possible? While there are a wide range of ethical considerations to consider when contemplating a career in forensic anthropology, this chapter will focus on two major categories: working with human remains and acting as an expert within the medicolegal system.
Working with Human Remains
Forensic anthropologists work with human remains in a number of contexts, including casework, excavation, research, and teaching. When working with human remains, it is always important to use proper handling techniques. To prevent damage to skeletal remains, bones should be handled over padded surfaces. Skulls should never be picked up by placing fingers in the eye orbits, foramen magnum (hole at the base of the skull for entry of the spinal cord), or through the zygomatic arches (cheekbones). Human remains, whether related to casework, fieldwork, donated skeletal collections, or research, were once living human beings. It is important to always bear in mind that work with remains should be ingrained with respect for the individual and their relatives. In addition to fieldwork, casework, and teaching, anthropologists are often invited to work with remains that come from a bioarchaeological context or from a human rights violation. While this discussion of ethics is not comprehensive, two case examples will be provided below in which an anthropologist must consider the ethical standards outlined above.
Modern Human Rights Violations
Forensic anthropologists may also be called to participate in criminal investigations involving human rights violations. Anthropological investigations may include assistance with identifications, determination of the number of victims, and trauma analyses. In this role, forensic anthropologists play an integral part in promoting human rights, preventing future human rights violations, and providing the evidence necessary to prosecute those responsible for past events. A few ethical considerations for the forensic anthropologist involved in human rights violations include the use of appropriate standards of identification, presenting reliable and unbiased testimony, and maintaining preservation of evidence. For a more comprehensive history of forensic anthropological contributions to human rights violations investigations, see Ubelaker 2018.
Acting as an Expert in the Medicolegal System
In addition to the ethical considerations involved in working with human skeletal remains, forensic anthropologists must abide by ethical standards when they act as experts within the medicolegal system. The role of the forensic anthropologist within the medicolegal system is primarily to provide information to the medical examiner or coroner that will aid in the identification process or determination of cause and manner of death. Forensic anthropologists also may be called to testify in a court of law. In this capacity, forensic anthropologists should always abide by a series of ethical guidelines that pertain to their interpretation, presentation, and preservation of evidence used in criminal investigations. First and foremost, practitioners should never misrepresent their training or education. When appropriate, outside opinions and assistance in casework should be requested (e.g., consulting a radiologist for radiological examinations or odontologist for dental exams). The best interest of the decedent should always take precedence. All casework should be conducted in an unbiased way, and financial compensation should never be accepted as it can act as an incentive to take a biased stance regarding casework. All anthropological findings should be kept confidential, and release of information is best done by the medical examiner or coroner. Finally, while upholding personal ethical standards, forensic anthropologists are also expected to report any perceived ethical violations committed by their peers.
Ethical standards for the field of forensic anthropology are outlined by the Organization of Scientific Area Committees (OSAC) for Forensic Science, administered by the National Institute of Standards and Technology (NIST). OSAC and NIST recently began an initiative to develop standards that would strengthen the practice of forensic science both in the United States and internationally. OSAC’s main objective is to “strengthen the nation’s use of forensic science by facilitating the development of technically sound forensic science standards and by promoting the adoption of those standards by the forensic science community” (NIST n.d.). Additionally, OSAC promotes the establishment of best practices and other guidelines to ensure that forensic science findings and their presentation are reliable and reproducible (NIST 2023).
Special Topic: Native American Graves Protection and Repatriation Act (NAGPRA)
There is a long history in the United States of systematic disenfranchisement of Native American people, including lack of respect for tribal sovereignty. This includes the egregious treatment of Native American human remains. Over several centuries, thousands of Native American remains were removed from tribal lands and held at institutions in the United States, such as museums and universities.
In 1990, a landmark human rights federal law, the Native American Graves Protection and Repatriation Act (NAGPRA), spurred change in the professional standards and practice of biological anthropology and archaeology. NAGPRA established a legal avenue to provide protection for and repatriation of Native American remains, cultural items, and sacred objects removed from Federal or tribal lands to Native American lineal descendants, Indian tribes, and Native Hawaiian organizations. Human remains and associated artifacts, curated in museum collections and federally funded institutions, are subject to three primary provisions outlined by the NAGPRA statute: (1) protection for Native graves on federal and private land; (2) recognition of tribal authority on such lands; and (3) the requirement that all Native skeletal remains and associated artifacts be inventoried and culturally affiliated groups be consulted concerning decisions related to ownership and final disposition (Rose, Green, and Green 1996). NAGPRA legislation was enacted to ensure ethical consideration and treatment of Native remains and to improve dialogue between scientists and Native groups.
- For more information about NAGPRA, visit the Bureau of Reclamation NAGPRA website
- To read the text of the law, visit the US Congress NAGPRA law website.
- For further discussion of NAGPRA history, please see TRACES: An Open Invitation to Archaeology open textbook website
Becoming a Forensic Anthropologist
What does it take to be a forensic anthropologist? Forensic anthropologists are first and foremost anthropologists. While many forensic anthropologists have an undergraduate degree in anthropology, they may also major in biology, criminal justice, pre-law, pre-med, and many other related fields. Practicing forensic anthropologists typically have an advanced degree, either a Master’s or Doctoral degree in Anthropology. Additional training and experience in archaeology, the medico-legal system, rules of evidence, and expert witness testimony are also common. Practicing forensic anthropologists are also encouraged to be board-certified through the American Board of Forensic Anthropology (ABFA). Learn more about the field and educational opportunities on the ABFA website: https://www.theabfa.org/coursework.
Review Questions
- What is forensic anthropology? What are the seven primary steps involved in a skeletal analysis?
- What are the major components of a biological profile? Why are forensic anthropologists often-tasked with creating biological profiles for unknown individuals?
- What are the four major types of skeletal trauma?
- What is taphonomy, and why is an understanding of taphonomy often critical in forensic anthropology analyses?
- What are some of the ethical considerations faced by forensic anthropologists?
About the Authors
Ashley Kendell, Ph.D.
California State University, Chico, akendell@csuchico.edu
Dr. Ashley Kendell is currently an associate professor and forensic anthropologist at Chico State. Prior to beginning her position at Chico State, she was a visiting professor at the University of Montana and the forensic anthropologist for the state of Montana. Dr. Kendell obtained her doctorate from Michigan State University, and her research interests include skeletal trauma analysis and digitization and curation methods for digital osteological data. She is also a Registry Diplomate of the American Board of Medicolegal Death Investigators. Throughout her doctoral program, she worked as a medicolegal death investigator for the greater Lansing, Michigan, area and was involved in the investigation of over 200 forensic cases.
Alex Perrone, M.A., M.S.N, R.N., P.H.N.
Butte Community College, perroneal@butte.edu
Alex Perrone is a lecturer in anthropology at Butte Community College. She is also a Registered Nurse and a certified Public Health Nurse. She is a former Supervisor of the Human Identification Laboratory in the Department of Anthropology at California State University, Chico. Her research interests include bioarchaeology, paleopathology, forensic anthropology, skeletal biology, California prehistory, and public health. She has worked on bioarchaeological and archaeological projects in Antigua, California, Hawaii, Greece, and the UK, and was an archaeological technician for the USDA Forest Service. She assisted with training courses for local and federal law enforcement agencies and assisted law enforcement agencies with the recovery and analysis of human remains.
Colleen Milligan, Ph.D.
California State University, Chico, cfmilligan@csuchico.edu
Dr. Colleen Milligan is a biological and forensic anthropologist with research interests in bioarchaeology, skeletal biology, and forensic anthropology. She has been a Fellow with the Department of Homeland Security and has assisted in forensic anthropology casework and recoveries in the State of Michigan and California. She has also assisted in community outreach programs in forensic anthropology and forensic science, as well as recovery training courses for local, state, and federal law enforcement officers. She is a certified instructor through Peace Officers Standards and Training (POST). Dr. Milligan serves as the current co-director of the Chico State Human Identification Laboratory.
For Further Exploration
The American Board of Forensic Anthropology (ABFA)
The American Academy of Forensic Sciences (AAFS)
The Organization of Scientific Area Committees for Forensic Science (OSAC)
References
Adams, Bradley J., and Lyle W. Konigsberg, eds. 2008. Recovery, Analysis, and Identification of Commingled Remains. Totowa, NJ: Humana Press.
Beatrice, Jared S., and Angela Soler. 2016. “Skeletal Indicators of Stress: A Component of the Biocultural Profile of Undocumented Migrants in Southern Arizona.” Journal of Forensic Sciences 61 (5): 1164–1172.
Berg, Gregory E. 2017. “Sex Estimation of Unknown Human Skeletal Remains.” In Forensic Anthropology: A Comprehensive Introduction, Second Edition, edited by Natalie R. Langley and MariaTeresa A. Tersigni-Tarrant, 143–159. Boca Raton, FL: CRC Press.
Bethard, Jonathan D., and Elizabeth A. DiGangi. 2020. “Letter to the Editor—Moving Beyond a Lost Cause: Forensic Anthropology and Ancestry Estimates in the United States.” Journal of Forensic Sciences 65 (5): 1791–1792.
Birkby, Walter H., Todd W. Fenton, and Bruce E. Anderson. 2008. “Identifying Southwest Hispanics Using Nonmetric Traits and the Cultural Profile.” Journal of Forensic Sciences 53 (1): 29–33.
Blatt, Samantha, Amy Michael, and Lisa Bright. Forthcoming. “Bioarchaeology: Interpreting Human Behavior from Skeletal Remains.” In TRACES: An Open Invitation to Archaeology. https://textbooks.whatcom.edu/tracesarchaeology/.
Brooks, S., and J. M. Suchey. 1990. “Skeletal Age Determination Based on the Os Pubis: A Comparison of the Acsádi-Nemeskéri and Suchey-Brooks Methods.” Human Evolution 5 (3): 227–238.
Buchanan, Shelby. 2014. “Bone Modification in Male to Female Transgender Surgeries: Considerations for the Forensic Anthropologist.” MA thesis, Department of Geography and Anthropology, Louisiana State University, Baton Rouge.
Cunningham, Craig, Louise Scheuer, and Sue Black. 2016. Developmental Juvenile Osteology, Second Edition. London: Elsevier Academic Press.
Galloway, Alison, ed. 1999. Broken Bones: Anthropological Analysis of Blunt Force Trauma. Springfield, IL: Charles C. Thomas Publisher, LTD.
Hefner, Joseph T., and Kandus C. Linde. 2018. Atlas of Human Cranial Macromorphoscopic Traits. San Diego: Academic Press.
İşcan, M. Y., S. R. Loth, and R. K. Wright. 1984. “Age Estimation from the Rib by Phase Analysis: White Males.” Journal of Forensic Sciences 29 (4): 1094–1104.
İşcan, M. Y., S. R. Loth, and R. K. Wright. 1985. “Age Estimation from the Rib by Phase Analysis: White Females.” Journal of Forensic Sciences 30 (3): 853–863.Katz, Darryl, and Judy Myers Suchey. 1986. “Age Determination of the Male Os Pubis.” American Journal of Physical Anthropology 69 (4): 427–435.
Komar, Debra A., and Jane E. Buikstra. 2008. Forensic Anthropology: Contemporary Theory and Practice. New York: Oxford University Press.
Langley, Natalie R., Alice F. Gooding, and MariaTeresa Tersigni-Tarrant. 2017. “Age Estimation Methods.” In Forensic Anthropology: A Comprehensive Introduction, Second Edition, edited by Natalie R. Langley and MariaTeresa A. Tersigni-Tarrant, 175–191. Boca Raton, FL: CRC Press.
Lovell, Nancy C. 1997. “Trauma Analysis in Paleopathology.” Yearbook of Physical Anthropology 104 (S25): 139–170.
Native American Graves Protection and Repatriation Act (NAGPRA) 1990 (25 U.S. Code 3001 et seq.)
NIST (National Institute of Standards and Technology). N.d. “The Organization of Scientific Area Committees for Forensic Science.” Accessed April 18, 2023. https://www.nist.gov/topics/organization-scientific-area-committees-forensic-science.
Ousley, Stephen. 1995. “Should We Estimate Biological or Forensic Stature?” Journal of Forensic Sciences 40(5): 768–773.
Phenice, T. W. 1969. “A Newly Developed Visual Method of Sexing the Os Pubis.” American Journal of Physical Anthropology 30 (2): 297–302.
Rose, Jerome C., Thomas J. Green, and Victoria D. Green. 1996. “NAGPRA Is Forever: Osteology and the Repatriation of Skeletons.” Annual Review of Anthropology 25: 81–103.
Schaefer, Maureen, Sue Black, and Louise Scheuer. Juvenile Osteology: A Laboratory and Field Manual. 2009. San Diego: Elsevier Academic Press.
Schall, Jenna L., Tracy L. Rogers, and Jordan D. Deschamps-Braly. 2020. “Breaking the Binary: The Identification of Trans-women in Forensic Anthropology.” Forensic Science International 309: 110220. https://doi.org/10.1016/j.forsciint.2020.110220.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2010a. “Personal Identification.” Last modified June 3, 2010. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_personal_identification.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2010b. “Sex Assessment.” Last modified June 3, 2010. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_sex_assessment.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2011. “Trauma Analysis.” Last modified May 27, 2011. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_trauma.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2012. “Stature Estimation.” Last modified August 2, 2012. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_stature_estimation.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2013. “Age Estimation.” Last modified January 22, 2013. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_age_estimation.pdf.
Soler, Angela, and Jared S. Beatrice. 2018. “Expanding the Role of Forensic Anthropology in Humanitarian Crisis: An Example from the USA-Mexico Border. In Sociopolitics of Migrant Death and Repatriation: Perspectives from Forensic Science, edited by Krista E. Latham and Alyson J. O’Daniel, 115–128. New York: Springer.
Soler, Angela, Robin Reineke, Jared Beatrice, and Bruce E. Anderson. 2019. “Etched in Bone: Embodied Suffering in the Remains of Undocumented Migrants.” In The Border and Its Bodies: The Embodiment of Risk along the U.S.-México Line, edited by Thomas E. Sheridan and Randall H. McGuire, 173–207. Tucson: University of Arizona Press.
Stull, Kyra E., Eric J. Bartelink, Alexandra R. Klales, Gregory E. Berg, Michael W. Kenyhercz, Erica N. L’Abbé, Matthew C. Go, et al.. 2021. “Commentary on: Bethard JD, DiGangi EA. Letter to the Editor—Moving Beyond a Lost Cause: Forensic Anthropology and Ancestry Estimates in the United States. J Forensic Sci. 2020;65(5):1791–2. doi: 10.1111/1556-4029.14513.” Journal of Forensic Sciences 66 (1): 417–420.
Tallman, Sean D., Caroline D. Kincer, and Eric D. Plemons. 2022. “Centering Transgender Individuals in Forensic Anthropology and Expanding Binary Sex Estimation in Casework and Research.” Special issue, “Diversity and Inclusion,” Forensic Anthropology 5 (2): 161–180.
Tersigni-Tarrant, MariaTeresa A., and Natalie R. Langley. 2017. “Human Osteology.” In Forensic Anthropology: A Comprehensive Introduction, Second Edition, edited by Natalie R. Langley and MariaTeresa A. Tersigni-Tarrant, 81–109. Boca Raton, FL: CRC Press.
Ubelaker, Douglas H. 2018. “A History of Forensic Anthropology.” Special issue, “Centennial Anniversary Issue of AJPA,” American Journal of Physical Anthropology 165 (4): 915–923.
White, Tim D., and Pieter A. Folkens. 2005. The Human Bone Manual. Burlington, MA: Elsevier Academic Press.
Winburn, Allysha P., and Bridget Algee-Hewitt. 2021. “Evaluating Population Affinity Estimates in Forensic Anthropology: Insights from the Forensic Anthropology Database for Assessing Methods Accuracy (FADAMA).” Journal of Forensic Sciences 66 (4): 1210–1219.
Winburn, Allysha Powanda, Sarah Kiley Schoff, and Michael W. Warren. 2016. “Assemblages of the Dead: Interpreting the Biocultural and Taphonomic Signature of Afro- Cuban Palo Practice in Florida.” Journal of African Diaspora Archaeology and Heritage 5 (1): 1–37.
Joylin Namie, Ph.D., Truckee Meadows Community College
Learning Objectives
- Identify and describe the major developments in scientific thought that led to the discovery of evolutionary processes.
- Explain how natural selection works and results in evolutionary change over time.
- Explain what is meant by the “Modern Synthesis” and its impacts on evolutionary thought.
- Discuss the teaching of human evolution in the U.S. and abroad.
The Beginnings of Evolutionary Thinking
Throughout our evolutionary history, humans have developed an understanding of the natural world as they interacted with and extracted resources from it.To survive, our earliest ancestors possessed an understanding of the physical environment, including weather patterns, animal behavior, edible and medicinal plants, locations of water, and seasonal cycles. Many ancient cultures, including those of the Americas (Dunbar-Ortiz 2014), Mesopotamia, and Egypt, left writings, hieroglyphics, and stories passed down through oral tradition detailing their understanding of the natural environment, human and zoological anatomy, botany, and medical practices (Moore 1993).
There are also over 2,000 years of organized thought and writing regarding evolution, including contributions from Greek, Roman, and Islamic scholars. Three examples of note are included here. The Greek philosopher Aristotle (384–322 BCE) studied the natural world, publishing several volumes on animals based on systematic observations, rather than attributing what he observed to divine intervention, as his contemporaries were doing (Figure 2.1). Aristotle’s system for the biological classification of nearly 500 species of animals was based on his own observations and dissections, interviews with specialists such as beekeepers and fishermen, and accounts of travelers. His nine book History of Animals, published in the 4th century BC (n.d.), was one of the first zoological taxonomies ever created. Aristotle’s primary contribution to the classification of biological species was to recognize that natural groups are based on structure, physiology, mode of reproduction, and behavior (Moore 1993, 39).
Aristotle’s History of Animals also placed animals in a hierarchy, ranking animals above plants due to what he claimed were their abilities to sense the world around them and to move. He also graded animals according to their modes of reproduction. Those giving birth to live young were placed above those who laid eggs. Warm-blooded animals ranked above invertebrates. This concept of “higher” and “lower” organisms was expanded upon by scholars in the Medieval period to form the Scala Naturae (Latin for “ladder of being”). This “Great Chain of Being,” depicting a hierarchy of beings with God at the top and minerals at the bottom (Figure 2.2), was thought by medieval Christians to have been decreed by God; in this Great Chain, humans were placed closer to God than other species. Aristotle’s works were rediscovered by Islamic scholars in the ninth century and translated into Arabic, Syriac, Persian, and later into Latin, becoming part of university curriculum in 13th-century Europe (Lindberg 1992), allowing Aristotle’s works and ideas to influence other thinkers for 2,000 years.
Science also owes a debt to many Arabic scholars. One such Islamic scholar and writer, who built upon the ideas of Aristotle, was Abū ʿUthman ʿAmr ibn Baḥr al-Kinānī al-Baṣrī / al-Jāḥiẓ, known as Al-Jahiz (776–868 CE), who authored over 200 books (El-Zaher 2018; Figure 2.3). His most well-known work was the seven-volume Kitab al-Hayawan or Book of Animals, in which he described over 350 species in zoological detail. Importantly, Al-Jahiz introduced the idea and mechanisms of biological evolution 1,000 years before Darwin proposed the concept of natural selection in 1859 (Love 2020). For instance, Al-Jahiz wrote about the struggle for existence, the transformation of species over time, and environmental factors that influence the process, all ideas that were later espoused by western European scientists in the 19th century. Al-Jahiz wrote:
Animals engage in a struggle for existing, and for resources, to avoid being eaten, and to breed. Environmental factors influence organisms to develop new characteristics to ensure survival, thus transforming them into new species. Animals that survive to breed can pass on their successful characteristics to their offspring. [Masood 2009]
Another important early Islamic scientist is Ibn al-Haytham (965–1040), a 10th-century Islamic scholar who contributed a great deal to our understanding of optics and how human vision works (Figure 2.4; Lindberg 1992, 177–180). Born in what is now Iraq, al-Haytham was a scholar of many disciplines, including mathematics, physics, mechanics, astronomy, philosophy, and medicine. He authored some 200 books in his lifetime and was an expert on Aristotle’s natural philosophy, logic, and metaphysics. In relation to evolution, al-Haytham’s methodology of investigation—specifically, using experiments to verify theory—is similar to what later became known as the modern scientific method. He is most famous for discovering the laws of reflection and refraction over 1,000 years ago and inventing the camera obscura, which was incredibly important for the eventual development of photography. His work is credited for its influence on astronomy, mathematics, and optics, inspiring Galileo, Johannes Kepler, and Sir Isaac Newton (Tasci 2020). As an inspirational scientific figure, al-Haytham was celebrated in 2016 by UNESCO as a trailblazer and the founder of modern optics (Figure 2.5). An International Year of Light was named in his honor and a scholarly conference on his contributions was held to coincide with the 1,000th anniversary of the publication of his Kitāb al-Manāẓir (Book of Optics; UNESCO.org 2015).
The writings of these Islamic scholars as well as similar scientific texts from other cultures were unknown to or unacknowledged by Western scientists until recently. Fortunately, many science teachers are now incorporating this content into their classes (Love 2020).
Western European Evolutionary Thought
Although there have been many different scientific traditions throughout world history, a new global discourse around science emerged in Western Europe in the 19th century. Europeans pointed to the continuing expansion of their colonial power, as well as their military and technological success, as evidence of the efficacy of Western science, which came to dominate on a global scale (Elshakry 2010). The movement toward a global science centered in Western Europe began with formulation of the Scientific Method.
The Scientific Method was first codified by Francis Bacon (1561–1626), an English politician who was likely influenced by the methods of inquiry established by Ibn al-Haytham centuries prior (Tbakhi and Amr 2007). Bacon has been called the founder of empiricism for proposing a system for weighing the truthfulness of knowledge based solely on inductive reasoning and careful observations of natural phenomena. Ironically, he died as a result of trying to scientifically observe the effects of cold on the putrefaction of meat. On a journey out of London, he purchased a chicken and stuffed it with snow for observation, catching a chill in the process. One week later, he died of bronchitis (Urbach, Quinton, and Lea 2023).
The second important development with regard to evolution was the concept of a species. John Ray (1627–1705), an English parson and naturalist, was the first person to publish a biological definition of species in his Historia Plantarum (History of Plants), a three volume work published in 1686, 1688, and 1704. Ray defined a species as a group of morphologically similar organisms arising from a common ancestor. However, we now define a species as a group of similar organisms capable of producing fertile offspring. In keeping with the scientific method, Ray classified plants according to similarities and differences that emerged from observation. He claimed that any seed from the same plant was the same species, even if it had slightly different traits.
The modern period of biological classification began with the work of Carl von Linne (“Carolus Linnaeus”) (1707–1778), a Swedish scientist who laid the foundations for the modern scheme of taxonomy used today. He established the system of binomial nomenclature, in which a species of animal or plant receives a name consisting of two terms: the first term identifies the genus to which it belongs and the second term identifies the species. His original Systema Naturae, published in 1736, went through several editions. By the tenth edition in 1758, mammals incorporated primates, including apes and humans, and the term Homo sapiens was introduced to signify the latter (Paterlini 2007).
Georges-Louis Leclerc, Comte de Buffon (1707–1788), was a prominent French naturalist whose work influenced prominent scientists in the second half of the 18th century. Buffon's idea that species change over time became a cornerstone of modern evolutionary theory. His technique of comparing similar structures across different species, called comparative anatomy, is still in use today in the study of evolution. He published 36 volumes of Histoire Naturelle during his lifetime and heavily influenced two prominent French thinkers who were to have significant impacts on our understanding of evolution, Georges Cuvier and Jean-Baptiste Lamarck.
Georges Cuvier (1769–1832) was a paleontologist and comparative anatomist (Figure 2.6). One of his first major contributions to the field of evolution was proof that some species had become extinct through detailed and comprehensive analyses of large fossil quadrupeds (Moore 1993, 111). The idea of extinction was not new, but it was challenging to demonstrate if a fossil species was truly extinct or still had living relatives elsewhere. It was also challenging in that it ran counter to religious beliefs of the time. The Bible’s Book of Genesis was interpreted as saying that all species had been created by God in the seven days it took to create the world and that all created species have survived to this day. Extinction was interpreted as implying imperfection, suggesting God’s work was flawed. Also, given that the Earth was calculated to have been created in 4004 B.C.E., based on biblical genealogies, there would not have been enough time for species to disappear (Moore 1993, 112).
Cuvier was so knowledgeable in this field that he became famous for his ability to reconstruct what an extinct animal looked like from fragmentary remains. He demonstrated that fossil mammoths differed from similar living creatures, such as elephants. His many examples of fossils telling the stories of animals that lived and then disappeared were taken as incontrovertible proof of extinctions (PBS 2001). Where Cuvier went awry was his hypothesis of how extinction worked and its causes. As part of his study of comparative anatomy, Cuvier made observations of stratified layers of rock, or sediment, each containing different species. From this, he drew conclusions that species were “fixed” and did not evolve, but then went extinct, and that different assemblages of fossils occurred at different times in the past, as evidenced by the sedimentary layers (Moore 1993, 118). Cuvier explained this through a theory of catastrophism, which stated that successive catastrophic deluges (akin to Biblical floods) swept over parts of the Earth periodically, exterminating all life. When the waters receded from a particular region, lifeforms from unaffected regions would repopulate the areas that were destroyed, giving rise to a new layer of species that looked different from the layer below it. This theory implied that species were fixed in place and did not evolve and that the Earth was young. In fact, Cuvier postulated that the last catastrophe was a deluge he believed occurred five to six thousand years ago, paving the way for the advent of humans (Moore 1993, 118). Cuvier’s catastrophism became part of an ongoing and vociferous debate between two schools of geology. The catastrophists believed the present state of the earth was the consequence of a series of violent catastrophes of short duration, while the uniformitarians thought it was the result of slow acting geological forces that continue to shape the earth.
James Hutton (1726–1797) was one prominent proponent of uniformitarianism. Based on evidence he found at sites in his native Scotland, Hutton argued that the Earth was much older than previously thought. Examining the geology of Siccar Point, a cliff site on the eastern coast of Scotland (Figure 2.7), Hutton concluded that the intersection of the vertical and horizontal rocks represented a gap in time of many millions of years, during which the lower rocks had been deformed and eroded before the upper layers were deposited on top. From this, Hutton argued sediments are deposited primarily in the oceans, where they become strata, or layers of sedimentary rock. Volcanic action uplifts these strata to form mountains, which are then subject to erosion from rain, rivers, and wind, returning sediment to the oceans (Moore 1993, 121). Hutton’s Theory of the Earth (1788) demanded vast periods of time (known as “deep time”) for such slow-working forces to shape the earth. At the time, he was heavily criticized for this view, as it contradicted the biblical version of the history of creation.
Another Scotsman, who was to become a highly influential geologist and a close friend of Darwin, was Charles Lyell (1797–1875). Lyell was originally a lawyer who began his studies of Geology at Oxford under the tutelage of catastrophist William Buckland, from whom he diverged when Buckland tried to find physical evidence of Noah’s flood from the Christian Bible. Lyell was instead intent on establishing geology as a science based on observation. Building upon Hutton’s ideas (published 50 years earlier), Lyell traveled throughout Europe, documenting evidence of uniformitarianism. During his travels, he cataloged evidence of sea level rise and fall and of volcanoes positioned atop much older rocks. He also found evidence of valleys formed through erosion, mountains resulting from earthquakes, and volcanic eruptions that had been witnessed or documented in the past (University of California Berkeley Museum of Paleontology n.d.). Lyell also espoused the principle that “rocks and strata (layers of rock) increase in age the further down they are in a geological sequence. Barring obvious upheavals or other evidence of disturbance, the same principle must apply to any fossils contained within the rock. The lower down in a sequence of rocks a fossil is, the older it is likely to be (Wood 2005, 12).”
Lyell published the first edition of his three-volume Principles of Geology in 1830–1833 (Figure 2.8). It established geology as a science, underwent constant revisions as new scientific evidence was discovered, and was published in 12 editions during Lyell’s lifetime. In it, he espoused the key concept of uniformitarianism—that "the present is the key to the past.” What this meant was that geological remains from the distant past can be explained by reference to geological processes now in operation and directly observable.
Jean-Baptiste Lamarck (1744–1829) was the first Western scientist to propose a mechanism explaining why and how traits changed in species over time, as well as to recognize the importance of the physical environment in acting on and shaping physical characteristics. Lamarck’s view of how and why species changed through time, known as the “Theory of Inheritance of Acquired Characteristics,” was first presented in the introductory lecture to students in his invertebrate zoology class at the Museum of Natural History in Paris in 1802 (Burkhardt 2013). It was based on the idea that as animals adapted to their environments through the use and disuse of characteristics, their adaptations were passed on to their offspring through reproduction (Figure 2.9). Lamarck was right about the environment having an influence on characteristics of species, as well as about variations being passed on through reproduction. He simply had the mechanism wrong.
Lamarck’s theory involved a three-step process. Step one involves an animal experiencing a radical change in its environment. Step two is the animal (either individual or species) responding with a new kind of behavior. Step three is how the behavioral change results in morphological (meaning physical) changes to the animal that are successfully passed on to subsequent generations (Ward 2018, 8). Lamarck’s most famous example was the proposition that giraffes actively stretched their necks to reach leaves on tall trees to eat. Over their lifetimes, the continuation of this habit resulted in gradual lengthening of the neck. These longer necks were then passed on to their offspring. Lamarck's theory was disproved when evolutionary biologist August Weismann published the results of an experiment involving mice (Figure 2.10). Weismann amputated the tails of 68 mice and then successively bred five generations of them, removing the tails of all offspring in each generation, eventually producing 901 mice, all of whom had perfectly healthy long tails in spite of having parents whose tails were missing (Weismann 1889).
How giraffes actually ended up with long necks is a different story. In an environment where the food supply is higher off the ground, and perhaps less available to competing species, giraffes who happened to have slightly longer necks (due to random individual variation and genetic mutation) would be more likely to survive. These giraffes would then be able to reproduce, passing along the slight variation in neck length that would allow their offspring to do the same. Over time, individuals with longer necks would be overrepresented in the population, and neck lengths overall would increase among giraffes. Unfortunately, Lamarck’s ideas challenged the scientific establishment of the time and were rejected. He was discredited and harassed “to the point of loss of money, reputation, and then health” (Ward 2018, 9).
The final piece in the evolutionary puzzle leading up to the theory of natural selection was put forth by Thomas Malthus (1766–1834), who published An Essay on Population in 1798. Malthus lived in England during the time of the Industrial Revolution. It was a time of great poverty and misery when many people migrated from the countryside to squalid, disease-ridden cities to work extremely long hours in dangerous conditions in factories, coal mines, and other industrial workplaces. Birth rates were high and starvation and disease were rampant. Malthus struggled to explain why. His answer was basically the idea of carrying capacity, an ecological concept still in use today. Malthus suggested the rate of population growth exceeded the rate of increase of the human food supply. In other words, people were outgrowing the available food crops. He also suggested that populations of animals and plants were naturally constrained by the food supply, resulting in reductions in population in times of scarcity, “restraining them within the prescribed bounds” (Moore 1993, 147). But, despite significant challenges, some individuals always survived. This was the key to later understandings of evolutionary change in species over time.
The Journey to Natural Selection
In Western European thought, the individual most closely associated with evolution is Charles Darwin (1809–1882). However, as one can see from the individuals and ideas presented in the prior section, he was not the first person to explore evolution and how it might work. In fact, Darwin built upon and synthesized many of the ideas—from geology to biology, ecology, and economy—discussed above. He was simply in the right place at the right time. If he had not worked out his ideas when he did, someone else would have. As a matter of fact, as noted below, someone else did, forcing Darwin to publicly reveal his theory.
Darwin’s journey to the discovery of natural selection began during a childhood spent being curious, experimenting, and collecting natural specimens. When Darwin was 12 years old, his nickname was “Gas” because of the foul-smelling chemistry experiments he and his older brother, Erasmus, performed late into the evenings in their makeshift laboratory in the garden of their parent’s home (Costa 2017). Darwin was also a lifelong collector of biological and geological specimens, most famously beetles, at times going to great lengths in pursuit of a new specimen, as one of his personal letters relates,
I will give a proof of my zeal: one day, on tearing off some old bark, I saw two rare beetles, and seized one in each hand; then I saw a third and new kind, which I could not bear to lose, so that I popped the one which I held in my right hand into my mouth. Alas! it ejected some intensely acrid fluid, which burnt my tongue so that I was forced to spit the beetle out, which was lost, as was the third one. [Darwin 2001 (1897), 50].
Darwin continued his observations and experiments during his formal education, culminating in his graduation from Cambridge in 1831, at which point he was invited to become a gentleman naturalist for a British Royal Navy surveying mission of the globe aboard the H.M.S. Beagle. It is worth noting that Darwin was only 22 years old and the captain’s third choice for the position (Costa 2017), but he proved extremely curious and methodical. The mission departed in December of 1831 and returned five years later (Figure 2.11). During this time, Darwin produced copious notebooks, observations, drawings, and reflections on the natural phenomena he encountered and the experiments he performed.
Discussing all of Darwin’s work aboard the Beagle is beyond the scope of this chapter, but his primary interests were in cataloging new varieties of plant and animal life and examining the geology of the places the ship made landfall. Part of Darwin’s success with regard to both ventures was due to his extreme seasickness, which began before the ship even left Plymouth Harbor. It never let up, encouraging Darwin to go ashore at every available opportunity. “In fact, of the nearly five years of the voyage, Darwin was actually on board the ship for just a year and a half altogether” (Costa 2017, 18).
During the voyage, the young Darwin tried to make sense of what he saw through the lens of the scientific paradigms he held when he left England, but he continually made observations that challenged these paradigms. For example, while the Beagle crewmen were charting the coast of Argentina, Darwin conducted fieldwork on land. There he observed species that were new to him, like armadillos. He also collected fossils, including those of extinct armadillos. Meaning, he had found both extant and extinct members of the same species in the same geographic location, which challenged the theory of catastrophism put forth by Cuvier, who argued that each variant of an animal, living or extinct, was its own distinct species (Moore 1993, 144). Darwin also observed geographic variation in the same species all along the east coast of South America, from Brazil to the southern tip of Argentina. He noted that some species were found in multiple localities and differed from place to place. Those living closer to each other exhibited only slight variations, while those living further apart might be cataloged as entirely different species if one did not know better.
He made similar observations in the Galapagos Islands located off the northwest coast of Ecuador, with regard to giant tortoises and finches (Figure 2.12). A local resident of the islands explained to Darwin that each island had its own variety of tortoise and that locals could discern which island a tortoise came from simply by looking at it. Darwin noted other such examples in both plants and animals, meaning geographic variation was occurring on separate, neighboring islands.
Prevailing views of time argued that variations in living species, and even the fossil armadillos and the living armadillos, were the result of separate creation events. According to this view, each variation, no matter how slight, was a different species. Challenging these ideas would mean challenging not only catastrophism, but the Fixity of Species and other well-accepted ideas of the time. Darwin was aware that he was a young, unestablished naturalist. He was also aware of the ruin that befell Lamarck when his theories were rejected. Lastly, Lyell, who was a good friend of Darwin’s, rejected evolution altogether. It is no wonder that Darwin published a great deal about the geological and fossil data he collected when he returned from the voyage, but not his early hypotheses about evolution.
Upon Darwin’s return to England, it took another twenty years of data collection and experimentation before he was ready to share his conclusions about evolution. Much of this work was conducted at Down House, his home of forty years, where he performed all sorts of experiments that laid the groundwork for his ideas about evolution. Darwin’s home was his laboratory, and he engaged the help of his children, neighbors, friends, and servants in collecting, dissecting, and experimenting. At one point in the 1850s, sheets of moistened paper covered with frogs eggs lined the hallways of the house, while flocks of sixteen different pigeon breeds cooed outside, glass jars filled with salt water and floating seeds filled the cellar, and the smell of dissected pigeon skeletons pervaded the air inside the house. There were also ongoing experiments in the yard, including piles of dissected flowers, beekeeping, and fenced-off plots of land where seedlings were under study. Darwin was a keen experimental scientist, observer, and a prolific writer and presenter of scientific papers. He regarded his work as “one long argument” that never really ended. In fact, Darwin published ten books after On the Origin of Species, addressing such far-ranging topics as animal behavior, orchids, and domestication, among others (Costa 2017).
Darwin may not have published Origins in 1859 had it not been for receiving a paper in June of 1858 from Alfred Russel Wallace, an English naturalist working in Malaysia, espousing the same ideas. Wallace had sent the paper to Darwin asking if it was worthy of publication and requesting he forward it to Lyell and the English botanist, Joseph Hooker. Darwin wrote to Lyell and Hooker about Wallace’s paper, entitled On the Tendency of Varieties to Depart Indefinitely from the Original Type. In recognition that both Wallace and Darwin had arrived at the same discovery, a “joint” paper composed of four parts (none of them actually coauthored) was read to the Linnaean Society by Lyell, then secretary of the Society, on July 1, 1858, and published on August 20. Darwin published On the Origin of Species 15 months later. (The original composite paper read before the Linnaean society is available to read for free from the Alfred Russell Wallace Website, on the 1858 Darwin-Wallace paper page.)
The Mechanism of Natural Selection
On the subject of natural selection and how it works, let’s hear from Darwin himself from the original publication of On the Origin of Species (1859):
A struggle for existence inevitably follows from the high rate at which all organic beings tend to increase. Every being, which during its natural lifetime produces several eggs or seeds, must suffer destruction during some period of its life, and during some season or occasional year, otherwise, on the principle of geometric increase, its numbers would quickly become so inordinately great that no country could support the product. Hence, as more individuals are produced than can possibly survive, there must in every case be a struggle for existence…It is the doctrine of Malthus applied with manifold force to the whole animal and vegetable kingdoms…There is no exception to the rule that every organic being naturally increases at so high a rate, that if not destroyed, the earth would soon be covered by the progeny of a single pair. …Owing to this struggle for life, any variation, however slight and from whatever cause proceeding, if it be in any degree profitable to an individual of any species…will tend to the preservation of that individual, and will be inherited by its offspring. The offspring, also, will thus have a better chance of surviving, for, of the many individuals of any species which are periodically born, but a small number can survive. I have called this principle, by which slight variation, if useful, is preserved, by the term Natural Selection. [Darwin 1859, 61–62]
Let us take a moment here to explore the mechanism of natural selection in more detail. Before we begin, it is important to recognize that Darwin defined evolution as descent with modification, by which he meant that species share a common ancestor yet change over time, giving rise to new species. Descent with modification refers to the fact that offspring from two parents look different from each of their parents, and from each other, meaning they descend with slight differences (“modifications”). If you have ever observed a litter of puppies or a field of flowers and stopped to examine each individual closely, you have seen that each differs from the next, and none look exactly like their parents. These variations are random, not specific, and may or may not be present in the following generations.
Darwin struggled to explain why some slight differences were preserved over time, while others were not. He turned to what he knew of animal breeding (artificial selection) for an explanation (Richards 1998). Darwin bred different breeds of pigeons at Down House, carefully documenting phenotypic differences across generations, including slight anatomical variations he observed through dissection. He also grew and crossbred species of flowers and dissected those too. Darwin was also very fond of hunting and of hunting dogs. In an early draft of his theory on speciation, he used greyhounds as an example of adaptation and selection, “noting how its every bone and muscle, instinct and habit, were fitted to run down hare (rabbits) (University of Cambridge n.d.).” In each case of plant and animal breeding Darwin observed, he noted that humans were selecting variants in each generation that had characteristics humans desired (i.e., sweetness of fruits, colors of flowers, fur type and color of animals). Breeders then continually bred plants and animals with the desired variants, over and over again. These small changes added up over time to create new species of plants and breeds of animals. Darwin also noted that artificial selection does not necessarily render plants or animals better adapted to their original environments. The characteristics humans desire often result in plants less likely to survive in the wild and animals more likely to suffer from certain behavioral or health problems. One has only to examine high rates of hip dysplasia in several modern breeds of dogs to observe what Darwin was referring to.
From his studies of artificial selection, Darwin drew the conclusion that nature also acts upon variations among members of the same species. Instead of human intervention, the forces of nature, such as heat, cold, predation, disease, and now climate change, determine which offspring, with which variants, survive and reproduce. These individuals then pass down these favorable variants to their own offspring. In this way, nature selects for traits that are beneficial within a particular environment and selects against traits that are disadvantageous within a particular environment. Over many generations, populations of a species become more and more adapted (or, in evolutionary terms, “fit”) for their specific environments. Darwin named this process natural selection.
This theory explained the variations in tortoises Darwin had observed years earlier in the Galapagos Islands (see Figure 2.12). Tortoises who lived on larger islands with lush vegetation to feed on were larger than those on smaller islands. They also had shorter necks and dome-shaped shells as their food was close to the ground. Tortoises on smaller, drier islands fed on cacti, which grew much taller. These tortoises had longer necks, longer front legs, and saddle-shaped shells, which allowed them to successfully stretch to reach the edible cactus pads that grew on the tops of the plants. How did these observable differences in the two tortoise populations emerge? Darwin would argue that, over time, small, random variations in the tortoises were differentially selected for by the distinct natural environments on different islands.
In addition to the biogeographical evidence Darwin offered from his research aboard the Beagle, as well as the evidence he documented from the artificial selection of plants and animals, he also relied, where possible, on fossil evidence. One example, mentioned above, were the fossil findings of extinct armadillos in Argentina in the same locations as living armadillos. Unfortunately, as Darwin himself noted, the geological record was incomplete, most often missing the transitional fossil forms that bridge extinct and living species. That issue has since been resolved with scientific research in geochronology and paleontology, among other fields. It is now well-established that life is far more ancient than was believed in Darwin’s time and that these ancient forms of life were the ancestors to all life on this planet (Kutschera and Niklas 2004).
What Darwin was Missing
Although the theory of evolution by natural selection gained traction in scientific circles in the decades after Darwin’s publication of Origins, he was never able to discover the mechanisms that caused variation within members of the same species or the means by which traits were inherited. This began later in 1892 with the publication of The Germ-Plasm: A Theory of Heredity by August Weismann, the same Weismann of the mouse tail experiment presented earlier in this chapter. In his book, Weissman proposed a theory of germ-plasm, which was a precursor to the later discovery and understanding of DNA. Weismann specialized in cytology, a branch of biology devoted to understanding the function of plant and animal cells. Germ-plasm, he proposed, was a substance in the germ cells (what we would call gametes, or sex cells, today) that carried hereditary information. He predicted that an offspring inherits half of its germ-plasm from each of its parents, and claimed that other cells (e.g. somatic, or body, cells) could not transmit genetic information from parents to offspring. This thereby erased the possibility that acquired traits (which he argued resided in somatic cells) could be inherited (Zou 2015). This contribution to evolutionary theory was an important step toward understanding genetic inheritance, but a connection between genetics and evolution was still lacking.
A series of lectures by a deceased Augustinian monk named Gregor Mendel (1822–1884), originally published in 1865, changed that perspective (Moore 1993, 285). Although Darwin was unknown to Mendel, he began a series of experiments with pea plants shortly after the publication of Darwin’s Origins, aiming to add to evolutionary understandings of heredity. As Mendel bred different generations of pea plants that had differences in seed shape and color, pod shape and color, flower position, and stem length, he documented consistent expression of some variations over others in subsequent generations. He meticulously documented the statistics of each crossing of plants and the percentages of phenotypes that resulted, eventually discovering the concept of dominance and recessiveness of characteristics. He also documented that there is no blending of inherited characteristics. For example, pea pod colors in the offspring of two parent plants, one with yellow pods and one with green, were either yellow or green, not yellowish green. Mendel also discovered that characteristics are inherited and expressed independently of each other, meaning the color of the pea pod was not necessarily expressed in conjunction with the pod being wrinkled or smooth. The recognition of the importance of Mendel’s work began with its rediscovery by Hugo de Vries and Carl Correns, both of whom were working on hypotheses regarding heredity in plants and had arrived at conclusions similar to Mendel’s. Both published papers supporting Mendel’s conclusions in 1900 (Moore 1993). Research into the inheritance of characteristics continued through the next three decades, and by the close of the 1930s, no major scientific questions remained regarding the transmission of heredity through genes, although what genes did and what chemicals they were made of were still under investigation.
The Modern Synthesis refers to the merging of Mendelian genetics with Darwinian evolution that took place between 1930 and 1950. The basic principles of the synthetic theory were influenced by scientists working in many different fields, including genetics, zoology, biology, paleontology, botany, and statistics. Although there were differences of opinion among them, evolution came to be defined as changes in allele frequencies within populations. Genetic mutations, changes in the genetic code that are the original source of variation in every living thing, were believed to be random, the sources of phenotypic variation, and transmitted through sexual reproduction. These assertions were supported by a growing body of field and laboratory research, as well as new work in mathematics in the field of population genetics that defined evolution as numerical changes in gene frequencies within an interbreeding population from one generation to the next (Corning 2020). These changes in gene frequencies were argued to be the result of natural selection, mutation, migration (gene flow), and genetic drift, or random chance. Empirical research and mathematics demonstrated that very small selective forces acting over a relatively long time were able to generate substantial evolutionary change, including speciation (Plutynski 2009). Thus, the Modern Synthesis encompassed both microevolution, which refers to changes in gene frequencies between generations within a population, and macroevolution, longer-term changes in a population that can eventually result in speciation, wherein individuals from different populations are no longer able to successfully interbreed and produce viable offspring.
Genetics and the Origin of Species, published in 1937 by Theodosius Dobzhansky (Figure 2.13), was a cornerstone of the modern synthesis, applying genetics to the study of natural selection in wild populations, appealing to both geneticists and field biologists. Dobzhansky was interested in speciation, particularly in finding out what kept one species distinct from another and how speciation occurred. His research involved fruit flies, the species Drosophila pseudoobscura. At the time he began in the 1920s, biologists assumed all members of the same species had nearly identical genes. Dobzhansky traveled from Canada to Mexico capturing wild members of D.pseudoobscura, discovering that different populations had different combinations of alleles (forms of a gene) that distinguished them from other populations, even though they were all members of the same species. What, then, led to the creation of new species? Dobzhansky realized it was sexual selection. Members of the same species are more likely to live among their own kind and to recognize, and prefer, them as mates. Over time, as a result of random mutations, natural selection in a given environment, and genetic drift, meaning random changes in allele frequencies, members of the same population accumulate mutations distinct to their own gene pool, eventually becoming genetically distinct from other populations. What this means is that they have become a new species.
From these studies, Dobzhansky and others developed the Bateson-Dobzhansky-Muller model, also known as Dobzhansky-Muller model (Figure 2.14). It is a model of the evolution of genetic incompatibility. Combining genetics with natural selection, the model is important in understanding the role of reproductive isolation during speciation and the role of natural selection in bringing it about. Due to sexual selection (mate preference), populations can become reproductively isolated from one another. Eventually, novel mutations may arise and be selected for in one or both populations, rendering members of each genetically incompatible with the other, resulting in two distinct species.
Evolution and Natural Selection Observable Today (Transform into Special Topic/Dig Deeper)
Although this chapter primarily focuses on the past, it is important to remember that natural selection and evolution are still ongoing processes. Climate change, deforestation, urbanization, and other human impacts on the planet are influencing evolution among many contemporary species of plants and animals. One such example occurs among crested anoles (Anolis cristatellus), small lizards of the Caribbean jungle that are increasingly making their home in cities (Figure 2.15).
As urban sprawl continues across the planet, shrinking the availability of wilderness habitat, many wild species have come to make their homes in cities. “Urbanization has dramatically transformed landscapes around the world—changing how animals interact with nature, creating "heat islands" with higher temperatures, and hurting local biodiversity. Yet many organisms survive and even thrive in these urban environments, taking advantage of new habitats created by humans (National Science Foundation 2023). A recent example of lizards in Puerto Rico demonstrates evolution happening quickly in both behavior and genes that has come about as a result of the pressures of urban life (Winchell et al. 2023). Crested anoles, who once lived only in forests, now scurry around towns and cities throughout the Caribbean. As a result of having to sprint across large open spaces, like hot streets and parking lots, they have developed longer limbs. City-living lizards also now sport longer toe pads with special scales that allow them to cling to smooth surfaces, like windows and walls (as well as the plastic patio furniture pictured in Figure 2.15), rather than to the rough surfaces of bark and rock that their forest-living relatives climb. These adaptations enhance their ability to escape predators and survive in cities.
Researchers were curious to see if these changes were the result of genetic changes in urban populations, so they captured 96 male lizards in three Puerto Rican regions and compared their genomes to each other and to forest specimens in each location. They found that members of the three city-living populations were genetically distinct from each other, as well as from forest populations in their respective regions. In total, 33 genes in the urban lizards’ genomes were different from their forest-living counterparts and were linked to urbanization. These changes are estimated to have occurred just within the last 30 to 80 generations, suggesting that selective pressures related to survival in urban environments is strong. As study coauthor Kristen Winchell put it, “We are watching evolution as it is unfolding” (National Public Radio 2023). (If you are interested in hearing more about the study, see “How Lizards Adapt to Urban Living,” an episode of Science Sessions, a free podcast from the Proceedings of the National Academy of Sciences (PNAS 2023) featuring Dr. Winchell and her work.)
Misconceptions About Evolution Through Natural Selection
After many years of teaching about evolution and natural selection, it continues to surprise me how many misconceptions exist about how the process works. If you do a web image search for “human evolution,” the following image is likely to appear (Figure 2.16).
What is wrong with this picture? First, it implies that humans evolved from chimpanzees, which is incorrect. Although, as primates, we share a common ancestor very far back in time, we split from other primates, including our closest relatives, the nonhuman apes, several million years ago. This image also suggests that evolution is gradual and progressive; that it is intentional and directional; and that there is an end to it—a stopping point. As you will be learning, evolution takes place in fits and starts, depending on the physical environment, changes in climate, food supply, predation, reproductive success, and other factors. It is also not intentional, in the sense that there is no predetermined end; in fact, if environmental conditions change, species can evolve in different directions or even go extinct. Evolution also does not necessarily progress in the same direction over time. One example is the eel-like creature Qikiqtania wakei that lived 375 million years ago. It was originally a fish that evolved to walk on land, then evolved to live back in the water. Early tetrapods like Qikiqtania were likely spending more and more time out of the water during this period. The arrangement of bones and joints in their fins was starting to resemble arms and legs, which would have allowed them to prop themselves up in shallow water and survive on mudflats. Qikiqtania’s skeletal morphology, however, suggests that it then evolved from having rudimentary fingers and toes back to fins that allowed them to again swim in open water (Stewart et al. 2022).
There is also the misperception that natural selection can create entirely new anatomical structures out of thin air in response to changes in environmental pressures. For example, when asked if they can think of ways in which modern humans are continuing to evolve biologically, students often postulate that, as a result of climate change, humans might rapidly develop gills, webbed hands and feet, and learn to breathe underwater in response to rising sea levels. Unfortunately, natural selection can only act on slight variations in anatomy that are already present, and we have no rudimentary physiological system for breathing underwater. Given that natural selection can only act upon existing variation, humans have evolved in such a way that many parts of our bodies are prone to injury. Our knees are one example. The anterior cruciate ligament (ACL) in our knees is “vulnerable to tearing in humans because our upright bipedal posture forces it to endure much more strain than it is designed to” (Lents 2018, 23). When our ancestors made the transition from quadrupedalism to upright walking, we shifted from four bent legs to two straight legs, relying more on our bones than our muscles to support our weight. This is functional for normal walking and running in a straight line, but sudden shifts in direction and momentum, combined with the sizes and weights of modern humans, result in tears in an ACL that is simply not strong enough to bear the stress. If evolution had the capability to engineer a knee from scratch, it would look quite different, and any ligaments involved would likely be larger, stronger, and more flexible. For an interesting look at what anatomically modern humans might look like if we had evolved to withstand the stresses our bodies undergo in our present environment, see “This is what the perfect body looks like - according to science,” which was proposed by biological anthropologist Alice Roberts (Harrison 2018).
Another misperception about evolution is that some species are “more evolved” than others. Every species currently alive on the planet today is the result of millennia of natural selection that has rendered current members of that species well-adapted to their respective environments. Humans are no more “evolved” than fruit flies or yeast. What sets us apart are our cultural and technological abilities, which have allowed us to successfully survive in a wide variety of physical environments, many of which are now becoming too hot, too wet, or too dry to sustain human life without a great deal of technological intervention (IPCC 2022).
There is also some confusion about what “fitness” actually means and a failure to grasp that it changes as environmental conditions change. Evolutionary “fitness” is different from physical fitness. “Fitness” in evolutionary terms refers to an individual’s ability to survive and reproduce viable offspring who also survive and reproduce. Evolutionary fitness and reproductive success are highly dependent on specific environmental conditions, which can shift over time, greatly affecting the relative fitness of individuals in a population. Recent research on the impacts of climate change on dragonflies will serve to illustrate the point (Figure 2.17).
Pictured here is a male dragonfly, who, you will notice, has distinctive black markings on its wings. This is due to melanization. Males control breeding, and those with more ornamentation tend to attract more mates and to successfully ward off male competitors. Higher levels of melanization, however, have negative consequences for males in warming climates. The black markings absorb heat, elevating body temperatures, which can cause overheating, reduce male fighting ability, and even lead to death (Moore et al. 2021). Females are not as adversely affected because they spend more time in shaded areas, while males are more often flying in sunlit areas, fending off rivals. However, as highly melanized males become less viable, wing coloration is undergoing selection in males. In other words, what constitutes being “fit” for males has changed, favoring those who have fewer of the black markings and, therefore, are less negatively impacted by warming temperatures. Note that natural selection acts on individuals, “selecting” those who happen to be fit for particular environmental conditions at a particular point in time. Evolution, though, happens at the level of the population. If the climate continues to warm, populations of dragonflies who inhabit warming areas will increasingly exhibit less ornamentation in males.
Lastly, natural selection can only act on characteristics that influence reproductive success. Deleterious traits that have nothing to do with one’s ability to reproduce and successfully rear offspring to reproductive age will continue to be passed on. For example, the author of this chapter is a natural redhead, and redheads are predisposed genetically to a number of conditions that can negatively affect health (Colliss Harvey 2015), but some of these conditions are not diagnosed until later in life. One example is Parkinson’s disease (Chen et al. 2017), which is a degenerative neurological disorder. The average age of diagnosis of Parkinson’s is 60 years of age, meaning redheads may encounter such a diagnosis well past childbearing age, having already passed on the genetic predisposition. Thus, Parkinson’s disease cannot be selected out from the redhead family tree.
Dig Deeper: Teaching Evolution Around the World (Put in Chapter 17 or refer such a subject to chap 17)
Evolution is recognized as a central organizing principle for all scientific disciplines and accepted without controversy among scientists and educated people around the world. The United States has historically been the exception (Lerner 2000). In some parts of the U.S. the teaching of evolution to K-12 students continues to evoke controversy, related to politics and religion. The problem is compounded by the degree of control individual states and local school boards exercise over curriculum in the nation’s public schools (Lerner 2000).
The debate over teaching evolution in schools in the United States first came to a head in 1925 after several states attempted to legislate a ban against its teaching. The state of Tennessee eventually did pass such a ban, and the American Civil Liberties Union (ACLU) offered to defend any science teacher who agreed to break the law. John Scopes, who taught in a small, rural Tennessee school, continued to teach evolution, resulting in the “Scopes Monkey Trial,” one of the most famous media trials in American history (Figure 2.18). The entire nation listened to its broadcast live on the radio and read about it daily in hundreds of newspapers. Scopes was defended by Clarence Darrow, the most famous lawyer in the country at the time; Scopes was eventually convicted, though the conviction was later overturned. The pro-Evolution movement benefited greatly from Darrow’s questioning of those on the anti-Evolution side, whose responses were perceived negatively by well-educated listeners in northern cities. The teaching of evolution was also bolstered by a Supreme Court decision in 1947 overriding states’ rights to make decisions on church-state issues and by the launching of Sputnik, the first Soviet (Russian) satellite, in 1957, making science education a national priority. During the 1960s to 1970s, creationism and intelligent design began to take hold as courts ruled in favor of “academic freedom” (Pew Research Center 2019). Since that time, states, and even local school boards, have pushed for the removal of evolution from science curriculum and textbooks or for teaching evolution on equal footing with concepts such as creationism and intelligent design (Masci 2019).
Partly in response to chronically low scores by students in the United States on international measures of ability in science, math, and reading (Desilver 2017), development of the Next Generation Science Standards (NGSS) for K-12 education began in 2011 and were implemented in 2013 in public schools across the nation (nextgenscience.org). The standards were developed collaboratively by The National Research Council (NRC; a branch of the National Academy of Sciences), the National Science Teachers Association, the American Association for the Advancement of Science, and Achieve, an independent, nonpartisan, nonprofit education-reform organization dedicated to working with states to raise academic standards and graduation rates. Twenty-six states also partnered together in the development of the standards. However, due to a number of issues, including funding and support for teachers, adoption of the standards has varied by state. States were also allowed to alter the curriculum. One of the main issues in several states with the original curriculum was the teaching of evolution (Pew Research Center 2014). One prominent aspect to note is that the NGSS do not specify that human evolution be taught, and high school standards do not require teaching about all of the forces of evolution, only mutation and natural selection.
The situation regarding teaching evolution has changed greatly in recent years. A 2007 survey found that only one in three public high school biology teachers in the United States presented evolution consistently with the recommendations of the nation’s leading scientific authorities, and 13% of these teachers emphasized creationism as a valid scientific alternative to modern evolutionary biology (Plutzer, Branch, and Reid 2020). A repeat of the survey in 2019 demonstrated marked improvement in the amount of time teachers devoted to teaching evolution, as well as more teacher training and preparedness to teach evolution. Such improvements were attributed to the need to meet the Next Generation Science Standards, as well as continuing outreach by the National Science Teaching Association, the National Association of Biology Teachers, and the National Academy of Sciences in producing classroom resources and providing professional development opportunities to advance the inclusion of evolution in the nation’s classrooms.
Public acceptance of evolution has also substantially improved in recent years (Miller et al. 2022). National samples of American adults have been asked at regular intervals since 1985 to agree or disagree with the following statement: “Human beings, as we know them today, developed from earlier species of animals (Miller et al. 2022)” During the last decade, the percentage of U.S. adults agreeing with this statement increased from 40% to 54%—a majority for the first time. This level of acceptance of evolution in the United States is atypically low for a developed nation. In a study of the acceptance of evolution in 34 developed nations in 2005, only Turkey—at 27%—scored lower than the United States (Miller, Scott and Okamoto 2006). There are also distinct differences among members of the U.S. population in terms of acceptance of evolution, with 68% of those ages 18–24, 58% of those with college degrees, and 65% of those who have taken four or more college-level science courses the most accepting of evolution. An increasing number of parents also report changing their unfavorable views of evolution due to helping their children with science homework and science fair projects.
Approaches to teaching evolution vary across the globe, with considerable differences within and between nations. One example is the United Kingdom, home of Darwin and Wallace. There, evolution is not introduced until ages 14–16, which is considered quite late by some educators. And, although evolution is taught in biology classes, it is addressed as a separate topic, rather than integrated into the curriculum as a foundational concept. As in the United States, there is also considerable variation between public and private schools, and between religious and secular institutions, in their treatment of the topic and the inclusion of alternative viewpoints, such as creationism (Harmon 2011). There are similar differences across the European Union, including within different populations in member countries.
There are also differences in the teaching of evolution across many predominantly Muslim nations. Salman Hameed, Professor of Integrated Science and Humanities at Hampshire College and Director of the Center for the Study of Science in Muslim Societies (SSiMS), whose research focuses on the acceptance of evolution among Muslims, has uncovered a great deal of variation among Muslims regarding beliefs about evolution. He points out that there is no central position within Islam regarding evolution, leaving it up to governments, textbook authors, and other entities to decide whether, and how, to address evolution in education. Saudi Arabia, Oman, Algeria, Morocco, and Lebanon all ban the teaching of evolution on religious grounds. Other Islamic nations, including Egypt, Malaysia, Syria, and Turkey, include evolutionary concepts like natural selection in their science curriculum but refrain from discussing human evolution (Asghar, Hameed, and Farahani 2014).
Impeding acceptance of evolution in science classes around the world is the adoption of textbooks from the E.U., the U.K., and the U.S. that include examples that are not culturally relevant to local populations (Harmon 2011). One classic example Hameed cites is that of the peppered moths in England whose predominant color pattern evolved from mostly white to mostly black due to pollutants darkening the tree bark of their habitat during the Industrial Revolution. Historical examples like this have little relevance for 21st-century students who grew up in non-Western countries and know little of England’s history or of the species that live there. It also privileges Western science over local science, to which many individuals in former European colonies and territories object (Jones 2017). Hameed suggests customizing textbooks to include local fossils and species whenever possible (Harmon 2011).
Are We Still Evolving?
After reading this chapter, many students are curious to know if humans are still evolving. The answer is yes. As a species, we continue to respond to selective pressures biologically and culturally. This final section will focus on three contemporary examples of human evolution. Before beginning, let’s review the conditions necessary for natural selection to operate on a trait. First, the trait must be heritable, meaning it is transmitted genetically from generation to generation. There must also be variation of the trait within the population and the trait must influence reproductive success. Three examples of traits that meet these criteria are immunity to the Human Immunodeficiency Virus (HIV), height, and wisdom teeth (Andrews, Kalinowski, and Leonard 2011).
AIDS is a potentially fatal infectious disease caused by HIV, a zoonosis believed to be derived from Simian Immunodeficiency Viruses (SIVs) found in chimpanzees and monkeys and most likely transmitted to humans through the butchering of infected animals (Sharp and Hahn 2011). In total, 40 million people have died from AIDS-related illnesses since the start of the global epidemic in the 1980s. There were 38.4 million people around the world living with AIDS as of 2021, including 1.5 million new cases and 650,000 deaths in that year alone (UNAIDS 2021). A disease causing this level of morbidity and mortality represents a major selective pressure, especially given that infection can occur before birth (Goulder et al. 2016), thereby affecting future reproductive success.
The majority of people in the world are highly susceptible to HIV infection, but some are not. These latter individuals are homozygous for a rare, recessive allele at the CCR5 locus that makes them immune to HIV. Heterozygotes who inherit a single copy of this allele are more resistant to infection and, when infected, the disease takes longer to progress in the event that they are infected. The mechanism by which the allele prevents infection involves a 32-base pair deletion in the DNA sequence of the CCR5 gene, creating a nonfunctioning receptor on the surface of the cell that prevents HIV from infecting the cell. The allele is inherited as a simple Mendelian trait, and there is variation in its prevalence, ranging as high as 14% of the population in northern Europe and Russia (Novembre, Galvani, and Slatkin 2005; see Figure 2.19). What is interesting about the allele’s geographic distribution is that it does not map onto parts of the world with the highest rates of HIV infection (Figure 2.20), suggesting that AIDS was not the original selective pressure favoring this allele (see Figures 2.19 and 2.20).
Given its current geographic distribution, the bubonic plague, which ravaged Europe repeatedly from the 14th to the 19th centuries (Pamuk 2007), was initially proposed as the selective agent. Subsequent research suggests smallpox, which killed up to 400,000 people annually in 18th-century Europe (Hays 2005), was more likely the selective pressure (Novembre, Galvani, and Slatkin 2005). Given the mortality rates for smallpox (Crosby 2003), an allele that conferred immunity was highly advantageous, as it is for those faced with the threat of HIV infection today.
Height is another example of a trait experiencing selective pressure. If you have ever toured a historical site, you have likely hit your head on a doorframe or become claustrophobic trying to squeeze down a narrow hallway under a lower-than-average ceiling. It is not your imagination. Humans have gotten taller in recent centuries. In fact, the average height of people in industrialized nations has increased approximately 10 centimeters (about four inches) in the past 150 years. This increase has been attributed to improvements in nutrition, sanitation, and access to medical care (Hatton 2014). But this is only part of the story.
Height is highly heritable. Studies indicate 80% of variation in height within populations is due to genetics, with 697 different genetic variances identified as having an effect on adult stature (Devuyst 2014). Multiple studies also demonstrate a positive relationship between height and reproductive success for men (Andrews, Kalinowsky, and Leonard 2011). This is primarily due to sexual selection and nonrandom mating, namely women’s preferences for taller men, which may explain why height is one characteristic men often lie about on dating websites (Guadagno, Okdie, and Kruse 2012). Sexual selection also plays out with regard to economic success in Western cultures, with taller men more likely to be in higher-level positions that pay well. Research demonstrates an inch of height is worth an additional $789 per year in salary, meaning a man who is six feet tall will earn on average $5,525 more per year than an identical man who is five foot five purely due to heightism bias (Gladwell 2007). Over the course of a career, this can add up to hundreds of thousands of dollars, likely allowing taller men to attract more potential mates, increasing their reproductive success.
Wisdom teeth are also undergoing evolutionary pressure. Have you or anyone in your family had their wisdom teeth removed? While it can be a painful and expensive process, it is a common experience in Western nations. It begs the question as to why there is no longer room in our mouths for all of our teeth? Biological anthropologist Daniel Lieberman offers several reasons, including that modern humans are growing faster and maturing earlier, which could be leading to impaction if skeletal growth takes place faster than dental growth. He also argues that the soft diets many modern humans consume generate insufficient strain to stimulate enough growth in our jaws to accommodate all of our teeth. Lastly, as the human brain has expanded over the past hundreds of thousands of years, it is taking up more space in the skull, causing the jaw to shrink, leaving no room for third molars (Lieberman 2011).
Conversely, do you know anyone whose wisdom teeth never came in? That is fairly common in some populations, suggesting evolutionary pressure favoring the absence of wisdom teeth has been culturally influenced. The oldest fossil evidence of skulls missing third molars was found in China and is 300,000 to 400,000 years old, suggesting the earliest mutation selecting against the eruption of wisdom teeth arose in Asia (Main 2013). Since that time, jaws have continued to decrease in size to the point they often cannot accommodate third molars, which can become impacted, painful, and even infected, a condition physical anthropologist Alan Main argues might have interfered with the ability to survive and reproduce in ancestral populations (Main 2013). As we have learned, a mutation that positively influences reproductive success—such as being born without the trait to develop wisdom teeth—would likely be selected for over time. Evidence in modern humans suggests that this is the case, with 40% of modern Asians and 45% of Native Alaskans and Greenlanders (populations descended from Asian populations) lacking wisdom teeth. The percentage among those of European descent ranges from 10 to 25% and for African Americans is 11% (Main 2013). Later chapters of this textbook emphasize that directional selection progresses along a particular path until the environment changes and a trait is no longer advantageous. In the case of wisdom teeth, the ability of modern dentistry to preempt impaction through surgery may, in fact, be what is preventing natural selection from doing away with wisdom teeth altogether.
Key Developments in Evolutionary Thought
4th century BCE |
Aristotle (384–322 BCE) |
“Founder of Biology.” Publishes History of Animals, a biological classification system of over 500 animals based on structure, physiology, reproduction, and behavior. Also creates the “Great Chain of Being,” ranking species and placing humans closest to God. |
8th–9th century CE |
Al-Jahiz (776–868 CE) |
Writes seven-volume Book of Animals, which includes animal classifications and food chains. Introduces concept of biological evolution and its mechanisms. |
1011–1021 |
Ibn al-Haythem (965–1040 CE) |
“Father of Modern Optics.” Uses experimental science to catalog how vision works and discovers laws of reflection and refraction. Publishes Book of Optics and invents camera obscura, the foundation for modern photography. |
1620 |
Francis Bacon (1561–1626) |
“Father of Empiricism.” Publishes The Novum Annum, formulating the scientific method based on observation and inductive reasoning. |
1686 |
John Ray (1627–1705) |
First to publish a biological definition of species in History of Plants. |
1749 |
Comte de Buffon (1707–1788) |
Publishes Histoire Naturelle, comparing anatomical structures across species using methods still in use today. Inspires Lamarck and Cuvier. |
1758 |
Carl von Linne (Carolus Linnaeus) (1707–1778) |
Introduces system of binomial nomenclature. Publishes Systema Naturae, the tenth edition of which introduces the designation Homo sapiens for humans. |
1788 |
James Hutton (1726–1797) |
“Father of Geology.” Publishes Theory of the Earth; introduces idea of Deep Time; explains how features of the earth were formed through the actions of rain, wind, rivers, and volcanic eruptions. |
1798 |
Thomas Malthus (1766–1834) |
Economist and “Father of Statistics.” Publishes An Essay on Population; introduces concept of carrying capacity; explains how populations outstrip the food supply, leaving some individuals to die off; inspires Darwin’s idea of “natural selection.” |
1809 |
Jean-Baptiste Lamarck (1744–1829) |
Publishes theory of the Inheritance of acquired characteristics; is the first Western scientist to propose a mechanism explaining how traits change in species over time and to recognize the importance of the physical environment in acting on species and their survival. |
1810 |
Georges Cuvier (1769–1832) |
Paleontologist/comparative anatomist; proved species went extinct; proposed the Theory of Catastrophism. |
1830 |
Charles Lyell (1797–1875) |
Establishes geology as a science. Publishes first edition of The Principles of Geology (1830–33); issuing 12 total editions in his lifetime, each updated according to new scientific data. |
1858 |
Alfred Russel Wallace (1823–1913) |
Sends scientific paper to Darwin titled “On the Tendency of Varieties to Depart Indefinitely from the Original Type,” essentially espousing the concept of natural selection; a reading of the papers by both Wallace and Darwin to the Linnaean Society is conducted by Lyell. |
1859 |
Charles Darwin (1809–1882) |
Publishes On the Origin of Species by Means of Natural Selection (1859). |
1865 |
Gregor Mendel (1822–1884) |
Publishes Experiments in Plant Hybridization (1865), outlining the fundamentals of genetic inheritance. |
1889 |
August Weismann (1834–1914) |
Publishes Essays Upon Heredity (1889), disproving the inheritance of acquired characteristics. Publishes The Germ Plasm (1892), postulating an early idea of inheritance through sexual reproduction. |
1937 |
Theodosius Dobzhansky (1900–1975) |
One of the founders of the Modern Synthesis of biology and genetics. Publishes Genetics and the Origin of Species (1937). Documents a genetic model of speciation through reproductive isolation. |
Review Questions
- Summarize the major scientific developments that led to the formulation of the theory of natural selection.
- Explain how natural selection operates and how it leads to evolution in populations.
- Explain the importance of genetics to an understanding of human evolution.
- Have you observed current examples of evolution taking place where you live? In which species? Which forces of evolution are involved?
Key Terms
Allele: A nonidentical DNA sequence found in the same gene location on a homologous chromosome, or gene copy, that codes for the same trait but produces a different phenotype.
Artificial selection: The identification by humans of desirable traits in plants and animals, and the subsequent steps taken to enhance and perpetuate those traits in future generations.
Binomial nomenclature: A system of classification in which a species of animal or plant receives a name consisting of two terms: the first identifies the genus to which it belongs, and the second identifies the species.
Carrying capacity: The number of living organisms, including animals, crops, and humans, that a geographic area can support without environmental degradation.
Catastrophism: The theory that the Earth’s geology has largely been shaped by sudden, short-lived, violent events, possibly worldwide in scope. Compare to uniformitarianism.
Comparative anatomy: Georges-Louis Leclerc’s technique of comparing similar anatomical structures across different species.
Creationism: The belief that the universe and all living organisms originate from specific acts of divine creation, as in the Biblical account, rather than by natural processes such as evolution.
Empiricism: The idea that all learning and knowledge derives from experience and observation. It became prominent in the 17th and 18th centuries in western Europe due to the rise of experimental science.
Evolution: In a biological sense, this term refers to cumulative inherited change in a population of organisms through time. More specifically, evolution is defined as a change in allele (gene) frequencies from one generation to the next among members of an interbreeding population.
Extant: Still in existence; surviving.
Extinct: Said of a species, family, or other group of animals or plants that has no living members; no longer in existence.
Fixity of Species: The idea that a species, once created, remains unchanged over time.
Gene: A sequence of DNA that provides coding information for the construction of proteins.
Genetic drift: Random changes in allele frequencies within a population from one generation to the next.
Gene flow: The introduction of new genetic material into a population through interbreeding between two distinct populations.
Gene pool: The entire collection of genetic material in a breeding community that can be passed from one generation to the next.
Genotype: The genotype of an organism is its complete set of genetic material—its unique sequence of DNA. Genotype also refers to the alleles or variants an individual carries in a particular gene or genetic location.
Hybrid: Offspring of parents that differ in genetically determined traits.
Intelligent design: A pseudoscientific set of beliefs based on the notion that life on earth is so complex that it cannot be explained by the scientific theory of evolution and therefore must have been designed by a supernatural entity.
Macroevolution: Large and often-complex changes in biological populations, such as species formation.
Microevolution: Changes in the frequency of a gene or allele in an interbreeding population.
Modern synthesis: The mid–20th century merging of Mendelian genetics with Darwinian evolution that resulted in a unified theory of evolution.
Natural selection: The natural process by which the survival and reproductive success of individuals or groups within an interbreeding population that are best adjusted to their environment leads to the perpetuation of genetic qualities best suited to that particular environment at that point in time.
Phenotype: The detectable or visible expression of an organism’s genotype.
Scientific method: A method of procedure that has characterized natural science since the 17th century, consisting of systematic observation, measurement, experimentation, and the formulation, testing, and modification of hypotheses.
Speciation: The process by which new genetically distinct species evolve from the main population, usually through geographic isolation or other barriers to gene flow.
Species: A group of living organisms consisting of similar individuals capable of exchanging genes or interbreeding. The species is the principal natural taxonomic unit, ranking below a genus and denoted by a Latin binomial (e.g., Homo sapiens).
Uniformitarianism: The theory that changes in the earth's crust during geologic history have resulted from the action of continuous and uniform processes—such as wind, precipitation, evaporation, condensation, erosion, and volcanic action—that continue to act in the present. Compare to catastrophism.
About the Author
Joylin Namie, Ph.D.
Truckee Meadows Community College, jnamie@tmcc.edu
Joylin Namie is Professor of Anthropology at Truckee Meadows Community College, where she teaches courses in biological and cultural anthropology. Her current research interest is in (un)sustainable tourism in desert environments, particularly in the country of Jordan and the U.S. state of Nevada. She was awarded a fellowship to Jordan from the Council of American Overseas Research Centers (CAORC) in 2020 to explore this topic, including visiting Petra and other important tourism destinations in Jordan. Dr. Namie’s favorite things in life are teaching, traveling, and spending time with her dog, Charley.
For Further Exploration
Costa, James T. 2017. Darwin’s Backyard: How Small Experiments Led to a Big Theory. New York: W.W. Norton.
Darwin, Charles. 1905. The Voyage of the Beagle. (Originally published in 1839 as Journal and Remarks). [Author’s note: Several editions exist with different publishers, including illustrated editions, paperback editions, and e-books.]
Moore, John A. 1993. Science as a Way of Knowing: The Foundations of Modern Biology. Cambridge, MA: Harvard University Press.
References
Al-Haytham, Ibn. 1011-1021. Kitāb al-Manāẓir (Book of Optics). Cairo, Egypt.
Al-Jahiz. 776–868 CE. Kitab al-Hayawan (Book of Animals).
Andrews, Tessa M., Steven T. Kalinowski, and Mary J. Leonard. 2011. “Are Humans Evolving? A Classroom Discussion to Change Students’ Misconceptions Regarding Natural Selection.” Evolution: Education and Outreach 4 (3): 456–466.
Aristotle. 384-322 BCE. History of Animals.
Asghar, Anila, Salman Hameed, and Najme Kashani Farahani. 2014. “Evolution in Biology Textbooks: A Comparative Analysis of Five Muslim Countries.” Religion & Education 41 (1). Accessed February 12, 2023. https://www.tandfonline.com/doi/abs/10.1080/15507394.2014.855081.
Associated Press. January 10, 2023. “Forest Lizards Have Genetically Morphed To Survive Life In The City, Researchers Say.” National Public Radio (NPR). Retrieved February 19, 2023 from https://www.npr.org/2023/01/10/1148150056/forest-lizards-genetically-morphed-cities.
Burkhardt, Richard W. 2013. “Lamarck, Evolution, and the Inheritance of Acquired Characters.” Genetics 194 (4): 793–805.
Chen, Xiqun, Danielle Feng, Michael A. Schwartzchild, and Xiang Gao. 2017. “Red Hair, MC1R Variants, and Risk for Parkinson’s Disease—A Meta-Analysis.” Annals of Clinical and Translational Neurology 4 (3): 212–216.https://doi.org/10.1002/acn3.381.
Colliss Harvey, Jacky. 2015. Red: A History of the Redhead. New York: Black Dog & Levanthal.
Corning, Peter A. 2020. “Beyond the Modern Synthesis: A Framework for a More Inclusive Biological Synthesis.” Progress in Biophysics and Molecular Biology 153: 5–12.
Costa, James T. 2017. Darwin’s Backyard: How Small Experiments Led to a Big Theory. New York: W. W. Norton.
Crosby, Alfred W., Jr. 2003. The Columbian Exchange: Biological and Cultural Consequences of 1492. 30th Anniversary Edition. Westport, CT: Praeger.
Darwin, Charles. 1859. On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life. First Edition. London: John Murray.
Darwin, Francis, ed. 2001[1897]. The Life & Letters of Charles Darwin, vol. 2. Honolulu: University Press of the Pacific.
Desilver, Drew. 2017. “U.S. Students’ Academic Achievement Still Lags That of Their Peers in Many Other Countries.” Pew Research Center, February 15. Accessed May 25, 2022. https://www.pewresearch.org/fact-tank/2017/02/15/u-s-students-internationally-math-science/.
Devuyst, Olivier. 2014. “High Time for Human Height.” Peritoneal Dialysis International 34 (7):685–686.
Dobzhansky, Theodosius. 1937. Genetics and the Origin of Species. Columbia University Biological Series (Volume 11). New York: Columbia University Press.
Dunbar-Ortiz, Roxanne. 2014. An Indigenous Peoples’ History of the United States. Boston: Beacon.
El-Zaher, Sumaya. 2018. “The Father of the Theory of Evolution: Al-Jahiz and His Book of Animals.” MVSLIM.com, October 9. Accessed August 27, 2022. https://mvslim.com/the-father-of-the-theory-of-evolution-al-jahiz-and-his-book-of-animals/.
Elshakry, Marwa. 2010. “When Science Became Western: Historiographical Reflections.” ISIS 101 (1). Accessed November 20, 2022. https://www.journals.uchicago.edu/doi/10.1086/652691.
Gladwell, Malcolm. 2007. Blink: The Power of Thinking without Thinking. New York: Back Bay
Guadagno, Rosanna E., Bradley M. Okdie, and Sara A. Kruse. 2012. “Dating Deception: Gender, Online Dating, and Exaggerated Self-Presentation.” Computers in Human Behavior 28 (2): 642–647.
Harmon, Katherine. 2011. “Evolution Abroad: Creationism Evolves in Science Classrooms around the Globe.” Scientific American, March 3. Accessed May 25, 2022. https://www.scientificamerican.com/article/evolution-education-abroad/.
Harrison, Ellie. 2018. “This is What the Perfect Body Looks Like - According to Science.” Radiotimes.com. Accessed June 14, 2023. https://www.radiotimes.com/tv/documentaries/this-is-what-the-perfect-body-looks-like-according-to-science/.
Hatton, Tim. 2014. “Why Did Humans Grow Four Inches in 100 Years? It Wasn’t Just Diet.” The Conversation, May 1. https://theconversation.com/why-did-humans-grow-four-inches-In-100-years-it-wasnt-just-diet-25919.
Hays, J. N. 2005. Epidemics and Pandemics: Their Impacts on Human History. Santa Barbara, CA: ABC-CLIO.
Hutton, James. 1788. Theory of the Earth. Transactions of the Royal Society of Edinburgh, Volume 1. Scotland: Royal Society of Edinburgh.
IPCC. 2022. Climate Change 2022: Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, edited by H.-O. Pörtner, D.C. Roberts, M. Tignor, E.S. Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem, and B. Rama. Cambridge: Cambridge University Press. https://www.ipcc.ch/report/ar6/wg2/.
Jones, Stephen. 2017. “Religion, Science, and Evolutionary Theory”[an interview with Salman Hameed]. Podcast, The Religious Studies Project, January 30. Accessed May 29, 2023. https://www.religiousstudiesproject.com/podcast/religion-science-and-evolutionary-theory/.
Kutschera, Ulrich, and Karl J. Niklas. 2004. “The Modern Theory of Biological Evolution: An Expanded Synthesis.” Naturwissenschaften 91: 255–276.
Leclerc, Georges-Louis, Comte de Buffon. 1749-1804. Histoire Naturelle. Volumes 1-36. Paris: Imprimerie Royale (Royal Printing House).
Lents, Nathan H. 2018. Human Errors: A Panorama of Our Glitches, from Pointless Bones to Broken Genes. Boston: Houghton Mifflin Harcourt.
Lerner, Lawrence S. 2000. Good Science, Bad Science: Teaching Evolution in the States. Washington, DC: Thomas B. Fordham Foundation.
Lieberman, Daniel E. 2011. The Evolution of the Human Head. Cambridge, MA: Harvard University Press.
Lindberg, David C. 1992. The Beginnings of Western Science: The European Scientific Tradition in Philosophical, Religious, and Institutional Context, 600 B.C. to A.D. 1450. Chicago: The University of Chicago Press.
Linnaeus, Carl. 1736. Systema Naturae. First Edition. Stockholm: Laurentius Salvius.
Love, Shayla. 2020. “A Thousand Years before Darwin, Islamic Scholars Were Writing about Evolution.” VICE World News, October 5. Accessed August 27, 2022. https://www.vice.com/en/article/ep4ykn/a-thousand-years-before-darwin-islamic-scholars-were-writing-about-natural-selection.
Main, Douglas. 2013. “Ancient Mutation Explains Missing Wisdom Teeth.” Live Science, March 13. https://www.livescience.com/27529-missing-wisdom-teeth.html.
Malthus, Thomas Robert. 1798. An Essay on the Principle of Population. London: J. Johnson.
Masci, David. 2019. “For Darwin Day, 6 Facts About the Evolution Debate.” Pew Research Center. Accessed June 14, 2023. https://www.pewresearch.org/short-reads/2019/02/11/darwin-day/.
Masood, Ehsan. 2009. “Islam’s Evolutionary Legacy.” The Guardian, March 1. Accessed February 13, 2023. https://www.theguardian.com/commentisfree/belief/2009/feb/27/islam-religion-evolution-science.
Miller, Jon D., Eugenie C. Scott, Mark S. Ackerman, Belen Laspra, Glenn Branch, Carmelo Polino, and Jordan S. Huffaker. 2022. “Public Acceptance of Evolution in the United States, 1985–2020.” Public Understanding of Science 31 (2): 223–238.
Miller, Jon D., Eugenie C. Scott, and Shinji Okamoto. 2006, August 11. “Public Acceptance of Evolution.” Science 313 (5788): 765-766. DOI: 10.1126/science.1126746.
Moore, John A. 1993. Science as a Way of Knowing: The Foundations of Modern Biology. Cambridge, MA: Harvard University Press.
Moore, Michael P., Kaitlyn Hersch, Chanont Sricharoen, Sarah Lee, Caitlin Reice, Paul Rice, Sophie Kronick, Kim A. Medley, and Kasey D. Fowler-Finn. 2021. “Sex-Specific Ornament Evolution Is a Consistent Feature of Climatic Adaptation across Space and Time in Dragonflies.” PNAS 118 (28): e2101458118. https://doi.org/10.1073/pnas.210145818.
National Public Radio. 2023. “Forest Lizards Have Genetically Morphed to Survive Life in the City, Researchers Say.” National Public Radio (NPR), January 10. Accessed February 19, 2023. https://www.npr.org/2023/01/10/1148150056/forest-lizards-genetically-morphed-cities.
National Science Foundation. 2023. “Urban Lizards Share Genomic Markers Not Found in Forest-Dwellers.” National Science Foundation, February 7. Accessed May 25, 2023. https://beta.nsf.gov/news/urban-lizards-share-genomic-markers-not-found.
Novembre, John, Alison P. Galvani, and Montgomery Slatkin. 2005. “The Geographic Spread of the CCR5 Δ32 HIV-Resistance Allele.” PLoS Biology 3 (11): e339.
Pamuk, Şevket. 2007. “The Black Death and the Origins of the ‘Great Divergence’ across Europe, 1300–1600.” European Review of Economic History 11 (3): 289–317.
Paterlini, Marta. 2007. “There Shall Be Order: The Legacy of Linnaeus in the Age of Molecular Biology.” EMBO Reports 8 (9): 814–816. https://doi.org/10.1038/sj.embor.7401061.
Pew Research Center. 2019. “Darwin in America: The Evolution Debate in the United States.” Pew Research Center, February 6, 2019. Accessed June 13, 2023. https://www.pewresearch.org/religion/2019/02/06/darwin-in-america-2/.
Pew Research Center. 2014. “Fighting Over Darwin, State by State.” Pew Research Center, February 4, 2009; updated February 3, 2014. Accessed May 25, 2022. https://www.pewresearch.org/religion/2009/02/04/fighting-over-darwin-state-by-state/.
Plutynski, Anya. 2009. “The Modern Synthesis.” Routledge Encyclopedia of Philosophy, November 15 (updated December 14, 2009). Accessed November 27, 2022. http://philsci-archive.pitt.edu/15335/.
Plutzer, Eric, Glenn Branch, and Ann Reid. 2020. “Teaching Evolution in the U.S. Public Schools: A Continuing Challenge.” Evolution: Education and Outreach13: Article 14. https://evolution-outreach.biomedcentral.com/articles/10.1186/s12052-020-00126-8.
Proceedings of the National Academy of Sciences (PNAS). 2023. “How Lizards Adapt to Urban Living.” Science Sessions Podcast, February 13, . Accessed February 19, 2023. https://www.pnas.org/post/podcast/lizards-adapt-urban-living.
Public Broadcasting System. 2001. “Georges Cuvier.” WGBH Evolution Library. Accessed May 26, 2022. https://www.pbs.org/wgbh/evolution/library/02/1/l_021_01.html# :~:text=With%20elegant%20studies%20of%20the,as%20incontrovertible%20proof%20of %20extinctions.
Ray, John. 1686-1704. Historia plantarum. London: Clark. Volume 1 (1686), Volume II (1688), Volume III (1704).
Richards, Richard A. 1998. “Darwin, Domestic Breeding, and Artificial Selection.” Endeavour 22 (3): 106–109.
Solloch, Ute V., Kathrin Lang, Vinzenz Lange, and Irena Böhme. 2017. “Frequencies of Gene Variant CCR5-Δ32 in 87 Countries Based on Next-Generation Sequencing of 1.3 Million Individuals Sampled from 3 National DKMS Donor Centers.” Human Immunology 78 (11–12): 701-717.
Stewart, Thomas A., Justin B. Lemberg, Ailis Daly, Edward B. Daeschler, and Neil H. Shubin. 2022. “A New Epistostegalian from the Late Devonian of the Canadian Arctic.” Nature 608 (7923): 563–568.
Tasci, Ufuk Necat. 2020. “How a 10-Century Muslim Physicist Discovered How Humans See.” TRT World, May 25. Accessed May 18, 2022. https://www.trtworld.com/magazine/how-a-10th-century-muslim-physicist-discovered-how-humans-see-36620.
Tbakhi, Abdelghani, and Samir S. Amr. 2007. “Ibn Al-Haytham: Father of Modern Optics.” Annals of Saudi Medicine 27 (6): 464–467. UNAIDS. 2021. “Global HIV & AIDS Statistics—Fact Sheet.” https://www.unaids.org/en/resources/fact-sheet.
UNESCO.org. 2015. “International Year of Light: Ibn al Haytham, Pioneer of Modern Optics celebrated at UNESCO.” UNESCO, September 8. Accessed May 18, 2022. https://www.unesco.org/en/articles/international-year-light-ibn-al-haytham-pioneer-modern-optics-celebrated-unesco.
University of California Berkeley Museum of Paleontology. N.d. “The History of Evolutionary Thought—Uniformitarianism: Charles Lyell.” Understanding Evolution website. Accessed February 13, 2023. https://evolution.berkeley.edu/the-history-of-evolutionary-thought/1800s/uniformitarianism-charles-lyell/.
University of Cambridge. N.d. “Darwin and Dogs.” The Darwin Correspondence Project website. Accessed February 17, 2023. https://www.darwinproject.ac.uk/commentary/curious/darwin-and-dogs.
Urbach, Peter Michael, Anthony M. Quinton, and Kathleen Marguerite Lea. “Francis Bacon: British Author, Philosopher, and Statesman.” Encyclopaedia Britannica. Last updated May 12, 2023. https://www.britannica.com/biography/Francis-Bacon-Viscount-Saint-Alban.
Ward, Peter. 2018. Lamarck’s Revenge: How Epigenetics Is Revolutionizing Our Understanding of Evolution’s Past and Present. New York: Bloomsbury.
Weismann, August. 1892. Das Keimplasma: Eine Theorie der Vererbung (The Germ Plasm: a Theory of Inheritance). Jena (Germany): Fischer.
Weismann, August. 1889. Translations. Essays upon Heredity. Oxford: Clarendon. Accessed November 27, 2022. E-copy available at http://www.esp.org/books/weismann/essays/facsimile/.
Winchell, Kristen M., Shane C. Campbell-Staton, Jonathan B. Losos, and Anthony Geneva. 2023. “Genome-Wide Parallelism Underlies Contemporary Adaptation In Urban Lizards.” PNAS 120 (3): e2216789120. https://doi.org/10.1073/pnas.2216789120.
Wood, Bernard. 2005. Human Evolution: A Very Short Introduction. Oxford: Oxford University Press.
Zou, Yawen. 2015. "The Germ-Plasm: a Theory of Heredity (1893), by August Weismann." Embryo Project Encyclopedia, January 26. Accessed February 18, 2023. http://embryo.asu.edu/handle/10776/8284.
This is where you can write your introduction.
Michael B. C. Rivera, Ph.D., University of Cambridge
This chapter is a revision from "Chapter 13: Race and Human Variation” by Michael B. C. Rivera. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Learning Objectives
- Illustrate the troubling history of “race” concepts.
- Explain human variation and evolution as the thematic roots of biological anthropology as a discipline.
- Critique earlier “race” concepts based on a contemporary understanding of the apportionment of human genetic variation.
- Explain how biological variation in humans is distributed clinally and in accordance with both isolation-by-distance and Out-of-Africa models.
- Identify phenotypic traits that reflect selective and neutral evolution.
- Extend this more-nuanced view of human variation to today’s research, the implications for biomedical studies, applications in forensic anthropology, and other social/cultural/political concerns.
Humans exhibit biological variation. Humans also have a universal desire to categorize other humans in order to make sense of the world around them. Since the birth of the discipline of biological anthropology (or physical anthropology, as referred to back then), we have been interested in studying how humans vary biologically and what the sources of this variation are. Before we tackle these big problems, first consider this question: Why should we study human variation?
There are certainly academic reasons for studying human variation. First, it is highly interesting and important to consider the evolution of our species (see Chapters 9–12) and how our biological variation may be similar to (or different from) that of other species (e.g., other primates and apes; see Chapters 5 and 6). Second, anthropologists study modern human variation to understand how different biological traits developed over evolutionary time (see Figure 13.1). Suppose we are able to grasp the evolutionary processes that produce the differences in biology, physiology, body chemistry, behavior, and culture (human variation). In that case, we can make more accurate inferences about evolution and adaptation among our hominin ancestors, complementing our study of fossil evidence and the archaeological record. Third, as will be discussed in more detail later on, it is important to consider that biological variation among humans has biomedical, forensic, and sociopolitical implications. For these reasons, the study of human variation and evolution has formed the basis of anthropological inquiry for centuries and continues to be a major source of intrigue and inspiration for scientific research conducted today.
An even-more-important role of the biological anthropologist is to improve public understanding of human evolution and variation—outside of academic circles. Terms such as race and ethnicity are used in everyday conversations and in formal settings within and outside academia. The division of humankind into smaller, discrete categories is a regular occurrence in day-to-day life. This can be seen regularly when governments acquire census data with a heading like “geographic origin” or “ethnicity.” Furthermore, such checkboxes and drop-down lists are commonly seen as part of the identifying information required for surveys and job applications.
According to professors of anthropology, ethnic studies, and sociology, race is often understood as rooted in biological differences, ranging from such familiar traits as skin color or eye shape to variations at the genetic level. However, race can also be studied as an “ideological construct” that goes beyond biological and genetic bases (Fuentes et al. 2019), at different times relating to our ethnicities, languages, religious beliefs, and cultural practices. Sometimes people associate racial identity with the concept of socioeconomic status or position, or they link ideas about race to what passport someone has, how long they have been in a country, or how well they have “integrated” into a population.
Some of these ideas about ethnicity have huge social and political impacts, and notions of race have been part of the motivation behind various forms of racism and prejudice today, as well as many wars and genocides throughout history. Racism manifests in many ways—from instances of bullying between kids on school playgrounds to underpaid minorities in the workforce, and from verbal abuse hurled at people of color to violent physical behaviors against those of a certain race. Prejudice can be characterized as negative views toward another group based on some perceived characteristic that makes all members of that group worthy of disdain, disrespect, or exclusion (not solely along racial lines but also according to [dis]ability, gender, sexual orientation, or socioeconomic background, for example). According to Shay-Akil McLean (2014), “Racism is not something particular to the United States and race is not the same everywhere in the world. Racial categories serve particular contextual purposes depending on the society they are used in, but generally follow the base logic of the supremacy of one type of human body over all others (ordering these human bodies in a hierarchical fashion).” Choosing which biological or nonbiological features to use when discussing race is always a social process (Omi and Winant 2014). Race concepts have no validity to them unless people continue to use them in their daily lives—and, in the worst cases, to use them to justify racist behaviors and problematic ideas about racial difference or superiority/inferiority. Recent work in anthropological genetics has revealed the similarities amongst humans on a molecular level and the relatively few differences that exist between populations (Omi and Winant 2014).
The role of the biological anthropologist becomes crucial in the public sphere, because we may be able to debunk myths surrounding human variation and shed light, for the nonanthropologists around us, on how human variation is actually distributed worldwide (see Figure 13.1). Rooted in scientific observations, our work can help nonanthropologists recognize how common ideas about “race” often have no biological or genetic basis. Many anthropologists work hard to educate students on the history of where race concepts come from, why and how they last in public consciousness, and how we become more conscious of racial issues and the need to fight against racism in our societies. Throughout this chapter, I will highlight how humans cannot actually be divided into discrete “races,” because most traits vary on a continuous basis and human biology is, in fact, very homogenous compared to the greater genetic variation we observe in closely related species. Molecular anthropology, or anthropological genetics, continues to add new layers to our understanding of human biological variation and the evolutionary processes that gave rise to the contemporary patterns of human variation. The study of human variation has not always been unbiased, and thinkers and scientists have always worked in their particular sociohistorical context. For this reason, this chapter opens with a brief overview of race concepts throughout history, many of which relied on unethical and unscientific notions about different human groups.
Special Topic: My Experiences as an Asian Academic
My name is Michael, and I am a biological anthropologist and archaeologist (Figure 13.2). What strikes me as most interesting to investigate is human biological variation today and the study of past human evolution. For instance, some of my research on ancient coastal populations has revealed positive effects of coastal living on dietary health and many unique adaptations in bones and teeth when living near rivers and beaches. I love talking to students and nonscientists about bioanthropologists’ work, through teaching, science communication events, and writing book chapters like this one. I grew up in Hong Kong, a city in southern China. My father is from the Philippines and my mother is from Hong Kong, which makes me a mixed Filipino-Chinese academic. Growing up, I noticed that people came in all shapes, sizes, and colors. My life is very different now in that I have gained the expertise to explain those differences, and I feel a great responsibility to guide those new to anthropology toward their own understandings of diversity.
Biological anthropology is not taught extensively in Hong Kong, so I moved to the United Kingdom to earn my bachelor’s, master’s, and doctorate degrees. It was fascinating to me that we could answer important questions about human variation and history using scientific methods. However, I did not have many minority academic role models to look up to while I was at university. My department was made up almost exclusively of white westerner faculty, and it was hard to imagine I could one day get a job at these western institutions. While pursuing my degrees, I also remember several instances of my research contributions being overlooked or dismissed. Sometimes professors and fellow students would make racist comments toward Asian scholars (including me) and other Black, Indigenous and researchers of color, making us greatly uncomfortable in those spaces. When one of us would work up the courage to tell university leaders our experiences of being stereotyped, dismissed, or insulted, we received little support and were further excluded from research and teaching activities. This is a common experience for Black, Indigenous, and other people of color who pursue biological anthropology, and we face the difficult choice between leaving the field or bearing with such unsafe spaces.
It became important to me at that time to find other academics of color with whom to share experiences and form community. I feel inspired by all my colleagues who advocate for greater representation and diversity in universities (whether they are minority academics or not). I admire many of my fellow researchers who are underrepresented and do a great job of representing minority groups through their cutting-edge research and quality teaching at the undergraduate and graduate levels. Although I no longer work in the West, it has remained my great hope that those in the West and the “Global North” will continue to improve university culture, and I support any efforts there to welcome all scholars.
My current work is based in Hong Kong, where I am deeply dedicated to helping develop biological anthropology in East and Southeast Asia and promoting research from our home regions on the international scene. The study of anthropology really highlights how we share a common humanity and history. Being somebody who is “mixed race” and Asian likely played a key role in steering me toward the study of human variation. As this chapter hopefully shows, there is a lot to discuss about race and ethnicity regarding the discipline’s history and current understandings of human diversity. Some scientific and technological advancements today are misused for reasons to do with money, politics, or the continuation of antiquated ideas. It is my belief, alongside many of my friends and fellow anthropologists, that science should be more about empathy toward all members of our species and contributing to the intellectual and technological nourishment of society. After speaking to many members of the public, as well as my own undergraduate students, I have received lovely messages from other individuals of color expressing thanks and appreciation for my presence and understanding as a minority scientist and mentor figure. Anthropology needs much more diversity as well as to make room for those who have traveled different routes into the discipline. All paths taken into anthropology are valid and valuable. I would encourage everyone to study anthropology—it is a field for understanding and celebrating the intricacies of human diversity.
The History of "Race" Concepts
“Race” in the Classical Era
The earliest classification systems used to understand human variation are evidenced by ancient manuscripts, scrolls, and stone tablets recovered through archaeological, historical, and literary research. The Ancient Egyptians had the Book of Gates, dated to the New Kingdom between 1550 B.C.E. and 1077 B.C.E (Figure 13.3). In one part of this tome dedicated to depictions of the underworld, scribes used pictures and hieroglyphics to illustrate a division of Egyptian people into the four categories known to them at the time: the Aamu (Asiatics), the Nehesu (Nubians), the Reth (Egyptians), and the Themehu (Libyans). Though not rooted in any scientific basis like our current understandings of human variation today, the Ancient Egyptians believed that each of these groups were made of a distinctive category of people, distinguishable by their skin color, place of origin, and even behavioral traits.
Another well-known early document is the Bible, where it is written that all humankind descends from one of three sons of Noah: Shem (the ancestor to all olive-skinned Asians), Japheth (the ancestor to pale-skinned Europeans), and Ham (the ancestor to darker-skinned Africans). Similar to the Ancient Egyptians, these distinctions were based on behavioral traits and skin color. More recent work in historiography and linguistics suggest that the branches of “Hamites,” “Japhethites,” and “Shemites” may also relate to the formation of three independent language groups some time between 1000 and 3000 B.C.E. With the continued proliferation of Christianity, this concept of approximately three racial groupings lasted until the Middle Ages and spread as far across Eurasia as crusaders and missionaries ventured at the time.
There is also the “Great Chain of Being,” conceived by ancient Greek philosophers like Plato (427‒348 B.C.E.) and Aristotle (384‒322 B.C.E.). They played a key role in laying the foundations of empirical science, whereby observations of everything from animals to humans were noted with the aim of creating taxonomic categories. Aristotle describes the Great Chain of Being as a ladder along which all objects, plants, animals, humans, and celestial bodies can be mapped in an overall hierarchy (in the order of existential importance, with humans placed near the top, just beneath divine beings; see Figure 13.4). When he writes about humans, Aristotle expressed the belief that certain people are inherently (or genetically) more instinctive rulers, while others are more natural fits for the life of a worker or enslaved person. Based on research by biological anthropologists, we currently recognize that these early systems of classification and hierarchization are unhelpful in studying human biological variation. Both behavioral traits and physical traits are coded for by multiple genes each, and how we exhibit those traits based on our genetics can vary significantly even between individuals of the same population.
“Race” during the Scientific Revolution
The 1400s to 1600s saw the beginnings of the Scientific Revolution in Western societies, with thinkers like Nicholas Copernicus, Galileo Galilei, and Leonardo Da Vinci publishing some of their most important findings. While by no means the first or only scholars globally to use observation and experimentation to understand the world around them, early scientists living at the end of the medieval period in Europe increasingly employed more experimentation, quantification, and rational thought in their work. This is the main difference between the work of the ancient Egyptians, Romans, and Greeks and that of later scientists such as Isaac Newton and Carl Linnaeus in the 1600s and 1700s.
Linnaeus is the author of Systema Naturae (1758), in which he classified all plants and animals under the formalized naming system known as binomial nomenclature (Figure 13.5). This system is typological, in that organisms are placed into groups according to how they are similar or different to others under study. What was most anthropologically notable about Linnaeus’s typological system was that he was one of the first to group humans with apes and monkeys, based on the anatomical similarities between humans and nonhuman primates. However, Linnaeus viewed the world in line with essentialism, a problematic concept that dictates that there are a unique set of characteristics that organisms of a specific kind must have and that would remove organisms from taxonomic categorizations if they lacked any of the required criteria.
Linnaeus subdivided the human species into four varieties, with overtly racist categories based on skin color and “inherent” behaviors. Some European scientists during this period were not aware of their own biases skewing their interpretations of biological variation, while others deliberately worked to shape public perceptions of human variation in ways that established “otherness” and enforced European domination and the subordination of non-European people. The conclusions and claims at which they arrived, consciously or subconsciously, often fit the times they were living through—the so-called Age of Discovery, when the superiority of European cultures over others was a pervasive idea throughout people’s social and political lives. Although much of Eurasia was linked by spice- and silk-trading routes, the European colonial period between the 1400s and 1700s was marked by many new and unfortunately violent encounters overseas (Figure 13.6). When Europeans arrived by ship on the shores of continents that were already inhabited, it was their first meeting with the Indigenous peoples of the Americas and Australasia, who looked, spoke, and behaved differently from peoples with whom they were familiar. Building on the idea of species and “subspecies,” natural historians of this time invented the term race, from the French rasse meaning “local strain.”
Another scientist of the times, Johann Friedrich Blumenbach (1752‒1840), classified humans into five races based on his observations of cranial form variation as well as skin color. He thus dubbed the “original” form of the human cranium the “Caucasian” form, with the idea that the ideal climate conditions for early humans would have been in the Caucasus region near the Caspian Sea. The key insight Blumenbach presented was that human variation in any particular trait should be more accurately viewed as falling along a gradation (Figure 13.7). While some of his theories were correct according to what we observe today with more knowledge in genetics, they erroneously believed that human “subspecies” were “degenerated” or “transformed” varieties of an ancestral Caucasian or European race. According to them, the Caucasian cranial dimensions were the least changed over human evolutionary time, while the other skull forms represented geographic variants of this “original.” As will be discussed in greater detail later in this chapter, we have genetic and craniometric evidence for sub-Saharan Africa being the origin of the human species instead (see Chapter 12 on the fossil record that places the origins of modern Homo sapiens in north and east Africa). Based on work that shows how most biological characteristics are coded for by nonassociated genes, it is not reasonable to draw links between individuals’ personalities and their skull shapes.
“Race” and the Dawn of Scientific Racism
Between the 1800s and mid-1900s, and contrary to what you might expect, an increased use of scientific methods to justify racial schemes developed in scholarship. Differing from earlier views, which saw all humans as environmentally deviated from one “original” humankind, classification systems after 1800 became more polygenetic (viewing all people as having separate origins) rather than monogenetic (viewing all people as having a single origin). Instead of moving closer to our modern-day understandings of human variation, there was increased support for the notion that each race was created separately and with different attributes (intelligence, temperament, and appearance).
The 1800s were an important precursor to modern biological anthropology as we know it, given that this was when the scientific measurement of human physical features (anthropometry) truly became popularized. However, empirical studies in the 1800s pushed even further the idea that Europeans were culturally and biologically superior to others. While considered one of the pioneers of American “physical” anthropology, Samuel George Morton (1799‒1851) was a scholar who had a large role in perpetuating 1800s scientific racism. By measuring cranial size and shape, he calculated that “Caucasians,” on average, have greater cranial volumes than other groups, such as the Indigenous peoples of the Americas and peoples Morton referred to collectively as “Negros.” Today, we know that cranial size variation depends on such factors as Allen’s and Bergmann’s rules, which give a more likely explanation: in colder environments, it is advantageous for those living there to have larger and rounder heads because they conserve heat more effectively than more slender heads (Beals et al. 1984). The leading figures in craniometry during the 1800s instead were linked heavily with powerful individuals and wealthy sociopolitical institutions and financial bodies. Theories in support of hierarchical racial schemes using “scientific” bases certainly helped continue the exploitative and unethical trafficking and enslavement of Africans between the 1500s and 1800s.
Morton went on to write in his publication Crania Americana (1839) a number of views that fit with a concept called biological determinism. The idea behind biological determinism is that an association exists between people’s physical characteristics and their behavior, intelligence, ability, values, and morals. If the idea is that some groups of people are essentially superior to others in cognitive ability and temperament, then it is easier to justify the unequal treatment of certain groups based on outward appearances. Another such problematic thinker was Paul Broca (1824‒1880), after whom a region of the frontal lobe related to language use is named (Broca’s area). Influenced by Morton, Broca likewise claimed that internal skull capacities could be linked with skin color and cognitive ability. He went on to justify the European colonization of other global territories by purporting it was noble for a biologically more “civilized” population to improve the “humanity” of more “barbaric” populations. Today, these theories of Morton, Broca, and others like them are known to have no scientific basis. If we could speak with them today, they would likely try to emphasize that their conclusions were based on empirical evidence and not a priori reasoning. However, we now can clearly see that their reasoning was biased and affected by prevailing societal views at the time.
“Race” and the Beginnings of Physical Anthropology
In the early 20th century, we saw a number of new figures coming into the science of human variation and shifting the theoretical focus within. Most notably, these included Aleš Hrdlička and Franz Boas.
Aleš Hrdlička (1869‒1943) was a Czech anthropologist who moved to the United States. In 1903, he established the physical anthropology section of the National Museum of Natural History (Figure 13.8). In 1918, he founded the American Journal of Physical Anthropology, which remains one of the foremost scientific journals disseminating bioanthropological research. As part of his work and the scope of the journal, he differentiated “physical anthropology” from other kinds of anthropology: he wrote that physical anthropology is “the study of racial anatomy, physiology, and pathology” and “the study of man’s variation” (Hrdlička 1918). In some ways, although the scope and technological capabilities of biological anthropologists have changed significantly, Hrdlička established an area of inquiry that has continued and prospered for over a hundred years.
Franz Boas (1858‒1942) was a German American anthropologist who established the four-field anthropology system in the United States and founded the American Anthropological Association in 1902. He argued that the scientific method should be used in the study of human cultures and the comparative method for looking at human biology worldwide. One of Boas’s significant contributions to biological anthropology was the study of skull dimensions with respect to race. After a long-term research project, he demonstrated how cranial form was highly dependent on cultural and environmental factors and that human behaviors were influenced primarily not by genes but by social learning. He wrote in one essay for the journal Science: “While individuals differ, biological differences between races are small. There is no reason to believe that one race is by nature so much more intelligent, endowed with great willpower, or emotionally more stable than another, that the difference would materially influence its culture” (Boas 1931:6). This conclusion directly contrasted with the theories of the past that relied on biological determinism. Biological anthropologists today have found evidence that corroborates Boas’s explanations: societies do not exist on a hierarchy or gradation of “civilizedness” but instead are shaped by the world around them, their demographic histories, and the interactions they have with other groups.
The first half of the 1900s still involved some research that was essentialist and focused on proving racial determinism. Anthropologists like Francis Galton (1822‒1911) and Earnest A. Hooton (1887‒1954) created the field of eugenics as an attempt to formalize the social scientific study of “fitness” and “superiority” among members of 19th-century Europe. As a way of “dealing with” criminals, diseased individuals, and “uncivilized” people, eugenicists recommended prohibiting parts of the population from being married or sterilizing these members of society so they could no longer procreate (Figure 13.9). They instead encouraged “reproduction in individual families with sound physiques, good mental endowments, and demonstrable social and economic capability” (Hooton 1936). In the 1930s, Nazi Germany used this false idea of there being “pure races” to highly destructive effect. The need to be protected against admixture from “unfit” groups was their justification for their blatant racism and purging of citizens that fell under their subjective criteria.
It should be noted that eugenicist ideas were popularly discussed and debated in many non-European contexts, as in the U.S., China, and South Africa, too. The Immigration Restriction Act of 1924 was passed in the United States, with the explicit aim of reducing the country’s “burden” of people considered inferior by restricting immigration of eastern European Jews, Italians, Africans, Arabs, and Asians. In the early 1900s, Chinese scientists and politicians showed great interest in eugenic ideologies, which came to dictate decisions in law-making, family life, and the medical field. Noted American anthropologist Ruth Benedict wrote extensively on Japanese culture and society during and after World War II. Her essentialist portrayals of Japanese people were heavily cited in popular discourse at the time. In 1950s South Africa, interracial marriages and sexual relations were banned by law; antimiscegenation became one of the huge focuses of apartheid resistance activists in later years. We still see the continuation of eugenics-based logic today around the world—in exclusionary immigration laws, cases of incarcerated prison inmates being forcibly sterilized, and the persistence of intelligence testing as a form of measuring people’s “fitness” in a society.
Shortly after World War II and the Nazi Holocaust, the full extent of essentialist, eugenicist thinking became clear. Social constructions of race, and the notion that one can predict psychological or behavioral traits based on external appearance, had become unpopular both within and outside the discipline. It was up to those in the field of physical anthropology at the time to separate physical anthropology from race concepts that supported unscientific and socially damaging agendas. This does not mean that there are no physiological or behavioral differences between different members of the human species. However, going forward, a number of physical anthropologists saw human biological variation as more complicated than simple typologies could describe.
“The New Physical Anthropology”
After 1950, focus steered away from the concept of “race” as a unit of variation and toward understanding why variation exists in populations from an evolutionary perspective. This was outlined by those pioneering the “new physical anthropology,” such as Sherwood Washburn, Theodosius Dobzhansky (Figure 13.10), and Julian Huxley, who borrowed this approach from contemporary population geneticists. Whether using genetic or phenotypic markers as the units of study, “groups” or “clusters” of humans differentiated by these became defined as populations that differ in the frequency of some gene or genes. Anthropologists consider what “makes” a population—a group of individuals potentially capable of or actually interbreeding due to shared geographic proximity, language, ethnicity, culture, and/or values. Put another way, a population is a local interbreeding group with reduced gene flow between themselves and other groups of humans. Members of the same population may be expected to share many genetic traits (and, as a result, many phenotypic traits that may or may not be visible outwardly).
Thinking of humans in terms of populations was part of Julian Huxley’s (1942) “Modern Synthesis”—so named because it helped to reconcile two fundamental principles about evolution that had not been made sense of together before (Figure 13.11). As discussed in Chapter 3, Gregor Mendel (1822‒1884) was able to show that inheritance was mediated by discrete particles (or genes) and not blended in the offspring. However, it was difficult for some 19th-century scientists to accept this model of genetic inheritance at the time because much of biological variation appeared to be continuous and not particulate (take skin color or height as examples). In the 1930s, it was demonstrated that traits could be polygenic and that multiple alleles could be responsible for any one phenotypic trait, thus producing the continuous variation in traits such as eye color that we see today. Thus, Huxley’s “Modern Synthesis” outlines not only how human populations are capable of exchanging genes at the microevolutionary level but also how multiple alleles for one trait (polygenic exchanges) can cause gradual macroevolutionary changes.
Human Variation in Biological Anthropology Today
Many Human Traits Are Nonconcordant
In our studies of human (genetic) variation today, we understand most human traits to be nonconcordant (Figure 13.12). “Nonconcordance” is a term used to describe how biological traits vary independent of each other—that is, they do not get inherited in a correlative manner with other genetically controlled traits. For example, if you knew an individual had genes that coded for tall height, you would not be able to predict if they are lighter-skinned or have red hair. This is different from earlier essentialist views of human variation: the idea that skin color could predict one’s brain function or even “temperament” and tendencies toward criminal behavior.
Human Variation Is Clinal/Continuous (Not Discrete)
Frank B. Livingstone (1928‒2005) wrote: “There are no races, only clines” (1962: 279). A cline is a gradation in the frequency of an allele/trait between populations living in different geographic regions. Human variation cannot be broken into discrete “races,” because most physical traits vary on a continuous or “clinal” basis. One obvious example of this is how human height does not only come in three values (“short,” “medium,” and “tall”) but instead varies across a spectrum of vertical heights achievable by humans all over the world. On the one hand, we can describe human height as exhibiting continuous variation, forming a continuous pattern, but height does not vary according to where people live across the globe and does not exhibit a clinal pattern. On the other hand, skin color variation between populations does show patterning that fits quite well on to how near or far they are from each other on a world map. This makes a trait like skin color clinally distributed worldwide. When large numbers of genetic loci for large numbers of samples were sampled from human populations distributed worldwide during the 1960s and 1970s, the view that certain facets of human diversity were clinally distributed was further supported by genetic data.
To study human traits that are clinally distributed, genetic tests must be performed to uncover the true frequencies of an allele or trait across a certain geographic space. One easily visible example of a clinal distribution seen worldwide is the patterning of human variation in skin color. Whether in southern Asia, sub-Saharan Africa, or Australia, dark brown skin is found. Paler skin tones are found in higher-latitude populations such as those who have lived in areas like Europe, Siberia, and Alaska for millennia. Skin color is easily observable as a phenotypic trait exhibiting continuous variation.
A clinal distribution still derives from genetic inheritance; however, clines often correspond to some gradually changing environmental factor. Clinal patterns arise when selective pressures in one geographic area differ from those in another as well as when people procreate and pass on genes together with their most immediate neighbors. There are several mechanisms, selective and neutral, that can lead to the clinal distribution of an allele or a biological trait. Natural selection is the mechanism that produced a global cline of skin color, whereby darker skin color protects equatorial populations from high amounts of UV radiation; there is a transition of lessening pigmentation in individuals that reside further and further away from the tropics (Jablonski 2004; Jablonski and Chaplin 2000; see Figure 13.13). The ability and inability to digest lactose (milk sugar) among different world communities varies according to differential practices and histories of milk and dairy-product consumption (Gerbault et al. 2011; Ingram et al. 2009). Where malaria seems to be most prevalent as a disease stressor on human populations, a clinal gradient of increasing sickle cell anemia experience toward these regions has been studied extensively by genetic anthropologists (Luzzatto 2012). Sometimes culturally defined mate selection based on some observable trait can lead to clinal variation between populations as well.
Two neutral microevolutionary processes that may produce a cline in a human allele or trait are gene flow and genetic drift (see Chapter 4). The ways in which neutral processes can produce clinal distributions is seen clearly when looking at clinal maps for different blood groups in the human ABO blood group system (Figure 13.14). For instance, scientists have identified an East-to-West cline in the distribution of the blood type B allele across Eurasia. The frequency of B allele carriers decreases gradually westward when we compare the blood groups of East and Southeast Asian populations with those in Europe. This shows how populations residing nearer to one another are more likely to interbreed and share genetic material (i.e., undergo gene flow). We also see 90%‒100% of native South American individuals, as well as between 70%‒90% of Aboriginal Australian groups, carrying the O allele (Mourant, Kopeć, and Domaniewska-Sobczak 1976). These high frequencies are likely due to random genetic drift and founder effects, in which population sizes were severely reduced by the earliest O allele-carrying individuals migrating into those areas. Over time, the O blood type has remained predominant.
Genetic Variation Is Greater Within Group than Between Groups
One problem with race-based classifications is they relied on an erroneous idea that individuals with particular characteristics would share more similar genes with each other within a particular “race” and share less with individuals of other “races” possessing different traits and genetic makeups. However, since around 50 years ago, scientific studies have shown that the majority of human genetic differences worldwide exist within groups (or “races”) individually rather than between groups. Indeed, most genetic variation we see occurs in Africa, and many variants are shared among individuals on all continents (Figure 13.15).
In 2002, a landmark article by Noah Rosenberg and colleagues explored worldwide human genetic variation using an even-greater genetic data set. They used 377 highly variable markers in the human genome and sampled from 1,056 individuals representative of 52 populations. The markers chosen for study were not ones that code for any expressed genes. Because these regions of the human genome were made of unexpressed genes, we may understand these markers as neutrally derived (as opposed to selectively derived) because they do not code for functional advantages or disadvantages. These neutral genetic markers likely reflect an intricate combination of regional founder effects and population histories. Analyses of these neutral markers allowed scientists to identify that 93%‒95% of global genetic differences, referred to as variance, can be accounted for by within-population differences, while only a small proportion of genetic variance (3%‒5%) can be attributed to differences among major groups (Rosenberg et al. 2002). This research supports the theory that distinct biological races do not exist, even though misguided concepts of race may still have real social and political consequences.
Biological Data Fit Isolation-By-Distance and Out-of-Africa Models
One further note is that the world’s population may be genetically divided into “groups,” “subsets,” “clumps,” or “clusters” that reflect some degree of genetic similarity. These identifiable clusters reflect genetic or geographic distances—either with gene flow facilitated by proximity between populations or impeded by obstacles like oceans or environmentally challenging habitats (Rosenberg et al. 2005). Sometimes, inferred clusters using multiple genetic loci are interpreted by nongeneticists literally as “ancestral populations.” However, it would be wrong to assume from these genetic results that highly differentiated and “pure” ancestral groups ever existed. These groupings reflect differences that have arisen over time due to clinal patterning, genetic drift, and/or restricted or unrestricted gene flow (Weiss and Long 2009). The clusters identified by scientists are arbitrary and the parameters used to split up the global population into groups is subjective and dependent on the particular questions or distinctions being brought into focus (Relethford 2009).
Additionally, research on worldwide genetic variation has shown that human variation decreases with increasing distance from sub-Saharan Africa, where there is evidence for this vast region being the geographical origin of anatomically modern humans (Liu et al. 2006; Prugnolle, Manica, and Balloux 2005; see Figures 13.16 and 13.17). Genetic differentiation decreases in human groups the further you sample data from relative to sub-Saharan Africa because of serial founder effects (Relethford 2004). Over the course of human colonization of the rest of the world outside Africa, populations broke away in expanding waves across continents into western Asia, then Europe and eastern Asia, followed by Oceania and the Americas. As a result, founder events occurred whereby genetic variation was lost, as the colonization of each new geographical region involved a smaller number of individuals moving from the original larger population to establish a new one (Relethford 2004). The most genetic variation is found across populations residing in different parts of sub-Saharan Africa, while other current populations in places like northern Europe and the southern tip of South America exhibit some of the least genetic differentiation relative to all global populations (Campbell and Tishkoff 2008).
Besides fitting nicely into the Out-of-Africa model, worldwide human genetic variation conforms to an isolation-by-distance model, which predicts that genetic similarity between groups will decrease exponentially as the geographic distance between them increases (Kanitz et al. 2018). This is because of the greater and greater restrictions to gene flow presented by geographic distance, as well as cultural and linguistic differences that occur as a result of certain degrees of isolation. Since genetic data conform to isolation-by-distance and Out-of-Africa models, these findings support the abolishment of “race” groupings. This research demonstrates that human variation is continuous and cannot be differentiated into geographically discrete categories. There are no “inherent” or “innate” differences between human groups; instead, variation derives from some degree of natural selection, as well as neutral processes like population bottle-necking (Figure 13.18), random mutations in the DNA, genetic drift, and gene flow through between-mate interbreeding.
Humans Have Higher Homogeneity Compared to Many Other Species
An important fact to bear in mind is that humans are 99.9% identical to one another. This means that the apportionments of human variation discussed above only concern that tiny 0.1% of difference that exists between all humans globally. Compared to other mammalian species, including the other great apes, human variation is remarkably lower. This may be surprising given that the worldwide human population has already exceeded seven billion, and, at least on the surface level, we appear to be quite phenotypically diverse. Molecular approaches to human and primate genetics tells us that external differences are merely superficial. For a proper appreciation of human variation, we have to look at our closest relatives in the primate order and mammalian class. Compared to chimpanzees, bonobos, gorillas and other primates, humans have remarkably low average genome-wide heterogeneity (Osada 2005).
When we look at chimpanzee genetic variation, it is fascinating that western, central, eastern, and Cameroonian chimpanzee groups have substantially more genetic variation between them than large global samples of human DNA (Bowden et al. 2012; Figure 13.19). This is surprising given that all of these chimpanzee groups live relatively near one another in Africa, while measurements of human genetic variation have been conducted using samples from entirely different continents. First, geneticists suppose that this could reflect differential experiences of the founder effect between humans and chimpanzees. Because all non-African human populations descended from a small number of anatomically modern humans who left Africa, it would be expected that all groups descended from that smaller ancestral group would be similar genetically. Second, our species is really young, given that we have only existed on the planet for around 150,000 to 300,000 years. This gave humans little time for random genetic mutations to occur as genes get passed down through genetic interbreeding and meiosis. Chimpanzees, however, have inhabited different ecological niches, and less interbreeding has occurred between the four chimpanzee groups over the past six to eight million years compared to the amount of gene flow that occurred between worldwide human populations (Bowden et al. 2012).
Recent advances have now enabled the attainment of genetic samples from the larger family of great apes and the evaluation of genetic variation among bonobos, orangutans, and gorillas alongside that of chimpanzees and humans (Prado-Martinez et al. 2013). Collecting such data and analyzing primate genetic variation has been important not only to elucidate how different ecological, demographic, and climatic factors have shaped our evolution but also to inform upon conservation efforts and medical research. Genes that may code for genetic susceptibilities to tropical diseases that affect multiple primates can be studied through genome-wide methods. Species differences in the genomes associated with speech, behavior, and cognition could tell us more about how human individuals may be affected by genetically derived neurological or speech-related disorders and conditions (Prado-Martinez et al. 2013; Staes et al. 2017). In 2018, a great ape genomic study also reported genetic differences between chimpanzees and humans related to brain cell divisions (Kronenberg et al. 2018). From these results, it may be inferred that cognitive or behavioral variation between humans and the great apes might relate to an increased number of cortical neurons being formed during human brain development (Kronenberg et al. 2018). Comparative studies of human and nonhuman great ape genetic variation highlight the complex interactions of population histories, environmental changes, and natural selection between and within species. When viewed in the context of overall great ape variation, we may reconsider how variable the human species is relatively and how unjustified previous “race” concepts really were.
Phenotypic Traits That Reflect Neutral Evolution
Depending on the trait being observed, different patterns of phenotypic variation may be found within and among groups worldwide. In this subsection, some phenotypic traits that reflect the aforementioned patterns of genetic variation will be discussed.
Looking beyond genetic variation briefly, recent studies have revisited biological anthropology’s earlier themes of externally observable traits, such as skull shape. In the last 20 or so years, anthropologists have evaluated the level to which human cranial shape variation reflects the results from genetic markers, such as those used previously to fit against Out-of-Africa models (Relethford 2004) or those used in the apportionment of human variation between and within groups (Lewontin 1972; Rosenberg et al. 2002). Using larger sample sizes of cranial data collected from thousands of skulls worldwide and a long list of cranial measurements, studies demonstrate a similar decrease in variation with distance from Africa and show that a majority of cranial variation occurs within populations rather than between populations (Betti et al. 2009; Betti et al. 2010; Manica et al. 2007; Relethford 2001; von Cramon-Taubadel and Lycett 2008; see Figure 13.20). The greatest cranial variation is found among skulls of sub-Saharan African origin, while the least variation is found among populations inhabiting places like Tierra del Fuego at the southern tip of Argentina and Chile. While ancient and historical thinkers previously thought “race” categories could reasonably be determined based on skull dimensions, modern-day analyses using more informative sets of cranial traits simply show that migrations out of Africa and the relative distances between populations can explain a majority of worldwide cranial variation (Betti et al. 2009).
This same patterning in phenotypic variation has even been found in studies examining shape variation of the pelvis (Betti et al. 2013; Betti et al. 2014), the teeth (Rathmann et al. 2017), and the human bony labyrinth of the ear (Ponce de León et al. 2018;Figure 13.21). The skeletal morphology of these bones still varies worldwide, but a greater proportion of that variation can still be attributed to the ways in which human populations migrated across the world and exchanged genes with those closer to them rather than those further away. Human skeletal variation in these parts of the body is continuous and nondiscrete. Given the important functions of the cranium and these other skeletal parts, we may infer that the genes that underpin their development have been relatively conserved by neutral evolutionary processes such as genetic drift and gene flow. It is also important to note that while some traits such as height, weight, cranial dimensions, and body composition are determined, in part, by genes, the underlying developmental processes behind these traits are underpinned by complex polygenic mechanisms that have led to the continuous spectrum of variation in such variables among modern-day human populations.
Phenotypic Traits That Reflect Natural Selection
Even though 99.9% of our DNA is the same across all humans worldwide, and many traits reflect neutral processes, there are parts of that remaining 0.1% of the human genome that code for individual and regional differences. Similarly to craniometric analyses that have been conducted in recent decades, human variation in skin color has also been reassessed using new methods and in light of greater knowledge of biological evolution.
New technologies allow scientists to use color photometry to sample and quantify the visible wavelength of skin color, in a way 19th- and 20th-century readers could not. In one report, it was found that 87.9% of global skin color variation can be attributed to genetic differences between groups, 3.2% to those among local populations within regions, and 8.9% within local populations (Relethford 2002). This apportionment differs significantly and is the reverse situation found in the distribution of genetic differences we see when we examine genetic markers such as blood type–related alleles. However, this pattern of human skin color worldwide is not surprising, given that we now understand that past selection has occurred for darker skin near the equator and lighter skin at higher latitudes (Jablonski 2004; Jablonski and Chaplin 2000). While most genetic variation reflects neutral variation due to population migrations, geographic isolation, and restricted gene flow dynamics, some human genetic/phenotypic variation is best explained as local adaptation to environmental conditions (i.e., selection). Given that skin color variation is atypical compared to other genetic markers and biological traits, this, in fact, goes against earlier “race” typologies. This is because recent studies ironically show how so much of genetic variation relates to neutral processes, while skin color does not. It follows that skin color cannot be viewed as useful in making inferences about other human traits.
It is also true that some populations have not been studied extensively in skin pigmentation genetics (e.g., African, Austronesian, Melanesian, Southeast Asian, Indigenous American, and Pacific Islander populations, according to Lasisi and Shriver 2018). Earlier dispersals of these populations, and their local genetic varition, will have contributed to worldwide genetic variation, inclusive of skin pigmentation variation. Gene loci we did not previously appreciate as being linked to pigmentation are now being recognized thanks to better tools, more diverse genetic samples, and more accessible datasets (Quillen et al. 2018). Biological anthropologists look forward to further discoveries elucidating the different selective pressures and population dynamics that influence skin pigmentation evolution.
Social Implications
To finish this chapter, we will consider the social, economic, political, and biological implications of poor understandings of race and the deliberate perpetuation of social and medical racism.
The Black Lives Matter movement (BLM) of 2013 began with the work of racial justice activists and community organizers Alicia Garza, Opal Tometi, and Patrissa Cullors. First incited by the murder of Trayvon Martin, a 17-year-old African American, and the acquittal of the man who shot him, BLM went on to protest against the deaths of numerous Black individuals, most of whom were killed by police officers (for example, Ahmaud Arbery was killed in February of 2020 by two white non-police officers). Some key characteristics of BLM from the start were its decentralized grassroots structure, the role of university students and social media in spreading awareness of the movement, and its embrace of other movements (e.g., climate justice, ending police brutality, feminist campaigns, queer activism, immigration reform, etc.). When George Floyd was murdered by a white police officer on May 25, 2020, the BLM gained new momentum, across 2,000-plus cities in the United States, and among many protesting against historic racism and police brutality in other contexts around the globe. Many in the biological anthropology community have responded to these events with a great dedication to working against systemic racism in society and institutions (American Association of Biological Anthropologists 2020).
BLM continues to be an important movement, as is evidenced in the degree of community organizing, mutual aid efforts, calls for political reform, progress toward curriculum reform and equality, inclusion and diversity (EDI) work in businesses and universities, the removal of monuments honoring historical figures associated with slavery and racism, and many other important actions. Garza (2016) writes: “The reality is that race in the United States operates on a spectrum from black to white … the closer you are to white on that spectrum, the better off you are.” Tometi (2016) has stated: “We need [a human rights movement that challenges systemic racism] because the global reality is that Black people are subject to all sorts of disparities in most of our challenging issues of our day. I think about climate change, and how six of the ten worst impacted nations by climate change are actually on the continent of Africa.” In the words of Cullors (2016), “Black Lives Matter is our call to action. It is a tool to reimagine a world where Black people are free to exist, free to live. It is a tool for our allies to show up differently for us.” We gather from their words the importance of learning from the egregious role that anthropologists have played in the past, recognizing the legacies of “scientific” justifications for eugenics and racism in our society today, and proactively working toward environmental and social equity.
Another major industry that engages in the quantification and interpretation of human variation is medical and clinical work (National Research Council [U.S.] Committee on Human Genome Diversity 1997). Large-scale genomic studies sampling from human populations distributed worldwide have produced detailed knowledge on variation in disease resistance or susceptibility between and within populations. Let’s think about drug companies who develop medicines for Black patients particularly. The predispositions to particular diseases are higher among people of African descent than some pharmaceutical businesses have taken into account. Through targeted sampling of various world groups, clinical geneticists may also identify genetic risk factors of certain common disorders such as chronic heart disease, asthma, diabetes, autoimmune diseases, and behavioral disorders. Having an understanding of population-specific biology is crucial in the development of therapies, medicines, and vaccinations, as not all treatments may be suitable for every human, depending on their genotype. During diagnosis and treatment, it is important to have an evolutionary perspective on gene-environment relationships in patients. Typological concepts of “race” are not useful, given that most racial groups (whether self-identified or not) popularly recognized lack homogeneity and are, in fact, variable. Cystic fibrosis, for instance, occurs in all world populations but can often be underdiagnosed in populations with African ancestry because it is thought of as a “white” disease (Yudell et al. 2016).
Sociologists, law scholars, and professors of race studies have written extensively on how genetic/technological/medical revolutions impact people of color. In her book, Fatal Invention: How Science, Politics, and Big Business Re-create Race in the Twenty-First Century (2013), Professor Dorothy E. Roberts writes about how technological advances have been used in resuscitating race as a biological category for dividing humans in essentialist ways (Figure 13.23). She notes how members of law enforcement have engaged in racial profiling, sometimes with the use of machine-learning and facial-recognition technologies. Ancestry-testing services also purport to tell us “what” we are and to insist that this information is “written” in our genes. Such advertising campaigns obscure the nuances of genetic variation with the primary motive of tapping into people’s desire to “know themselves” and driving up profits for their businesses. Commercial genetic testing reinforces the idea that genes map neatly onto race, all while generating massive stores of data in DNA databases. In Roberts’s view, the myth of the biological concept of race being perpetuated in these ways undermines a just society and reproduces racial inequalities.
The COVID-19 pandemic has had a significant impact on the world’s population, particularly people living in the economic Global South and many Black, Indigenous and communities of color residing in the Global North. We have witnessed disproportionately high numbers of COVID-related deaths and infection cases among marginalized groups. Many immigrants and ethnic minorities in various societies have also experienced scapegoating and blame directed at them for being the source of COVID-19 spread.
To inform us on how to interpret this current worldwide pandemic, historians and anthropologists are looking back at the lessons learned from past instances of racist medicine (discriminatory practices based on broader social discrimination) and medical racism (application of discriminatory practices justified on medical grounds). Historically, who could become doctors and medical professionals was often racialized, gendered, and class specific. This made it difficult for many to overcome prejudices against women, Black people, Indigenous individuals, or other people of color from becoming doctors and clinical researchers in places such as South Africa and the United States. This, in turn, affects the sorts of information we know about health levels and health outcomes among these very groups. In the past decade, long-overdue attention is finally being paid to how race affects biological outcomes. For instance, researchers have focused on the negative legacies of racial discrimination and racism-induced stress on hormone (im)balances, mental health disorders, cardiovascular disease prevalence, and other health outcomes (Kuzawa and Sweet 2009; Shonkoff, Slopen, and WIlliams 2021; Williams 2018). The technology and standards of protocol in medical testing have been scrutinized (for more on how pulse oximeters were not designed with nonwhite patients in mind, for example, see Sjoding et al. 2020). Scholars of race and medicine have also written on how illness and disease spread have often been used to perpetuate societal prejudices. This manifests as xenophobic tendencies at a societal level, such as the blaming of “outgroups” and increased “in-group” protectiveness. Overreliance on the idea that people are “inherently” disease carriers due to genetic or biological reasons leads to improper accounting for socioeconomic or infrastructural issues that lead to differential disease prevalence amongst minority communities. (For more on race and COVID, see Tsai 2021 as well as this textbook’s Chapter 16: Contemporary Topics: Human Biology and Health.)
Lastly, consider the changing field of forensic anthropology. In the past, forensic anthropologists ascribed ancestry or racial categories to sets of skeletons, reliant on the belief that different human groups will exhibit biologically “discrete” assortments so as to divide along culturally constructed categories (Sauer 1992). Now, a number of forensic anthropologists have argued that we should abandon these methods, both because it is unscientific and because it further validates and perpetuates this idea that race is biologically meaningful. As scientists, whether we affirm biological race as real has huge influence on the beliefs of members of the public, the judicolegal system, and law enforcement. Not all forensic experts agree with abandoning ancestry estimation. Some prefer to refocus on the neutral or selective causes of human biological variation, and assess how probabilistic it may be to assign bones of certain dimensions to one of several identified racial categories. These debates continue today as this textbook chapter is being written. More details on population affinity may be found in Chapter 15: Bioarchaeology and Forensic Anthropology.
It is important to remember that while it is possible to look for clues about one’s ancestry or geographic origin based on skull morphology, again, the amount of distinctiveness in any given sample makes it impossible to distinguish whether a cranium belongs to one group (Relethford 2009). Individuals can vary in their skeletal dimensions by continental origin, country origin, regional origin, sex, age, environmental factors, and the time period in which they lived, making it difficult to assign individuals to particular categories in a completely meaningful way (Ousley, Jantz, and Freid 2009). When forensic reports and scientific journal articles give an estimation of ancestry, it is crucial to keep in mind that responsible assignments of ancestry will be done through robust statistical testing and stated as a probability estimate. Today, we also live in a more globalized world where a skeletal individual may have been born originally to parents of two separate traditional racial categories. In contexts of great heterogeneity within populations, this definitely adds difficulty to the work of forensic scientists and anthropologists preparing results for the courtroom (genetic testing may be comparatively more helpful in such situations).
Did Deeper: Measuring FST
Richard Lewontin (1929‒) is a biologist and evolutionary geneticist who authored an article evaluating where the total genetic variation in humans lies. Titled “The Apportionment of Human Diversity” (Lewontin 1972), the article addressed the following question: On average, how genetically similar are two randomly chosen people from the same group when compared to two randomly chosen people from different groups?
Lewontin studied this problem by using genetic data. He obtained data for a large number of different human populations worldwide using 17 genetic markers (including alleles that code for various important enzymes and proteins, such as blood-group proteins). The statistical analysis he ran used a measure of human genetic differences in and among populations known as the fixation index (FST).
Technically, FST can be defined as the proportion of total genetic variance within a subpopulation relative to the total genetic variance from an entire population. Therefore, FST values range from 0 to 1 (or, sometimes you will see this stated as a percentage between 0% and 100%). The closer the FST value of a population (e.g., the world’s population) approaches 1, the higher the degree of genetic differentiation among subpopulations relative to the overall population (see Figure 13.24 for a detailed illustration).
In his article, Lewontin (1972) identified that most of human genetic differences (85.4%) were found within local subpopulations (e.g., the Germans or Easter Islanders), whereas 8.3% were found between populations within continental human groups, and 6.3% were attributable to traditional “race” groups (e.g., “Caucasian” or “Amerind”). These findings have been important for scientifically rejecting the existence of biological races (Long and Kittles 2003).
Talking About Human Biological Variation Going Forward
To conclude, utilizing the term races to describe human biological variation is not accurate or productive. Using a select few hundred genetic loci, or perhaps a number of phenotypic traits, it may be possible to assign individuals to a geographic ancestry, but what constitutes a bounded genetic or geographical grouping is both arbitrary and potentially harmful owing to ethical and historical reasons. The discipline of biological anthropology has moved past typological frameworks that shoehorn continuously variable human populations into discrete and socially constructed subsets. Improvements in the number of markers, the genetic technologies used to study variation, and the number of worldwide populations sampled have led to more nuanced understandings of human variation. It is of utmost importance that scientists make the following points clear to the public:
- Today, we refer to different local human groups as “populations.” What constitutes a population should be carefully defined in scientific reports based on some geographical, linguistic, or cultural criteria and some degree of relativity to other closely or distantly related human groups.
- Humans have significantly less genetic variation than other primates and mammals, and all human beings on Earth share 99.9% of their overall DNA. Some of the remaining 0.1% of human variation varies on a clinal or continuous basis, such as can be seen when looking at ABO blood-type polymorphisms worldwide.
- Many biological characteristics in humans are actually determined nonconcordantly and/or polygenically. Therefore, superiority or inferiority in human behavior or body form cannot justifiably be linked to fixed and innate differences between groups.
- Genetic distances are correlated with geographic distances among the global human population. This is especially apparent when we consider that genetic variation is highest in sub-Saharan Africa, and average genetic heterogeneity decreases in populations further away from the African continent in accordance with the migratory history of anatomically modern Homo sapiens.
- The effects of gene flow, genetic drift, and population bottlenecking are reflected in some phenotypic traits, such as cranial shape.
- We recognize other traits, like skin color and lactase persistence, to be the products of many millennia of natural selective pressures influencing human biology from the external environment.
Taken together, genetic analyses of human variation do not support 20th-century (or even-earlier) concepts of race. In discussions about human variation, these genomic results help clarify how biological variation is distributed across the human population today. Taking care to think about and debate the nature of human variation is important, because although the effects and events that produced genetic differences among groups occurred in the ancient past, sociocultural concepts about race and ethnicity continue to have real social, economic, and political consequences.
Beyond talking about variation in the university setting, it is important that teachers, researchers, and students of anthropology recognize and assume the responsibility of influencing public perspectives of human variation. Race-based classification systems were developed during the colonial era, the transatlantic trafficking of kidnapped Africans and the so-called “Scientific Revolution” by the first “anthropologists” and scholars of humankind’s variation. Unfortunately, some of their early ideas have persisted and evolved into present-day lived realities. Some of today’s politicians and socioeconomic bodies have racially charged agendas that promote racism or certain kinds of economic or racial inequalities. As anthropologists, we must acknowledge that while human “races” are not a biological reality, their status as a (misguided) social construction does have real consequences for many people (Antrosio 2011).
In other words, while “race” is a sociocultural invention, the treatment different individuals receive due to their perceived “race” can have significant financial, emotional, sociopolitical, and physiological costs. However—and importantly assuming a “color-blind” position when it comes to the topics of “race” and ethnicity (especially in political discussions) is actually counterproductive, because the negative social consequences of modern “race” ideas could be ignored, making it harder to examine and address instances of discrimination properly (Wise 2010). Rather than shy away from these topics, we can use our scientific findings to establish socially relevant and biologically accurate ideas concerning human diversity. Today, research into genetic and phenotypic differentiation among and within various human populations continues to expand in its scope, its technological capabilities, its sample sizes, and its ethical concerns. It is thanks to such scientific work done in the past few decades that we now have a deeper understanding not only of how humans vary but also of how we are biologically a rather homogenous, intermixing world population.
Review Questions
- How is the genetic variation of the human species distributed worldwide?
- What evolutionary processes are responsible for producing genotypic/phenotypic variation within and between human populations?
- Should we continue to attribute any value to “race” concepts older than 1950, based on our current understandings of human biological variation?
- How should we communicate scientific findings about human biological variation more accurately and responsibly to those outside the anthropological discipline?
Key Terms
Age of Discovery: A period between the late 1400s and late 1700s when European explorers and ships sailed extensively across the globe in pursuit of new trading routes and territorial conquest.
Ancestry: Biogeographical information about an individual, traced either through the study of an individual’s genome, skeletal characteristics, or some other form of forensic/archaeological evidence. Anthropologists carry out probabilistic estimates of ancestry. They attribute sets of human remains to distinctive “ancestral” groups using careful statistical testing and should report ancestry estimations with statistical probability values.
Binomial nomenclature: A system of naming living things developed by Linnaeus in the 1700s. It employs a scientific name made up of two italicized Latin or Greek words, with the first word capitalized and representative of an organism’s genus and the second word indicating an organism’s species (e.g., Homo sapiens, Australopithecus afarensis, Pongo tapanuliensis, etc.).
Biological anthropology: A branch of study under anthropology (the study of humankind) that focuses on when and where humans and our human ancestors first originated, how we have evolved and adapted globally over time, and the reasons why we see biological variation among humans worldwide today.
Biological determinism: The erroneous concept that an individual’s behavioral characteristics are innate and determined by genes, brain size, or other physiological attributes—and, notably, without the influence of social learning or the environment around the individual during development.
Bony labyrinth: A system of interconnected canals within the auditory (ear- or hearing-related) apparatus, located in the inner ear and responsible for balance and the reception of sound waves.
Cline: A gradient of physiological or morphological change in a single character or allele frequency among a group of species across environmental or geographical lines (e.g., skin color varies clinally, as, over many generations, human groups living nearer the equator have adapted to have more skin pigmentation).
Continuous variation: This term refers to variation that exists between individuals and cannot be measured using distinct categories. Instead, differences between individuals within a population in relation to one particular trait are measurable along a smooth, continuous gradient.
Cystic fibrosis: A genetic disorder in which one defective gene causes overproduction and buildup of mucus in the lungs and other bodily organs. It is most common in northern Europeans (but also occurs in other world populations).
Ecological niche: The position or status of an organism within its community and/or ecosystem, resulting from the organism’s structural and functional adaptations (e.g., bipedalism, omnivorous diets, lactose digestion, etc.).
Essentialism: A belief or view that an entity, organism, or human grouping has a specific set of characteristics that are fundamentally necessary to its being and classification into definitive categories.
Ethnicity: A term used commonly in an interchangeable way with the term race, complicated because how different people define this term depends on the qualities and characteristics they use to assign a label or identity to themselves and/or others (which may include aspects of family background, skin color, language(s) spoken, religion, physical proportions, behavior and temperament, etc.).
Eugenics: A set of beliefs and practices that involves the controlled selective breeding of human populations with the hope of improving their heritable qualities, especially through surgical procedures like sterilization and legal rulings that affect marriage rights for interracial couples.
Gene flow: A neutral (or nonselective) evolutionary process that occurs when genes get shared between populations.
Genetic drift: A neutral evolutionary process in which allele frequencies change from generation to generation due to random chance.
Heterogeneity: The quality of being diverse genetically.
Homogenous: The quality of being uniform genetically.
Human diversity: Human diversity is a measure of variation that may describe how many different forms of human there are, separated or clustered into groups according to some genetic, phenotypic, or cultural trait(s). The term can be applied to culture (in which case humans can be described as significantly diverse) or genetics (in which case humans are not diverse because all humans on Earth share a majority of their genes).
Human variation: Differences in biology, physiology, body chemistry, behavior, and culture. By measuring these differences, we understand the degrees of variation between individuals, groups, populations, or species.
Isolation-by-distance model: A model that predicts a positive relationship between genetic distances and geographical distances between pairs of populations.
Monogenetic: Pertaining to the idea that the origin of a species is situated in one geographic region or time (as opposed to polygenetic).
Mutation: A gene alteration in the DNA sequence of an organism. As a random, neutral evolutionary process that occurs over the course of meiosis and early cell development, gene mutations are possible sources of variation in any given human gene pool. Genetic mutations that occur in more than 1% of a population are termed polymorphisms.
Natural selection: An evolutionary process whereby certain traits are perpetuated through successive generations, likely owing to the advantages they give organisms in terms of chances of survival and/or reproduction.
Nonconcordance: The fact of genes or traits not varying with one another and instead being inherited independently.
Otherness: In postcolonial anthropology, we now understand “othering” to mean any action by someone or some group that establishes a division between “us” and “them” in relation to other individuals or populations. This could be based on linguistic or cultural differences, and it has largely been based on external characteristics throughout history.
Out-of-Africa model: A model that suggests that all humans originate from one single group of Homo sapiens in (sub-Saharan) Africa who lived between 100,000 and 315,000 years ago and who subsequently diverged and migrated to other regions across the globe.
Physical anthropology: This used to be the more common name given to the subdiscipline of anthropology centered upon the study of human origins, evolution and variation (also see biological anthropology above). This name for the field has gradually become less popular due to two reasons: first, it may not reflect our interests in other aspects of humankind that are not physical (such as those behavioral, cultural and spiritual), and second, using this term popular in the early decades of our field may be viewed by some as harkening back to a time when biological anthropologists conducted their work in unethical ways.
Polygenetic: Having many different ancestries, as in older theories about human origins that involved multiple traditional groupings of humans evolving concurrently in different parts of the world before they merged into one species through interbreeding and/or intergroup warfare. These earlier suggestions have now been overwhelmed by insurmountable evidence for a single origin of the human species in Africa (see the “Out-of-Africa model”).
Polymorphism: A genetic variant within a population (caused either by a single gene or multiple genes) that occurs at a rate of over 1% among the population. Polymorphisms are responsible for variation in phenotypic traits such as blood type and skin color.
Population: A group of humans living in a particular geographical area, with more local interbreeding within-group than interbreeding with other groups. A limited or restricted amount of gene flow between populations can occur due to geographical, cultural, linguistic, or environmental factors.
Population bottlenecking: An event in which genetic variation is significantly reduced owing to a sharp reduction in population size. This can occur when environmental disaster strikes or as a result of human activities (e.g., genocides or group migrations). An important example of this loss in genetic variation occurred over the first human migrations out of Africa and into other continental regions.
Prejudice: An unjustified attitude toward an individual or group that is not based on reason, whether positive (and showing preference for one group of people over another) or negative (and resulting in harm or injury to others).
Race: The identification of a group based on a perceived distinctiveness that makes that group more similar to each other than they are to others outside the group. This may be based on cultural differences, genetic parentage, physical characteristics, behavioral attributes, or something arbitrarily and socially constructed. As a social or demographic category, perceptions of “race” can have real and serious consequences for different groups of people. This is despite the fact that biological anthropologists and geneticists have demonstrated that all humans are genetically homogenous and that more differences can be found within populations than between them in the overall apportionment of human biological variation. This term is sometimes used interchangeably with ethnicity.
Racism: Any action or belief that discriminates against someone based on perceived differences in race or ethnicity.
Scientific Revolution: A period between the 1400s and 1600s when substantial shifts occurred in the social, technological, and philosophical sense, when a scientific method based on the collection of empirical evidence through experimentation was emphasized and inductive reasoning was used to test hypotheses and interpret their results.
Typological: Of or describing an assortment system that relies on the interpretation of qualitative similarities or differences in the study of variation among objects or people. The categorization of cultures or human groups according to “race” was performed with a typological approach in the earliest practice of anthropology, but this practice has since been discredited and abandoned.
Variance: In statistics, variance measures the dispersal of a set of data around the mean or average value.
About the Author
Michael B. C. Rivera, Ph.D.
University of Hong Kong, mrivera@hku.hk
Michael B. C. Rivera is a biological anthropologist and human bioarchaeologist who studies human evolution and history and works to develop these disciplines in Hong Kong, East/Southeast Asia, and the “Global South.” His doctoral thesis focused on the transition into agriculture in coastal environments and adaptations of ancient people along the beach. He is the only biological anthropologist working at the University of Hong Kong and the lead archaeologist managing the excavation of a WWII military aircraft that crashed in Hong Kong in 1945. Michael is also an advocate for greater inclusion, diversity, equality, and access to learning in academia. Much of his work also includes science communication and public engagement activities online, in schools, and in collaboration with museums.
For Further Exploration
Videos
American Medical Association (AMA). 2020. “Examining Race-Based Medicine.” YouTube, October 29. Accessed June 4, 2023.
Crenshaw, Kimberlé. 2016. “The Urgency of Intersectionality.” YouTube, December 7. Accessed June 4, 2023.
Golash-Boza, Tanya. 2018. “What Is Race? What Is Ethnicity? Is There a Difference?.” YouTube, October 28. Accessed June 4, 2023.
Lasisi, Tina. 2020. “How Hair Reveals the Futility of Race Categories.” National Museum of Natural History webinar, October 21.
Lasisi, Tina. 2022. “Where Does My Skin Color Come From?.” PBS Terra, August 18. Accessed June 4, 2023.
PBS Origins. 2018. “The Origin of Race in the USA.” YouTube, April 3. Accessed June 4, 2023.
Roberts, Dorothy. 2016. “The Problem with Race-Based Medicine.” YouTube, March 4. Accessed June 4, 2023.
Vox. 2015. “The Myth of Race, Debunked in 3 Minutes.” YouTube, January 13. Accessed June 4, 2023.
Podcast Episodes
Kwong, Emily, and Rebecca Ramirez. 2021. “Here’s a Better Way to Talk about Hair: A 16 Minute Listen with Tina, Lasisi” NPR Short Wave, October 6. Accessed June 4, 2023.
Speaking of Race. 2020. “Race and Health series.” Speaking of Race, April 10. Accessed June 4, 2023.
Websites
Choices Program. 2023. “An Interactive Timeline: Black Activism and the Long Fight for Racial Justice.” Choices Program, Brown University [Interactive Timeline], Updated February, 2023.
References
American Association of Biological Anthropologists. 2020. “An Open Letter to Our Community in Response to Police Brutality against African-Americans and a Call to Antiracist Action”. American Association of Biological Anthropologists, June 10, 2020. Accessed June 4, 2023.
Antrosio, Jason. 2011. “‘Race Reconciled’: Race Isn’t Skin Color, Biology, or Genetics.” Living Anthropologically (website), June 5, 2011; updated May 20, 2020. Accessed June 4, 2023.
Beals, Kenneth L., Courtland L. Smith, Stephen M. Dodd, J. Lawrence Angel, Este Armstrong, Bennett Blumenberg, Fakhry G. Girgis, et al. 1984. “Brain Size, Cranial Morphology, Climate, and Time Machines [and Comments and Reply].” Current Anthropology 25 (3): 301‒330.
Betti, Lia, François Balloux, Tsunehiko Hanihara, and Andrea Manica. 2010. “The Relative Role of Drift and Selection in Shaping the Human Skull.” American Journal of Physical Anthropology 141 (1): 76‒82. https://doi.org/10.1002/ajpa.21115.
Betti, Lia, François Balloux, William Amos, Tsunehiko Hanihara, and Andrea Manica. 2009. “Distance from Africa, Not Climate, Explains Within-Population Phenotypic Diversity in Humans.” Proceedings: Biological Sciences 276 (1658): 809‒814. https://doi.org/10.1098/rspb.2008.1563.
Betti, Lia, Noreen von Cramon-Taubadel, Andrea Manica, and Stephen J. Lycett. 2013. “Global Geometric Morphometric Analyses of the Human Pelvis Reveal Substantial Neutral Population History Effects, Even across Sexes.” PloS ONE 8 (2): e55909. https://doi.org/10.1371/journal.pone.0055909.
Betti, Lia, Noreen von Cramon-Taubadel, Andrea Manica, and Stephen J. Lycett. 2014. “The Interaction of Neutral Evolutionary Processes with Climatically Driven Adaptive Changes in the 3D Shape of the Human Os Coxae.” Journal of Human Evolution 73 (August): 64‒74. https://doi.org/10.1016/j.jhevol.2014.02.021.
Boas, Franz. 1931. “Race and Progress.” Science 74 1905): 1‒8.
Bowden, Rory, Tammie S. MacFie, Simon Myers, Garrett Hellenthal, Eric Nerrienet, Ronald E. Bontrop, Colin Freeman, Peter Donnelly, and Nicholas I. Mundy. 2012. “Genomic Tools for Evolution and Conservation in the Chimpanzee: Pan troglodytes ellioti Is a Genetically Distinct Population.” PLoS Genetics 8 (3): e1002504. https://doi.org/10.1371/journal.pgen.1002504.
Campbell, Michael C., and Sarah A. Tishkoff. 2008. “African Genetic Diversity: Implications for Human Demographic History, Modern Human Origins, and Complex Disease Mapping.” Annual Review of Genomics and Human Genetics 9: 403‒433.
Clee, Paul R. Sesink, Ekwoge E. Abwe, Ruffin D. Ambahe, Nicola M. Anthony, Roger Forso, Sabrina Locatelli, Fiona Maisels, et al. 2015. “Chimpanzee Population Structure in Cameroon and Nigeria Is Associated with Habitat Variation That May Be Lost Under Climate Change.” BMC Evolutionary Biology 15: 2. https://doi.org/10.1186/s12862-014-0275-z.
Cullors, Patrisse. 2016. “An Interview with the Founders of Black Lives Matter.” TED Talks 2016, October 26‒28. Accessed June 15, 2023. https://www.ted.com/talks/alicia_garza_patrisse_cullors_and_opal_tometi_an_interview_with_the_founders_of_black_lives_matter/up-next.
Fuentes, Agustín, Rebecca Rogers Ackermann, Sheela Athreya, Deborah Bolnick, Tina Lasisi, Sang-Hee Lee, Shay-Akil McLean, and Robin Nelson. 2019. “AAPA Statement on Race and Racism.” American Journal of Physical Anthropology 169 (3): 400‒402.
Garza, Alicia. 2016. “An Interview with the Founders of Black Lives Matter.” TED Talks 2016, October 26‒28. Accessed June 15, 2023. https://www.ted.com/talks/alicia_garza_patrisse_cullors_and_opal_tometi_an_interview_with_the_founders_of_black_lives_matter/up-next.
Gerbault, Pascale, Anke Liebert, Yuval Itan, Adam Powell, Mathias Currat, Joachim Burger, Dallas M. Swallow, and Mark G. Thomas. 2011. “Evolution of Lactase Persistence: An Example of Human Niche Construction.” Philosophical Transactions of the Royal Society B 366 (1566): 863‒877. https://doi.org/10.1098/rstb.2010.0268.
Hooton, Earnest A. 1936. “Plain Statements about Race.” Science 83 (2161): 511‒513.
Hrdlička, Aleš. 1918. “Physical Anthropology: Its Scope and Aims; Its History and Present Status in America. A: Physical Anthropology; Its Scopes and Aims.” American Journal of Physical Anthropology 1 (1): 3‒23.
Huxley, Julian. 1942. Evolution: The Modern Synthesis. London: Allen and Unwin.
Ingram, Catherine J. E., Charlotte A. Mulcare, Yuval Itan, Mark G. Thomas, and Dallas M. Swallow. 2009. “Lactose Digestion and the Evolutionary Genetics of Lactase Persistence.” Human Genetics 124 (6): 579‒591. https://doi.org/10.1007/s00439-008-0593-6.
Jablonski, Nina G. 2004. “The Evolution of Human Skin and Skin Color.” Annual Review of Anthropology 33: 585‒623. https://doi.org/10.1146/annurev.anthro.33.070203.143955.
Jablonski, Nina G., and George Chaplin. 2000. “The Evolution of Human Skin Coloration.” Journal of Human Evolution 39 (1): 57‒106. https://doi.org/10.1006/jhev.2000.0403.
Kanitz, Ricardo, Elsa G. Guillot, Sylvain Antoniazza, Samuel Neuenschwander, and Jérôme Gedout. 2018. “Complex Genetic Patterns in Human Arise from a Simple Range-Expansion Model over Continental Landmasses.” PLoS ONE 13 (2): e0192460.
Kronenberg, Zev N., Ian T. Fiddes, David Gordon, Shwetha Murali, Stuart Cantsilieris, Olivia S. Meyerson, Jason G. Underwood, et al. 2018. “High-Resolution Comparative Analysis of Great Ape Genomes.” Science 360 (6393): eaar6343. https://doi.org/10.1126/science.aar6343.
Kuzawa, Christopher W., and Elizabeth Sweet. 2009. “Epigenetics and the Embodiment of Race: Development Origins of US Racial Disparities in Cardiovascular Health.” American Journal of Human Biology 21 (1) : 2‒15.
Lasisi, Tina, and Mark D. Shriver. 2018. “Focus on African Diversity Confirms Complexity of Skin Pigmentation Genetics.” Genomic Biology 19: 13.
Lewontin, Richard. 1972. “The Apportionment of Human Diversity.” In Evolutionary Biology, vol. 6, edited by Theodosius Dobzhansky, Max K. Hecht, and William C. Steere, 381‒398. New York: Springer.
Linnaeus, Carl. 1758. Systema Naturae. Stockholm: Laurentius Salvius. https://www.cabdirect.org/abstracts/20057000018.html.
Liu, Hua, Franck Prugnolle, Andrea Manica, and François Balloux. 2006. “A Geographically Explicit Genetic Model of Worldwide Human-Settlement History.” American Journal of Human Genetics 79 (2): 230‒237.
Livingstone, Frank B. 1962. “On the Nonexistence of Human Races.” Current Anthropology 3 (3): 279‒281.
Long, Jeffery C., and Rick A. Kittles. 2003. “Human Genetic Diversity and the Nonexistence of Biological Races.” Human Biology 75 (4): 449‒471.
Luzzatto, Lucio. 2012. “Sickle Cell Anaemia and Malaria.” Mediterranean Journal of Hematology and Infectious Diseases 4 (1). https://doi.org/10.4084/MJHID.2012.065.
Manica, Andrea, William Amos, François Balloux, and Tsunehiko Hanihara. 2007. “The Effect of Ancient Population Bottlenecks on Human Phenotypic Variation.” Nature 448 (7151): 346‒348. https://doi.org/10.1038/nature05951.
McLean, Shay-Akil. 2014. “‘Race, Ethnicity, & Racism.” Decolonize ALL The Things Website, Accessed January 10, 2023. https://decolonizeallthethings.com/learning-tools/race-ethnicity-racism/.
Morton, Samuel George. 1839. Crania Americana, or, A Comparative View of the Skulls of Various Aboriginal Nations of North and South America. Philadelphia: J. Dobson.
Mourant, A. E., Ada C. Kopeć, and Kazimiera Domaniewska-Sobczak. 1976. The Distribution of the Human Blood Groups and Other Polymorphisms, 2nd edition. Oxford: Oxford University Press.
National Research Council (U.S.) Committee on Human Genome Diversity. 1997. Evaluating Human Genetic Diversity. Washington, D.C.: National Academies Press.
Omi, Michael, and Howard Winant. 2014. “The Theory of Racial Formation.” In Racial Formation in the United States,3rd edition, edited by Michael Omi and Howard Winant, 105‒126. Routledge: New York.
Osada, Naoki. 2015. “Genetic Diversity in Humans and Non-Human Primates and Its Evolutionary Consequences.” Genes and Genetic Systems 90 (3): 133‒145.
Ousley, Stephen D., Richard L. Jantz, and Donna Freid. 2009. “Understanding Race and Human Variation: Why Forensic Anthropologists Are Good at Identifying Race.” American Journal of Physical Anthropology 139 (1): 68‒76. https://doi.org/10.1002/ajpa.21006.
Ponce de León, Marcia S., Toetik Koesbardiati, John David Weissmann, Marco Millela, Carlos S. Reyna-Blanco, Gen Suwa, Osamu Kondo, Anna-Sapfo Malaspinas, Tim D. White, and Christoph P. E. Zollikofer. 2018. “Human Bony Labyrinth Is an Indicator of Population History and Dispersal from Africa.” Proceedings of the National Academy of Sciences 115 (16): 4128‒4133. https://doi.org/10.1073/pnas.1808125115.
Prado-Martinez, Javier, Peter H. Sudmant, Jeffrey M. Kidd, Heng Li, Joanna L. Kelley, Belen Lorente-Galdos, Krishna R. Veeramah, et al. 2013. “Great Ape Genetic Diversity and Population History.” Nature 499 (7459): 471–475. https://doi.org/10.1038/nature12228.
Prugnolle, Franck, Andrea Manica, and François Balloux. 2005. “Geography Predicts Neutral Genetic Diversity of Human Populations.” Current Biology 15 (5): 159‒160.
Quillen, Ellen E., Heather L. Norton, Esteban J. Parra, Frida Lona-Durazo, Khai C. Ang, Florin Mircea Illiescu, Laurel N. Pearson, et al. 2018. “Shades of Complexity: New Perspectives on the Evolution and Genetic Architecture of Human Skin.” American Journal of Physical Anthropology 168 (S67): 4–26.
Rathmann, Hannes, Hugo Reyes-Centeno, Silvia Ghirotto, Nicole Creanza, Tsunehiko Hanihara, and Katerina Harvati. 2017. “Reconstructing Human Population History from Dental Phenotypes.” Scientific Reports 7: 12495. https://doi.org/10.1038/s41598-017-12621-y.
Relethford, John H. 2001. “Global Analysis of Regional Differences in Craniometric Diversity and Population Substructure.” Human Biology 73 (5): 629‒636. https://doi.org/10.1353/hub.2001.0073.
Relethford, John H. 2002. “Apportionment of Global Human Genetic Diversity Based on Craniometrics and Skin Color.” American Journal of Physical Anthropology 118 (4): 393‒398. https://doi.org/10.1002/ajpa.10079.
Relethford, John H. 2004. “Global Patterns of Isolation by Distance Based on Genetic and Morphological Data.” Human Biology 76 (4): 499‒513. https://doi.org/10.1353/hub.2004.0060.
Relethford, John H. 2009. “Race and Global Patterns of Phenotypic Variation.” American Journal of Physical Anthropology 139 (1): 16‒22. https://doi.org/10.1002/ajpa.20900.
Roberts, Dorothy. 2013. Fatal Invention: How Science, Politics, and Big Business Re-Create Race in the Twenty-First Century. New York: The New Press.
Rosenberg, Noah A., Saurabh Mahajan, Sohini Ramachandran, Chengfeng Zhao, Jonathan K. Pritchard, and Marcus W. Feldman. 2005. “Clines, Clusters, and the Effect of Study Design on the Inference of Human Population Structure.” PLoS Genetics 1 (6): e70. https://doi.org/10.1371 /journal.pgen.0010070.
Rosenberg, Noah A., Jonathan K. Pritchard, James L. Weber, Howard M. Cann, Kenneth K. Kidd, Lev A. Zhivotovsky, and Marcus W. Feldman. 2002. “Genetic Structure of Human Populations.” Science 298 (5602): 2381‒2385.
Sauer, Norman J. 1992. “Forensic Anthropology and the Concept of Race: If Races Don’t Exist, Why Are Forensic Anthropologists So Good at Identifying Them?” Social Science and Medicine 34 (2): 107‒111. https://doi.org/10.1016/0277-9536(92)90086-6.
Shonkoff, Jack P., Natalie Slopen, and David R. Williams. 2021. “Early Childhood Adversity, Toxic Stress, and the Impacts of Racism on the Foundations of Health.” Annual Review of Public Health 42: 115‒134.
Sjoding, Michael W., Robert P. Dickson, Theodore J. Iwashyna, Steven E. Gay, and Thomas S. Valley. 2020. “Racial Bias in Pulse Oximetry Measurement.” The New England Journal of Medicine 383: 2477-2478.
Staes, Nicky, Chet C. Sherwood, Katharine Wright, Marc de Manuel, Elaine E. Guevara, Tomas Marques-Bonet, Michael Krützen, et al. 2017. “FOXP2 Variation in Great Ape Populations Offers Insight into the Evolution of Communication Skills.” Scientific Reports 7 (1): 1‒10. https://doi.org/10.1038/s41598-017-16844-x.
Tomati, Opal. 2016. “An Interview with the Founders of Black Lives Matter.” TED Talks 2016, October 26‒28. Accessed June 15, 2023. https://www.ted.com/talks/alicia_garza_patrisse_cullors_and_opal_tometi_an_interview_with_the_founders_of_black_lives_matter/up-next.
Tsai, Jennifer. 2021. “COVID-19 Is Not a Story of Race, but a Record of Racism—Our Scholarship Should Reflect That Reality.” The American Journal of Bioethics 21 (2): 43‒47. https://doi.org/10.1080/15265161.2020.1861377.
von Cramon-Taubadel, Noreen, and Stephen J. Lycett. 2008. “Brief Communication: Human Cranial Variation Fits Iterative Founder Effect Model with African Origin.” American Journal of Physical Anthropology 136 (1): 108‒113. https://doi.org/10.1002/ajpa.20775.
Weiss, Kenneth M., and Jeffrey C. Long. 2009. “Non-Darwinian Estimation: My Ancestors, My Genes’ Ancestors.” Genome Research 19: 703‒710. https://doi.org/10.1101/gr.076539.108.19.
Williams, David W. 2018. “Stress and the Mental Health of Populations of Color: Advancing Our Understanding of Race-related Stressors.” Journal of Health and Social Behavior 59 (4): 466‒485.
Wise, Tim. 2010. Colorblind: The Rise of Post-Racial Politics and the Retreat from Racial Equity. San Francisco: City Lights.
Yudell, Michael, Dorothy Roberts, Rob DeSalle, and Sarah Tishkoff. 2016. “Taking Race out of Human Genetics.” Science 351 (6273): 564‒565. https://doi.org/10.1126/science.aac4951.
Single-celled marine organisms with shells.
Hayley Mann, M.A., Binghamton University
This chapter is a revision from "Chapter 3: Molecular Biology and Genetics" by Hayley Mann, Xazmin Lowman, and Malaina Gaddis. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Learning Objectives
- Explain and identify the purpose of both DNA replication and the cell cycle.
- Identify key differences between mitosis and meiosis.
- Outline the process of protein synthesis, including transcription and translation.
- Use principles of Mendelian inheritance to predict genotypes and phenotypes of future generations.
- Explain complexities surrounding patterns of genetic inheritance and polygenic traits.
- Discuss challenges to and bioethical concerns of genetic testing.
I [Hayley Mann] started my Bachelor’s degree in 2003, which was the same year the Human Genome Project released its first draft sequence. I initially declared a genetics major because I thought it sounded cool. However, upon taking an actual class, I discovered that genetics was challenging. In addition to my genetics major, I signed up for biological anthropology classes and soon learned that anthropology could bring all those molecular lessons to life. For instance, we are composed of cells, proteins, nucleic acids, carbohydrates, and lipids. Anthropologists often include these molecules in their studies to identify how humans vary; if there are meaningful differences, they propose theories to explain them. Anthropologists study biomolecules in both living and ancient individuals. Ancient biomolecules can also be found on artifacts such as stone tools and cooking vessels. Over the years, scientific techniques for studying organic molecules have improved, which has unlocked new insights into the deep human past.
This chapter provides the basics for understanding human variation and how the evolutionary process works. A few advanced genetics topics are also presented because biotechnology is now commonplace in health and society. Understanding the science behind this remarkable field means you will be able to participate in bioethical and anthropological discussions as well as make more informed decisions regarding genetic testing.
Cells and Molecules
Molecules of Life
All organisms are composed of four basic types of molecules that are essential for cell structure and function: proteins, lipids, carbohydrates, and nucleic acids (Figure 3.1). Proteins are crucial for cell shape and nearly all cellular tasks, including receiving signals from outside the cell and mobilizing intra-cellular responses. Lipids are a class of organic compounds that include fats, oils, and hormones. As discussed later in the chapter, lipids are also responsible for the characteristic phospholipid bilayer structure of the cell membrane. Carbohydrates are sugar molecules and serve as energy to cells in the form of glucose. Lastly, nucleic acids, including deoxyribonucleic acid (DNA), carry genetic information about a living organism.
Molecule |
Definition |
Example |
Proteins |
Composed of one or more long chains of amino acids (i.e., basic units of protein) Often folded into complex 3D shapes that relate to function Proteins interact with other types of proteins and molecules |
Proteins come in different categories including structural (e.g., collagen, keratin, lactase, hemoglobin, cell membrane proteins), defense proteins (e.g, antibodies), enzymes (e.g., lactase), hormones (e.g., insulin), and motor proteins (e.g., actin) |
Lipids |
Insoluble in water due to hydrophilic (water-loving) head and a hydrophobic (water-repelling) tail |
Fats, such as triglycerides, store energy for your body Steroid hormones (e.g., estrogen and testosterone) act as chemical messengers to communicate between cells and tissues, as well as biochemical pathways inside of the cell |
Carbohydrates |
Large group of organic molecules that are composed of carbon and hydrogen atoms |
Starches and sugars, including blood glucose, provide cells with energy |
Nucleic Acids |
Carries the genetic information of an organism |
DNA RNA |
Cells
In 1665, Robert Hooke observed slices of plant cork using a microscope. Hooke noted that the microscopic plant structures he saw resembled cella, meaning “a small room” in Latin. Approximately two centuries later, biologists recognized the cell as being the most fundamental unit of life and that all life is composed of cells. Cellular organisms can be characterized as two main cell types: prokaryotes and eukaryotes (Figure 3.2).
Prokaryotes include bacteria and archaea, and they are composed of a single cell. Additionally, their DNA and organelles are not surrounded by individual membranes. Thus, no compartments separate their DNA from the rest of the cell (see Figure 3.2). It is well known that some bacteria can cause illness in humans. For instance, Escherichia coli (E. coli) and Salmonella contamination can result in food poisoning symptoms. Pneumonia and strep throat are caused by Streptococcal bacteria. Neisseria gonorrhoeae is a sexually transmitted bacterial disease. Although bacteria are commonly associated with illness, not all bacteria are harmful. For example, researchers are studying the relationship between the microbiome and human health. The bacteria that are part of the healthy human microbiome perform beneficial roles, such as digesting food, boosting the immune system, and even making vitamins (e.g., B12 and K).
Eukaryotes can be single-celled or multi-celled in their body composition. In contrast to prokaryotes, eukaryotes possess membranes that surround their DNA and organelles. An example of a single-celled eukaryote is the microscopic algae found in ponds (phytoplankton), which can produce oxygen from the sun. Yeasts are also single-celled, and fungi can be single- or multicellular. Plants and animals are all multicellular.
Although plant and animal cells have a surprising number of similarities, there are some key differences (Figure 3.3). For example, plant cells possess a thick outer cell membrane made of a fibrous carbohydrate called cellulose. Animal and plant cells also have different tissues. A tissue is an aggregation of cells that are morphologically similar and perform the same task. For most plants, the outermost layer of cells forms a waxy cuticle that helps to protect the cells and to prevent water loss. Humans have skin, which is the outermost cell layer that is predominantly composed of a tough protein called keratin. Overall, humans have a diversity of tissue types (e.g., cartilage, brain, and heart).
Animal Cell Organelles
An animal cell is surrounded by a double membrane called the phospholipid bilayer (Figure 3.4). A closer look reveals that this protective barrier is made of lipids and proteins that provide structure and function for cellular activities, such as regulating the passage of molecules and ions (e.g., H2O and sodium) into and out of the cell. Cytoplasm is the jelly-like matrix inside of the cell membrane. Part of the cytoplasm comprises organelles, which perform different specialized tasks for the cell (Figure 3.5). An example of an organelle is the nucleus, where the cell’s DNA is located.
Another organelle is the mitochondrion. Mitochondria are often referred to as “powerhouse centers” because they produce energy for the cell in the form of adenosine triphosphate (ATP). Depending on the species and tissue type, multicellular eukaryotes can have hundreds to thousands of mitochondria in each of their cells. Scientists have determined that mitochondria were once symbiotic prokaryotic organisms (i.e., helpful bacteria) that transformed into cellular organelles over time. This evolutionary explanation helps explain why mitochondria also have their own DNA, called mitochondrial DNA (mtDNA). All organelles have important physiological functions and disease can occur when organelles do not perform their role optimally. Figure 3.6 lists other organelles found in the cell and their specialized cellular roles.
Cell structure |
Description |
Centrioles |
Assist with the organization of mitotic spindles, which extend and contract for the purpose of cellular movement during mitosis and meiosis. |
Cytoplasm |
Gelatinous fluid located inside of cell membrane that contains organelles. |
Endoplasmic reticulum (ER) |
Continuous membrane with the nucleus that helps transport, synthesize, modify, and fold proteins. Rough ER has embedded ribosomes, whereas smooth ER lacks ribosomes. |
Golgi body |
Layers of flattened sacs that receive and transmit messages from the ER to secrete and transport proteins within the cell. |
Lysosome |
Located in the cytoplasm; contains enzymes to degrade cellular components. |
Microtubule |
Involved with cellular movement including intracellular transport and cell division. |
Mitochondrion |
Responsible for cellular respiration, where energy is produced by converting nutrients into ATP. |
Nucleolus |
Resides inside of the nucleus and is the site of ribosomal RNA (rRNA) transcription, processing, and assembly. |
Nucleopore |
Pores in the nuclear envelope that are selectively permeable. |
Nucleus |
Contains the cell’s DNA and is surrounded by the nuclear envelope. |
Ribosome |
Located in the cytoplasm and also the membrane of the rough endoplasmic reticulum. Messenger RNA (mRNA) binds to ribosomes and proteins are synthesized. |
Introduction to Genetics
Genetics is the study of heredity. Biological parents pass down their genetic traits to their offspring. Although children resemble their parents, genetic traits often vary in appearance or molecular function. For example, two parents with normal color vision can sometimes produce a son with red-green colorblindness. Patterns of genetic inheritance will be discussed in a later section. Molecular geneticists study the biological mechanisms responsible for creating variation between individuals, such as DNA mutations (see Chapter 4), cell division, and genetic regulation.
Molecular anthropologists use genetic data to test anthropological questions. Some of these anthropologists utilize ancient DNA (aDNA), which is DNA that is extracted from anything once living, including human, animal, and plant remains. Over time, DNA becomes degraded (i.e., less intact), but specialized laboratory techniques can make copies of short degraded aDNA segments, which can then be reassembled to provide more complete DNA information. A recent example of an aDNA study is provided in Special Topic: Native American Immunity and European Diseases, and aDNA is also explored in Appendix D.
DNA Structure
The discovery, in 1953, of the molecular structure of deoxyribonucleic acid (DNA) was one of the greatest scientific achievements of all time. Using X-ray crystallography, Rosalind Franklin (Figure 3.7) provided the image that clearly showed the double helix shape of DNA. However, due to a great deal of controversy, Franklin’s colleague and outside associates received greater publicity for the discovery. In 1962, James Watson, Francis Crick, and Maurice Wilkins received a Nobel Prize for developing a biochemical model of DNA. Unfortunately, Rosalind Franklin had passed away in 1958 from ovarian cancer. In current times, Franklin’s important contribution and her reputation as a skilled scientist are widely acknowledged.
The double helix shape of DNA can be described as a twisted ladder (Figure 3.8). More specifically, DNA is a double-stranded molecule with its two strands oriented in opposite directions (i.e., antiparallel). Each strand is composed of nucleotides with a sugar phosphate backbone. There are four different types of DNA nucleotides: adenine (A), thymine (T), cytosine (C), and guanine (G). The two DNA strands are held together by nucleotide base pairs, which have chemical bonding rules. The complementary base-pairing rules are as follows: A and T bond with each other, while C and G form a bond. The chemical bonds between A-T and C-G are formed by “weak” hydrogen atom interactions, which means the two strands can be easily separated. A DNA sequence is the order of nucleotide bases (A, T, G, C) along only one DNA strand. If one DNA strand has the sequence CATGCT, then the other strand will have a complementary sequence GTACGA. This is an example of a short DNA sequence. In reality, there are approximately three billion DNA base pairs in human cells.
DNA Is Highly Organized within the Nucleus
If you removed the DNA from a single human cell and stretched it out completely, it would measure approximately two meters (about 6.5 feet). Therefore, DNA molecules must be compactly organized in the nucleus. To achieve this, the double helix configuration of DNA undergoes coiling. An analogy would be twisting a string until coils are formed and then continuing to twist so that secondary coils are formed, and so on. To assist with coiling, DNA is first wrapped around proteins called histones. This creates a complex called chromatin, which resembles “beads on a string” (Figure 3.9). Next, chromatin is further coiled into a chromosome. Another important feature of DNA is that chromosomes can be altered from tightly coiled (chromatin) to loosely coiled (euchromatin). Most of the time, chromosomes in the nucleus remain in a euchromatin state so that DNA sequences are accessible for regulatory processes to occur.
Human body cells typically have 23 pairs of chromosomes, for a total of 46 chromosomes in each cell’s nucleus. An interesting fact is that the number of chromosomes an organism possesses varies by species, and this figure is not dependent upon the size or complexity of the organism. For instance, chimpanzees have a total of 48 chromosomes, while hermit crabs have 254. Chromosomes also have a distinct physical structure, including centromeres (the “center”) and telomeres (the ends) (Figure 3.10). Because of the centromeric region, chromosomes are described as having two different “arms,” where one arm is long and the other is shorter. Centromeres play an important role during cell division, which will be discussed in the next section. Telomeres are located at the ends of chromosomes; they help protect the chromosomes from degradation after every round of cell division.
Special Topic: Native American Immunity and European Diseases—A Study of Ancient DNA
Beginning in the early fifteenth century, Native Americans progressively suffered from high mortality rates as the result of colonization from foreign powers. European-borne diseases such as measles, tuberculosis, influenza, and smallpox are largely responsible for the population collapse of indigenous peoples in the Americas. Many Europeans who immigrated to the Americas had lived in large sedentary populations, which also included coexisting with domestic animals and pests. Although a few prehistoric Native American populations can be characterized as large agricultural societies (especially in Mesoamerica), their overall culture, community lifestyle, and subsistence practices were markedly different from that of Europeans. Therefore, because they did not share the same urban living environments as Europeans, it is believed that Native Americans were susceptible to many European diseases.
In 2016, a Nature article published by John Lindo and colleagues was the first to investigate whether pre-contact Native Americans possessed a genetic susceptibility to European diseases. Their study included Tsimshians, a First Nation community from British Columbia (Figure 3.11a-b). DNA from both present-day and ancient individuals (who lived between 500 and 6,000 years ago) was analyzed. The research team discovered that a change occurred in the HLA-DQA1 gene, which is a member of the major histocompatibility complex (MHC) immune system molecules. MHC molecules are responsible for detecting and triggering an immune response against pathogens. Lindo and colleagues (2016) concluded that HLA-DQA1 gene helped Native Americans adapt to their local environmental ecology. However, when European-borne epidemics occurred in the Northwest during the 1800s, a certain HLA-DQA1 DNA sequence variant (allele) associated with ancient Tsimshian immunity was no longer adaptive. As the result of past selective pressures from European diseases, present-day Tsimshians have different HLA-DQA1 allele frequencies. The precise role that HLA-DQA1 plays in immune adaptation requires further investigation. But overall, this study serves as an example of how studying ancient DNA from the remains of deceased individuals can help provide insight into living human populations and historical events.
DNA Replication and Cell Division
For life to continue and flourish, cells must be able to divide. Tissue growth and cellular damage repair are also necessary to maintain an organism throughout its life. All these rely on the dynamic processes of DNA replication and the cell cycle. The mechanisms highlighted in this section are tightly regulated and represent only part of the life cycle of a cell.
DNA Replication
DNA replication is the process by which new DNA is copied from an original DNA template. It is one phase of the highly coordinated cell cycle, and it requires a variety of enzymes with special functions. The creation of a complementary DNA strand from a template strand is described as semi-conservative replication. The result of semi-conservative replication is two separate double-stranded DNA molecules, each of which is composed of an original “parent” template strand and a newly synthesized “daughter” DNA strand.
DNA replication progresses in three steps referred to as initiation, elongation, and termination. During initiation, enzymes are recruited to specific sites along the DNA sequence (Figure 3.12). For example, an initiator enzyme, called helicase, “unwinds” DNA by breaking the hydrogen bonds between the two parent strands. The unraveling of the helix into two separated strands exposes the strands and creates a fork, which is the active site of DNA replication.
Elongation is the assembly of new DNA daughter strands from the exposed original parent strands. The two parent strands can further be classified as leading strand or lagging strand and are distinguished by the direction of replication. Enzymes called DNA polymerases read parent template strands in a specific direction. Complementary nucleotides are added, and the newly formed daughter strands will grow. On the leading parent strand, a DNA polymerase will create one continuous strand. The lagging parent strand is created in several disconnected sections and other enzymes fill in the missing nucleotide gaps between these sections.
Finally, termination refers to the end of DNA replication activity. It is signaled by a stop sequence in the DNA that is recognized by machinery at the replication fork. The end result of DNA replication is that the number of chromosomes are doubled so that the cell can divide into two.
DNA Mutations
DNA replication should result in the creation of two identical DNA nucleotide sequences. However, although DNA polymerases are quite precise during DNA replication, copying mistakes are estimated to occur every 107 DNA nucleotides. Variation from the original DNA sequence is known as a mutation. The different types of mutations will be discussed in greater detail in Chapter 4. Briefly, mutations can result in single nucleotide changes, as well as the insertion or deletion of nucleotides and repeated sequences. Depending on where they occur in the genome, mutations can be deleterious (harmful). For example, mutations may occur in regions that control cell cycle regulation, which can result in cancer (see Special Topic: The Cell Cycle and Immortality of Cancer Cells). Many other types of mutations, however, are not harmful to an organism.
Regardless of their effect, the cell attempts to reduce the frequency of mutations that occur during DNA replication. To accomplish this, there are polymerases with proofreading capacities that can identify and correct mismatched nucleotides. These safeguards reduce the frequency of DNA mutations so that they only occur every 109 nucleotides.
Mitotic Cell Division
There are two types of cells in the body: germ cells (sperm and egg) and somatic cells. The body and its various tissues comprises somatic cells. Organisms that contain two sets of chromosomes in their somatic cells are called diploid organisms. Humans have 46 chromosomes and they are diploid because they inherit one set of chromosomes (n = 23) from each parent. As a result, they have 23 matching pairs of chromosomes, which are known as homologous chromosomes. As seen in Figure 3.13, homologous chromosome pairs vary in size and are generally numbered from largest (chromosome 1) to smallest (chromosome 22) with the exception of the 23rd pair, which is made up of the sex chromosomes (X and Y). Typically, the female sex is XX and the male sex is XY. Individuals inherit an X chromosome from their chromosomal mother and an X or Y from their chromosomal father.
To grow and repair tissues, somatic cells must divide. As discussed previously, for cell division to occur, a cell must first replicate its genetic material. During DNA replication, each chromosome produces double the amount of genetic information. The duplicated arms of chromosomes are known as sister chromatids, and they are attached at the centromeric region. To elaborate, the number of chromosomes stays the same (n = 46); however, the amount of genetic material is doubled in the cell as the result of replication.
Mitosis is the process of somatic cell division that gives rise to two diploid daughter cells. Figure 3.14 includes a brief overview of mitosis. Once DNA and other organelles in the cell have finished replication, mitotic spindle fibers physically align each chromosome at the center of the cell. Next, the spindle fibers divide the sister chromatids and move each one to opposite sides of the cell. At this phase, there are 46 chromosomes on each side of a human cell. The cell can now divide into two fully separated daughter cells.
Meiotic Cell Division
Gametogenesis is the production of gametes (sperm and egg cells); it involves two rounds of cell division called meiosis. Similar to mitosis, the parent cell in meiosis is diploid. However, meiosis has a few key differences, including the number of daughter cells produced (four cells, which require two rounds of cell division to produce) and the number of chromosomes each daughter cell has (see Figure 3.14).
During the first round of division (known as meiosis I), each chromosome (n = 46) replicates its DNA so that sister chromatids are formed. Next, with the help of spindle fibers, homologous chromosomes align near the center of the cell and sister chromatids physically swap genetic material. In other words, the sister chromatids of matching chromosomes cross over with each other at matching DNA nucleotide positions. The occurrence of homologous chromosomes crossing over, swapping DNA, and then rejoining segments is called genetic recombination. The “genetic shuffling” that occurs in gametes increases organismal genetic diversity by creating new combinations of genes on chromosomes that are different from the parent cell. Genetic mutations can also arise during recombination. For example, there may be an unequal swapping of genetic material that occurs between the two sister chromatids, which can result in deletions or duplications of DNA nucleotides. Once genetic recombination is complete, homologous chromosomes are separated and two daughter cells are formed.
The daughter cells after the first round of meiosis are haploid, meaning they only have one set of chromosomes (n = 23). During the second round of cell division (known as meiosis II), sister chromatids are separated and two additional haploid daughter cells are formed. Therefore, the four resulting daughter cells have one set of chromosomes (n = 23), and they also have a genetic composition that is not identical to the parent cells nor to each other.
Although both sperm and egg gamete production undergo meiosis, they differ in the final number of viable daughter cells. In the case of spermatogenesis, four mature sperm cells are produced. Although four egg cells are also produced in oogenesis, only one of these egg cells will result in an ovum (mature egg). During fertilization, an egg cell and sperm cell fuse, which creates a diploid cell that develops into an embryo. The ovum also provides the cellular organelles necessary for embryonic cell division. This includes mitochondria, which is why humans, and most other multicellular eukaryotes, have the same mtDNA sequence as their mothers.
Chromosomal Disorders: Aneuploidies
During mitosis or meiosis, entire deletions or duplications of chromosomes can occur due to error. For example, homologous chromosomes may fail to separate properly, so one daughter cell may end up with an extra chromosome while the other daughter cell has one less. Cells with an unexpected (or abnormal) number of chromosomes are known as aneuploid. Adult or embryonic cells can be tested for chromosome number (karyotyping). Aneuploid cells are typically detrimental to a dividing cell or developing embryo, which can lead to a loss of pregnancy. However, the occurrence of individuals being born with three copies of the 21st chromosome is relatively common; this genetic condition is known as Down Syndrome. Moreover, individuals can also be born with aneuploid sex chromosome conditions such as XXY, XXX, and XO (referring to only one X chromosome).
Special Topic: The Cell Cycle and Immortality of Cancer Cells
DNA replication is part of a series of preparatory phases that a cell undergoes prior to cell division, collectively known as interphase (Figure 3.15). During interphase, the cell not only doubles its chromosomes through DNA replication, but it also increases its metabolic capacity to provide energy for growth and division. Transition into each phase of the cell cycle is tightly controlled by proteins that serve as checkpoints. If a cell fails to pass a checkpoint, then DNA replication and/or cell division will not continue. Some of the reasons why a cell may fail at a checkpoint is DNA damage, lack of nutrients to continue the process, or insufficient size. In turn, a cell may undergo apoptosis, which is a mechanism for cell death.
Unchecked cellular growth is a distinguishing hallmark of cancer. In other words, as cancer cells grow and proliferate, they acquire the capacity to avoid death and replicate indefinitely. This uncontrolled and continuous cell division is also known as “immortality.” As previously mentioned, most cells lose the ability to divide due to shortening of telomeres on the ends of chromosomes over time. One way in which cancer cells retain replicative immortality is that the length of their telomeres is continuously protected. Chemotherapy, often used to treat cancer, targets the cell cycle (especially cell division) to halt the propagation of genetically abnormal cells. Another therapeutic approach that continues to be investigated is targeting telomere activity to stop the division of cancer cells.
Researchers have exploited the immortality of cancer cells for molecular research. The oldest immortal cell line is HeLa cells (Figure 3.16), which were harvested from Henrietta Lacks, an African American woman diagnosed with cervical cancer in 1955. At that time, extracted cells frequently died during experiments, but surprisingly HeLa cells continued to replicate. Propagation of Lacks’s cell line has significantly contributed to medical research, including contributing to ongoing cancer research and helping to test the polio vaccine in the 1950s. However, Lacks had not given her consent for her tumor biopsy to be used in cell culture research. Moreover, her family was unaware of the extraction and remarkable application of her cells for two decades. The history of HeLa cell origin was first revealed in 1976. The controversy voiced by the Lacks family was included in an extensive account of HeLa cells published in Rebecca Skloot’s 2010 book, The Immortal Life of Henrietta Lacks. A film based on the book was also released in 2017 (Wolfe 2017).
Protein Synthesis
At the beginning of the chapter, we defined proteins as strings of amino acids that fold into complex 3-D shapes. There are 20 standard amino acids that can be strung together in different combinations in humans, and the result is that proteins can perform an impressive amount of different functions. For instance, muscle fibers are proteins that help facilitate movement. A special class of proteins (immunoglobulins) help protect the organism by detecting disease-causing pathogens in the body. Protein hormones, such as insulin, help regulate physiological activity. Blood hemoglobin is a protein that transports oxygen throughout the body. Enzymes are also proteins, and they are catalysts for biochemical reactions that occur in the cell (e.g., metabolism). Larger-scale protein structures can be visibly seen as physical features of an organism (e.g., hair and nails).
Transcription and Translation
Nucleotides in our DNA provide the coding instructions on how to make proteins. Making proteins, also known as protein synthesis, can be broken down into two main steps referred to as transcription and translation. The purpose of transcription, the first step, is to make an ribonucleic acid (RNA) copy of our genetic code. Although there are many different types of RNA molecules that have a variety of functions within the cell, we will mainly focus on messenger RNA (mRNA). Transcription concludes with the processing (splicing) of the mRNA. The second step, translation, uses mRNA as the instructions for chaining together amino acids into a new protein molecule (Figure 3.17).
Unlike double-stranded DNA, RNA molecules are single-stranded nucleotide sequences (Figure 3.18). Additionally, while DNA contains the nucleotide thymine (T), RNA does not—instead its fourth nucleotide is uracil (U). Uracil is complementary to (or can pair with) adenine (A), while cytosine (C) and guanine (G) continue to be complementary to each other.
For transcription to proceed, a gene must first be turned “on” by the cell. A gene is a segment of DNA that codes for RNA, and genes can vary in length from a few hundred to as many as two million base pairs in length. The double-stranded DNA is then separated, and one side of the DNA is used as a coding template that is read by RNA polymerase. Next, complementary free-floating RNA nucleotides are linked together (Figure 3.19) to form a single-stranded mRNA. For example, if a DNA template is TACGGATGC, then the newly constructed mRNA sequence will be AUGCCUACG.
Genes contain segments called introns and exons. Exons are considered “coding” while introns are considered “noncoding”—meaning the information they contain will not be needed to construct proteins. When a gene is first transcribed into pre-mRNA, introns and exons are both included (Figure 3.20). However, once transcription is finished, introns are removed in a process called splicing. During splicing, a protein/RNA complex attaches itself to the pre-mRNA. Next, introns are removed and the remaining exons are connected, thus creating a shorter mature mRNA that serves as a template for building proteins.
As described above, the result of transcription is a single-stranded mRNA copy of a gene. Translation is the process by which amino acids are chained together to form a new protein. During translation, the mature mRNA is transported outside of the nucleus, where it is bound to a ribosome (Figure 3.21). The nucleotides in the mRNA are read in triplets, which are called codons. Each mRNA codon corresponds to an amino acid, which is carried to the ribosome by a transfer RNA (tRNA). Thus, tRNAs is the link between the mRNA molecule and the growing amino acid chain.
Continuing with our mRNA sequence example from above, the mRNA sequence AUG-CCU-ACG codes for three amino acids. Using a codon table (Figure 3.22), AUG is a codon for methionine (Met), CCU is proline (Pro), and ACG is threonine (Thr). Therefore, the protein sequence is Met-Pro-Thr. Methionine is the most common “start codon” (AUG) for the initiation of protein translation in eukaryotes. As the ribosome moves along the mRNA, the growing amino acid chain exits the ribosome and folds into a protein. When the ribosome reaches a “stop” codon (UAA, UAG, or UGA), the ribosome stops adding any new amino acids, detaches from the mRNA, and the protein is released. Depending upon the amino acid sequence, a linear protein may undergo additional “folding.” The final three-dimensional protein shape is integral to completing a specific structural or functional task.
Dig Deeper: Protein Synthesis
To see protein synthesis in animation, please check out the From DNA to Protein video on YourGenome.org.
Special Topic: Genetic Regulation of the Lactase (LCT) Gene (chapter 14)
The LCT gene codes for a protein called lactase, an enzyme produced in the small intestine. It is responsible for breaking down the sugar “lactose,” which is found in milk. Lactose intolerance occurs when not enough lactase enzyme is produced and, in turn, digestive symptoms occur. To avoid this discomfort, individuals may take lactase supplements, drink lactose-free milk, or avoid milk products altogether.
The LCT gene is a good example of how cells regulate protein synthesis. The promoter region of the LCT gene helps regulate whether it is transcribed or not transcribed (i.e., turned “on” or “off,” respectively). Lactase production is initiated when a regulatory protein known as a transcription factor binds to a site on the LCT promoter. RNA polymerases are then recruited; they read DNA and string together nucleotides to make RNA molecules. An LCT pre-mRNA is synthesized (made) in the nucleus, and further chemical modifications flank the ends of the mRNA to ensure the molecule will not be degraded in the cell. Next, a spliceosome complex removes the introns from the LCT pre-mRNA and connects the exons to form a mature mRNA. Translation of the LCT mRNA occurs and the growing protein then folds into the lactase enzyme, which can break down lactose.
Most animals lose their ability to digest milk as they mature due to the decreasing transcriptional “silence” of the LCT gene over time. However, some humans have the ability to digest lactose into adulthood (also known as “lactase persistence”). This means they have a genetic mutation that leads to continuous transcriptional activity of LCT. Lactase persistence mutations are common in populations with a long history of pastoral farming, such as northern European and North African populations. It is believed that lactase persistence evolved because the ability to digest milk was nutritionally beneficial. More information about lactase persistence will be covered in Chapter 14.
Mendelian Genetics
Gregor Johann Mendel (1822–1884) is often described as the “Father of Genetics.” Mendel was a monk who conducted pea plant breeding experiments in a monastery located in the present-day Czech Republic (Figure 3.23). After several years of experiments, Mendel presented his work to a local scientific community in 1865 and published his findings the following year. Although his meticulous effort was notable, the importance of his work was not recognized for another 35 years. One reason for this delay in recognition is that his findings did not agree with the predominant scientific viewpoints on inheritance at the time. For example, it was believed that parental physical traits “blended” together and offspring inherited an intermediate form of that trait. In contrast, Mendel showed that certain pea plant physical traits (e.g., flower color) were passed down separately to the next generation in a statistically predictable manner. Mendel also observed that some parental traits disappeared in offspring but then reappeared in later generations. He explained this occurrence by introducing the concept of “dominant” and “recessive” traits. Mendel established a few fundamental laws of inheritance, and this section reviews some of these concepts. Moreover, the study of traits and diseases that are controlled by a single gene is commonly referred to as Mendelian genetics.
The physical appearance of a trait is called an organism’s phenotype. Figure 3.24 shows pea plant (Pisum sativum) phenotypes that were studied by Mendel, and in each of these cases the physical traits are controlled by a single gene. In the case of Mendelian genetics, a phenotype is determined by an organism’s genotype. A genotype consists of two gene copies, wherein one copy was inherited from each parent. Gene copies are also known as alleles (Figure 3.25), which means they are found in the same gene location on homologous chromosomes. Alleles have a nonidentical DNA sequence, which means their phenotypic effect can be different. In other words, although alleles code for the same trait, different phenotypes can be produced depending on which two alleles (i.e., genotypes) an organism possesses. For example, Mendel’s pea plants all have flowers, but their flower color can be purple or white. Flower color is therefore dependent upon which two color alleles are present in a genotype.
A Punnett square is a diagram that can help visualize Mendelian inheritance patterns. For instance, when parents of known genotypes mate, a Punnett square can help predict the ratio of Mendelian genotypes and phenotypes that their offspring would possess. When discussing genotype, biologists use upper and lower case letters to denote the different allele copies. Figure 3.26 is a Punnett square that includes two heterozygous parents for flower color (Bb). A heterozygous genotype means there are two different alleles for the same gene. Therefore, a pea plant that is heterozygous for flower color has one purple allele and one white allele. When an organism is homozygous for a specific trait, it means their genotype consists of two copies of the same allele. Using the Punnett square example, the two heterozygous pea plant parents can produce offspring with two different homozygous genotypes (BB or bb) or offspring that are heterozygous (Bb).
A pea plant with purple flowers could be heterozygous (Bb) or homozygous (BB). This is because the purple color allele (B) is dominant to the white color allele (b), and therefore it only needs one copy of that allele to phenotypically express purple flowers. Because the white flower allele is recessive, a pea plant must be homozygous for the recessive allele in order to have a white color phenotype (bb). As seen by the Punnett square example (Figure 3.26), three of four offspring will have purple flowers and the other one will have white flowers.
The Law of Segregation was introduced by Mendel to explain why we can predict the ratio of genotypes and phenotypes in offspring. As discussed previously, a parent will have two alleles for a certain gene (with each copy on a different homologous chromosome). The Law of Segregation states that the two copies will be segregated from each other and will each be distributed to their own gamete. We now know that the process where that occurs is meiosis.
Offspring are the products of two gametes combining, which means the offspring inherits one allele from each gamete for most genes. When multiple offspring are produced (like with pea plant breeding), the predicted phenotype ratios are more clearly observed. The pea plants Mendel studied provide a simplistic model to understand single-gene genetics. While many traits anthropologists are interested in have a more complicated inheritance (e.g., are informed by many genes), there are a few known Mendelian traits in humans. Additionally, some human diseases also follow a Mendelian pattern of inheritance (Figure 3.27). Because humans do not have as many offspring as other organisms, we may not recognize Mendelian patterns as easily. However, understanding these principles and being able to calculate the probability that an offspring will have a Mendelian phenotype is still important.
Mendelian disorder | Gene |
Alpha Thalassemia | HBA1 |
Cystic Fibrosis | CFTR |
Fragile X Syndrome | FMR1 |
Glucose-6-Phosphate Dehydrogenase Deficiency | G6PD |
Hemophilia A | F8 |
Huntington disease | HTT |
Mitochondrial DNA Depletion Syndrome | TYMP |
Oculocutaneous Albinism: Type 1 | TYR |
Polycystic Kidney Disease | PKHD1 |
Sickle-cell anemia | HBB |
Spinal Muscular Atrophy: SMN1 Linked | SMN1 |
Tay-Sachs Disease | HEXA |
Wilson Disease | ATP7B |
Example of Mendelian Inheritance: The ABO Blood Group System
In 1901, Karl Landsteiner at the University of Vienna published his discovery of ABO blood groups. While conducting blood immunology experiments in which he combined the blood of individuals who possess different blood cell types, he observed an agglutination (clotting) reaction. The presence of agglutination implies there is an incompatible immunological reaction; no agglutination will occur in individuals with the same blood type. This work was clearly important because it resulted in a higher survival rate of patients who received blood transfusions. Blood transfusions from someone with a different type of blood causes agglutinations, and the resulting coagulated blood can not easily pass through blood vessels, resulting in death. Landsteiner received the Nobel Prize (1930) for his discovery and explaination of the ABO blood group system.
Blood cell surface antigens are proteins that coat the surface of red blood cells, and antibodies are specifically “against” or “anti” to the antigens from other blood types. Thus, antibodies are responsible for causing agglutination between incompatible blood types. Understanding the interaction of antigens and antibodies helps to determine ABO compatibility amongst blood donors and recipients. To better comprehend blood phenotypes and ABO compatibility, blood cell antigens and plasma antibodies are presented in Figure 3.28. Individuals that are blood type A have A antigens on the red blood cell surface, and anti-B antibodies, which will bind to B antigens should they come in contact. Alternatively, individuals with blood type B have B antigens and anti-A antibodies. Individuals with blood type AB have both A and B antigens but do not produce antibodies for the ABO system. This does not mean type AB does not have any antibodies present, just that specifically anti-A and anti-B antibodies are not produced. Individuals who are blood type O have nonspecific antigens and produce both anti-A and anti-B antibodies.
Figure 3.29 shows a table of the ABO allele system, which has a Mendelian pattern of inheritance. Both the A and B alleles function as dominant alleles, so the A allele always codes for the A antigen, and the B allele codes for the B antigen. The O allele differs from A and B, because it codes for a nonfunctional antigen protein, which means there is no antigen present on the cell surface of O blood cells. To have blood type O, two copies of the O allele must be inherited, one from each parent, thus the O allele is considered recessive. Therefore, someone who is a heterozygous AO genotype is phenotypically blood type A, and a genotype of BO is blood type B. The ABO blood system also provides an example of codominance, which is when both alleles are observed in the phenotype. This is true for blood type AB: when an individual inherits both the A and B alleles, then both A and B antigens will be present on the cell surface.
Also found on the surface of red blood cells is the rhesus group antigen, known as “Rh factor.” In reality, there are several antigens on red blood cells independent from the ABO blood system, however, the Rh factor is the second most important antigen to consider when determining blood donor and recipient compatibility. Rh antigens must also be considered when a pregnant mother and her baby have incompatible Rh factors. In such cases, a doctor can administer necessary treatment steps to prevent pregnancy complications and hemolytic disease, which is when the mother’s antibodies break down the newborn’s red blood cells.
An individual can possess the Rh antigen (be Rh positive) or lack the Rh antigen (be Rh negative). The Rh factor is controlled by a single gene and is inherited independently of the ABO alleles. Therefore, all blood types can either be positive (O+, A+, B+, AB+) or negative (O-, A-, B-, AB-).
Individuals with O+ red blood cells can donate blood to A+, B+, AB+, and O+ blood type recipients. Because O- individuals do not have AB or Rh antigens, they are compatible with all blood cell types and are referred to as “universal donors.” Individuals that are AB+ are considered to be “universal recipients” because they do not possess antibodies against other blood types.
Mendelian Patterns of Inheritance and Pedigrees
A pedigree can be used to investigate a family’s medical history by determining if a health issue is inheritable and will possibly require medical intervention. A pedigree can also help determine if it is a Mendelian recessive or dominant genetic condition. Figure 3.30 is a pedigree example of a family with Huntington’s disease, which has a Mendelian dominant pattern of inheritance. In a standard pedigree, males are represented by a square and females are represented by a circle. Biological family members are connected to a horizontal line, with biological parents above and offspring below. When an individual is affected with a certain condition, the square or circle is filled in as a solid color. With a dominant condition, at least one of the parents will have the disease and an offspring will have a 50% chance of inheriting the affected chromosome. Therefore, dominant genetic conditions tend to be present in every generation. In the case of Huntington’s, some individuals may not be diagnosed until later in adulthood, so parents may unknowingly pass this dominantly inherited disease to their children.
Because the probability of inheriting a disease-causing recessive allele is more rare, recessive medical conditions can skip generations. Figure 3.31 is an example of a family that carries a recessive cystic fibrosis mutation. A parent that is heterozygous for the cystic fibrosis allele has a 50% chance of passing down their affected chromosome to the next generation. If a child has a recessive disease, then it means both of their parents are carriers (heterozygous) for that condition. In most cases, carriers for recessive conditions show no serious medical symptoms. Individuals whose family have a known medical history for certain conditions sometimes seek family planning services (see the Genetic Testing section).
Pedigrees can also help distinguish if a health issue has either an autosomal or X-linked pattern of inheritance. As previously discussed, there are 23 pairs of chromosomes and 22 of these pairs are known as autosomes. The provided pedigree examples (Figure 3.30–31) are autosomally linked genetic diseases. This means the genes that cause the disease are on one of the chromosomes numbered 1 to 22. The conditions caused by genes located on the X chromosome are referred to as X-linked diseases.
Figure 3.32 depicts a family in which the mother is a carrier for the X-linked recessive disease Duchenne Muscular Dystrophy (DMD). The mother is a carrier for DMD, so daughters and sons will have a 50% chance of inheriting the pathogenic DMD allele. Because females have two X chromosomes, females who inherit only one copy will not have the disease (although in rare cases, female carriers may show some symptoms of the disease). On the other hand, males who inherit a copy of an X-linked pathogenic DMD allele will typically be affected with the condition. Thus, males are more susceptible to X-linked conditions because they only have one X chromosome. Therefore, when evaluating a pedigree, if a higher proportion of males are affected with the disease, this could suggest the disease is X-linked recessive.
Compared to the X chromosome, the Y chromosome is smaller with only a few genes. Y-linked traits are therefore rare and can only be passed from a chromosomal father to a biological XY child.
Other Patterns of Inheritance
Complexity Surrounding Mendelian Inheritance
Pea plant trait genetics are relatively simple compared to what we know about genetic inheritance today. The vast majority of genetically controlled traits are not strictly dominant or recessive, so the relationship among alleles and predicting phenotype is often more complicated. For example, traits that exhibit incomplete dominance occur when a heterozygote exhibits a phenotype that is an intermediate phenotype of both alleles. In snapdragon flowers, the red flower color (R) is dominant and white is recessive (r). Therefore, the homozygous dominant RR is red and homozygous recessive rr is white. However, because the R allele is not completely dominant, the heterozygote Rr is a blend of red and white, which results in a pink flower (Figure 3.33).
An example of incomplete dominance in humans is the enzyme β-hexosaminidase A (Hex A), which is encoded by the gene HEXA. Patients with two dysfunctional HEXA alleles are unable to metabolize a specific lipid-sugar molecule (GM2 ganglioside); because of this, the molecule builds up and causes damage to nerve cells in the brain and spinal cord. This condition is known as Tay-Sachs disease, and it usually appears in infants who are three to six months old. Most children with Tay-Sachs do not live past early childhood. Individuals who are heterozygous for the functional type HEXA allele and one dysfunctional allele have reduced Hex A activity. However, the amount of enzyme activity is still sufficient, so carriers do not exhibit any neurological phenotypes and appear healthy.
Some genes and alleles can also have higher penetrance than others. Penetrance can be defined as the proportion of individuals who have a certain allele and also express an expected phenotype. If a genotype always produces an expected phenotype, then those alleles are said to be fully penetrant. However, in the case of incomplete (or reduced) penetrance, an expected phenotype may not occur even if an individual possesses the alleles that are known to control a trait or cause a disease.
A well-studied example of genetic penetrance is the cancer-related genes BRCA1 and BRCA2. Mutations in these genes can affect crucial processes such as DNA repair, which can lead to breast and ovarian cancers. Although BRCA1 and BRCA2 mutations have an autosomal dominant pattern of inheritance, it does not mean an individual will develop cancer if they inherit a pathogenic allele. Several lifestyle and environmental factors can also influence the risk for developing cancer. Regardless, if a family has a history of certain types of cancers, then it is often recommended that genetic testing be performed for individuals who are at risk. Moreover, publically available genetic testing companies are now offering health reports that include BRCA1 and BRCA2 allele testing (see the Genetic Testing section).
Polygenic Traits
While Mendelian traits tend to be influenced by a single gene, the vast majority of human phenotypes are polygenic traits. The term polygenic means “many genes.” Therefore, a polygenic trait is influenced by many genes that work together to produce the phenotype. Human phenotypes such as hair color, eye color, height, and weight are examples of polygenic traits. Hair color, for example, is largely determined by the type and quantity of a pigment called melanin, which is produced by a specialized cell type within the skin called melanocytes.The quantity and ratio of melanin pigments determine black, brown, blond, and red hair colors. MC1R is a well-studied gene that encodes a protein expressed on the surface of melanocytes that is involved in the production of eumelanin pigment. Typically, people with two functional copies of MC1R have brown hair. People with reduced functioning MC1R allele copies tend to produce pheomelanin, which results in blond or red hair. However, MC1R alleles have variable penetrance, and studies are continually identifying new genes (e.g., TYR, TYRP1, SLC24A5, and KITLG) that also influence hair color. Individuals with two nonfunctioning copies of the gene TYR have a condition called oculocutaneous albinism—their melanocytes are unable to produce melanin so these individuals have white hair, light eyes, and pale skin.
In comparison to Mendelian diseases, complex diseases (e.g., Type II diabetes, coronary heart disease, Alzheimer's, and schizophrenia) are more prevalent in humans. Complex diseases are polygenic, but their development is also influenced by physical, environmental, sociocultural, and individual lifestyle factors. Families can be more predisposed to certain diseases; however, complex diseases often do not have a clear pattern of inheritance.
Although research of complex traits and diseases continue, geneticists may not know all of the genes involved with a given complex disease. Additionally, how much genetic versus nongenetic determinants contribute to a disease phenotype can be difficult to decipher. Therefore, predicting individual medical risk and risk across different human populations is often a significant challenge. For instance, cardiovascular diseases (CVDs) continue to be one of the leading causes of death around the world. Development of CVDs has been linked to nutrient exposure during fetal development, high fat and sedentary lifestyles, drug usage, adverse socioeconomic conditions, and various genes. Human environments are diverse, and public health research including the field of Human Biology can help identify risk factors and behaviors associated with chronic diseases. Large-scale clinical genetic studies with powerful bioinformatic approaches can also help elucidate some of these complex relationships.
Genomics and Epigenetics
A genome is all of the genetic material of an organism. In the case of humans, this includes 46 chromosomes and mtDNA. The human genome contains approximately three billion base pairs of DNA and has regions that are both noncoding and coding. Scientists now estimate that the human genome contains 20,000–25,000 protein-coding genes, with each chromosome containing a few hundred to a few thousand genes. As our knowledge of heredity increases, researchers have begun to realize the importance of epigenetics, or changes in gene expression that do not result in a change of the underlying DNA sequence. Epigenetics research is also crucial for unraveling gene regulation, which involves complex interactions between DNA, RNA, proteins, and the environment.
Genomics
The vast majority of the human genome is noncoding, meaning there are no instructions to make a protein or RNA product in these regions. Historically, noncoding DNA was referred to as “junk DNA” because these vast segments of the genome were thought to be irrelevant and nonfunctional. However, continual improvement of DNA sequencing technology along with worldwide scientific collaborations and consortia have contributed to our increased understanding of how the genome functions. Through these technological advances and collaborations, we have since discovered that many of these noncoding DNA regions are involved in dynamic genetic regulatory processes.
Genomics is a diverse field of molecular biology that focuses on genomic evolution, structure, and function; gene mapping; and genotyping (determining the alleles present). Evolutionary genomics determined that humans share about 98.8% percent of their DNA with chimpanzees. Given the phenotypic differences between humans and chimpanzees, having a DNA sequence difference of 1.2% seems surprising. However, a lot of genomics research is also focused on understanding how noncoding genomic regions influence how individual genes are turned “on” and “off” (i.e., regulated). Therefore, although DNA sequences are identical, regulatory differences in noncoding genetic regions (e.g., promoters) are believed to be largely responsible for the physical differences between humans and chimpanzees.
Further understanding of genomic regulatory elements can lead to new therapies and personalized treatments for a broad range of diseases. For example, targeting the regulatory region of a pathogenic gene to “turn off” its expression can prevent its otherwise harmful effects. Such molecular targeting approaches can be personalized based on an individual’s genetic makeup. Genome-wide association studies (GWAS), which seek to determine genes that are linked to complex traits and diseases, typically require significant computational efforts. This is because millions of DNA sequences must be analyzed and GWAS sometimes include thousands of participants. During the beginning of the genomics field, most of the large-scale genomics studies only included North American, European, and East Asian participants and patients. Researchers are now focusing on increasing ethnic diversity in genomic studies and databases. In turn, accuracy of individual disease risk across all human populations will be improved and more rare disease–causing alleles will be identified.
Epigenetics
All cells within your body have the same copy of DNA. For example, a brain neuron has the same DNA blueprint as does a skin cell on your arm. Although these cells have the same genetic information, they are considered specialized. The reason all cells within the body have the same DNA but different morphologies and functions is that different subsets of genes are turned “on” and “off” within the different cell types. A more precise explanation is that there is differential expression of genes among different cell types. In the case of neuronal cells, a unique subset of genes are active that allow them to grow axons to send and receive messages. This subset of genes will be inactive in non-neuronal cell types such as skin cells. Epigenetics is a branch of genetics that studies how these genes are regulated through mechanisms that do not change the underlying DNA sequence. “Special Topics: Epigenetics and X Chromosome Inactivation” details a well-known example of epigenetics regulation.
The prefix epi- means “on, above, or near,” and epigenetic mechanisms such as DNA methylation and histone modifications occur on, above, or near DNA. The addition of a methyl group (— CH₃) to DNA is known as DNA methylation (Figure 3.34). DNA methylation and other modifications made to the histones around which DNA are wrapped are thought to make chromatin more compact. This DNA is inaccessible to transcription factors and RNA polymerases, thus preventing genes from being turned on (i.e., transcribed). Other histone modifications have the opposite effect by loosening chromatin, which makes genes accessible to transcription factors.
It is important to note that environmental factors can alter DNA methylation and histone modifications and also that these changes can be passed from generation to generation. For example, someone’s epigenetic profile can be altered during a stressful time (e.g., natural disasters, famine, etc.), and those regulatory changes can be inherited by the next generation. Moreover, our epigenetic expression profile changes as we age. For example, certain places in our genome become “hyper” or “hypo” methylated over time. Identical twins also have epigenetic profiles that become more different as they age. Researchers are only beginning to understand what all of these genome-wide epigenetic changes mean. Scientists have also discovered that changes in epigenetic modifications can alter gene expression in ways that contribute to diseases. It is also important to note that, unlike DNA mutations (which permanently change the nucleotide sequence), epigenetic changes can be easily reversed. A lot of research now focuses on how drugs can alter or modulate changes in DNA methylation and histone modifications to treat diseases such as cancer.
Special Topic: Epigenetics and X Chromosome Inactivation
Mary Lyon was a British geneticist who presented a hypothesis for X chromosome inactivation (called the Lyon hypothesis) based on her work and other studies of the day. Females inherit two X chromosomes, one from each parent. Males have one functional X chromosome; however, this does not mean females have more active genes than males. During the genetic embryonic development of many female mammals, one of the X chromosomes is inactivated at random, so females have one functional X chromosome. The process of X chromosome inactivation in females occurs through epigenetic mechanisms, such as DNA methylation and histone modifications. Recent studies have analyzed the role of a long noncoding RNA called X-inactive specific transcript (XIST), which is largely responsible for the random silencing of one of the X chromosomes. The presence of two X chromosomes is the signal for XIST RNA to be expressed so that one X chromosome can be inactivated. However, some cells may have an active paternal X chromosome while other cells may have an active maternal X chromosome. This phenomenon is easily seen in calico and tortoiseshell cats (Figure 3.35). In cats, the gene that controls coat color is found on the X chromosome. During early embryo development, random inactivation of X chromosomes gives rise to populations of cells that express black or orange, which results in the unique coat patterning. Therefore, calico cats are typically always female.
Genetic Testing
To assist with public health efforts, newborn screening for genetic diseases has been available in the United States for over 50 years. One of the first available genetic tests was to confirm a phenylketonuria (PKU) diagnosis in infants, which is easily treatable with a dietary change. Currently, each state decides what genes are included on newborn screening panels and some states even have programs to help with infant medical follow-ups.
There are now hundreds of laboratories that provide testing for a few thousand different genes that can inform medical decisions for infants and adults. What has made this industry possible are the advancements in technology and decreased cost to patients. Moreover, genetic testing has been made available publicly to anyone without the assistance of medical professionals.
Clinical Testing
Clinical genetics tests assist patients with making medically informed decisions about family planning and health. Applications of this technology include assistance with in vitro fertilization (IVF) procedures, embryo genetic screening, and personalized medicine such as matching patients to cancer therapies. To ensure accuracy of patient genetic screening, it is important that all clinical laboratories are regulated. The Clinical Laboratory Improvement Amendments (CLIA) are United States federal standards that all human laboratory testing clinics must follow. A major benefit provided by some clinical genetic testing companies is access to genetic counselors, who have specialized education and training in medical genetics and counseling. For individuals with a family history of genetic disease, a physician may recommend genetic carrier screening to see if there is a risk for passing on a disease to a child. Genetic counselors provide expertise with interpretation of genetic testing results, as well as help guide and support patients when making impactful medical decisions.
Direct-to-Consumer (DTC) Genetic Testing
Genetic testing that is performed without the guidance of medical professionals is called direct-to-consumer (DTC) genetic testing. Companies that sell affordable genome sequencing products to the public continue to increase in popularity. These companies have marketing campaigns typically based on the notion of personal empowerment, which can be achieved by knowing more about your DNA. For example, if you are identified as having a slightly increased risk for developing celiac disease (Figure 3.36), then you may be motivated to modify your dietary consumption by removing gluten from your diet. Another scenario is that you could test positive for a known pathogenic BRCA1 or BRCA2 cancer-predisposing allele. In this case, you may want to follow up with a physician and obtain additional clinical testing, which could lead to life-altering decisions. DNA sequencing products for entertainment and lifestyle purposes are also available. For example, some DTC companies offer customized genetic reports for health and fitness, wherein recommendations for optimal exercise workout and meal plans are provided.
DTC testing typically lacks genetic counselor services to consumers, and regulations for nonclinical laboratories are not as strict. This has led to some controversies regarding company genetic products that provide health information. The company 23andMe was the first on the market to offer DTC health testing, and in 2013, the U.S. Food and Drug Administration (FDA) intervened. 23andMe worked toward complying with FDA regulations and then gained approval to offer testing on a few medically related genes. In 2017, 23andMe offered a “Late-Onset Alzheimer's Disease” genetic risk report. Such offerings have been criticized because customers could receive results they may not fully be able to interpret without professional assistance and advice. In turn, this could increase the stress of participants (sometimes called the “burden of knowing”) and could lead to unnecessary medical intervention.To address this issue, 23andMe now provides disclaimers and also interactive learning modules that customers must complete if they wish to view certain genotyping results. However, individuals who tested positive for a disease-causing allele have also been able to successfully seek medical help. The potential for harm and the proposed benefits of DTC testing continue to be a topic of debate and investigation.
Ancestry percentage tests are also widely popular (Figure 3.37). Customers are genotyped and their alleles are assigned to different groups from around the world (Chapter 4 will discuss human biological variation in further detail). However, the scientific significance and potential harm of ancestry percentage tests have been called into question. For example, most alleles tested are not exclusive to one population, and populations may be defined differently depending on the testing companies. If an allele is assigned to the “Irish” population, there is a good chance that the allele may have evolved in a different cultural group or region that pre-dates the formation of the country Ireland. In other words, genetic variation often pre-dates the origins of the population and geographical names of the region used by genetic testing companies. Another critique is that someone’s identity need not include biological relationships. In using the tests, customers have the option to find and connect online with other individuals with whom they share portions of their genome, which has resulted in both positive and negative personal experiences. Another interesting development in this field is that law enforcement is currently developing forensic techniques that involve mining DTC genomic databases for the purpose of identifying suspects linked to crimes. Regardless of these various considerations, there are now millions of individuals worldwide who have “unlocked the secrets” of their DNA, and the multibillion-dollar genomics market only continues to grow.
As you have seen in this chapter, DNA provides instructions to our cells, which results in the creation and regulation of proteins. Understanding these fundamental mechanisms is important to being able to understand how the evolutionary process works (see Chapter 4) and how humans vary from one another (see Chapters 13 and 14). In addition, advancement in genetic technologies—including ancient DNA studies, genomics, and epigenetics—has led to new anthropological understandings about our biological relationships to other living (extant) and extinct primates. Many of these genetic discoveries will be covered in the chapters to come.
Special Topic: Genetic Biotechnology
Polymerase Chain Reaction (PCR) and Sanger Sequencing
One of the most important inventions in the genetics field was polymerase chain reaction (PCR). In order for researchers to visualize and therefore analyze DNA, the concentration must meet certain thresholds. In 1985, Kary Mullis developed PCR, which can amplify millions of copies of DNA from a very small amount of template DNA (Figure 3.38). For example, a trace amount of DNA at a crime scene can be amplified and tested for a DNA match. Also, aDNA is typically degraded, so a few remaining molecules of DNA can be amplified to reconstruct ancient genomes. The PCR assay uses similar biochemical reactions to our own cells during DNA replication.
In Sanger sequencing, PCR sequences can be analyzed at the nucleotide level with the help of fluorescent labeling. Several different types of alleles and genetic changes can be detected in DNA by using this analysis. Figure 3.39 shows someone who is heterozygous for a single nucleotide allele. These methods continue to be used extensively alongside larger-scale genome technologies.
Genetic innovations are transforming the healthcare industry. However, the different types of technology and the results of these tests often include a learning curve for patients, the public, and medical practitioners. Microarray technology, by which DNA samples are genotyped (or “screened”) for specific alleles, has been available for quite some time (Figure 3.40). Presently, microarray chips can include hundreds of alleles that are known to be associated with various diseases. The microarray chip only binds with a DNA sample if it is “positive” for that particular allele and a fluorescent signal is emitted, which can be further analyzed.
Next-generation sequencing (NGS) is a newer technology that can screen the entire genome by analyzing millions of sequences within a single machine run. If a patient is suspected of having a rare genetic condition that cannot be easily diagnosed or the diagnosis is entirely unknown, whole genome sequencing may be recommended by a doctor. However, sequencing the entire genome is still not a cost-effective healthcare approach. Therefore, clinical NGS genetic testing typically only includes a smaller subset of the genome known to have pathogenic disease-causing mutations (i.e., the gene-coding, or “exonic,” regions of the genome). Sequencing cancer tumor genomes is another significant application of this technology. To better understand how genetic mutations affect gene expression patterns, tumor genomic analysis also often involves RNA sequencing (known as the “transcriptome”). The primary goal of this complex “multi-omics” analysis is to provide personalized medicine, where patient outcome can be improved by administering tailored targeted therapies.
Review Questions
- What is the purpose of DNA replication? Explain in a few sentences what happens during DNA replication. When do DNA mutations happen? And how does this create phenotypic variation (i.e., different phenotypes of the same physical trait)?
- Using your own words, what are homologous chromosomes and sister chromatids? What are the key differences between mitosis and meiosis?
- Determine if the pedigree diagram below (Figure 3.41) represents an autosomal dominant, autosomal recessive, or X-linked recessive pattern of inheritance. You should write the genotype (i.e., AA, Aa, or aa) above each square to help you (note: there may sometimes be two possible answers for a square’s genotype). Please also explain why you concluded a particular pattern of inheritance.
- Use base pairing rules to transcribe the following DNA template sequence into mRNA: GTAAAGGTGCTGGCCATC. Next, use the protein codon table (see Figure 3.21) to translate the sequence. In regard to transcription, explain what the significance is of the first and last codon/protein in the sequence.
- In your opinion, what do you think the benefits are of direct-to-consumer (DTC) genetic testing? What are the drawbacks and/or greater ethical concerns? Do you think benefits outweigh concerns?
- Imagine that you submit your DNA sample to a genetic testing company and among the various diseases for which they test, there is an allele that is associated with late-onset Alzheimer’s disease. You have the option to view your Alzheimer’s result or to not view your result. What do you do and why?
Key Terms
Adenosine triphosphate (ATP): A high-energy compound produced by mitochondria that powers cellular processes.
Allele: A nonidentical DNA sequence found in the same gene location on a homologous chromosome, or gene copy, that codes for the same trait but produces a different phenotype.
Amino acids: Organic molecules that are the building blocks of protein. Each of the 20 different amino acids have their own unique chemical property. Amino acids are chained together to form proteins.
Ancient DNA (aDNA): DNA that is extracted from organic remains and that often dates from hundreds to thousands of years ago. Also, aDNA is typically degraded (i.e., damaged) due to exposure to the elements such as heat, acidity, and humidity.
Aneuploid: A cell with an unexpected amount of chromosomes. The loss or gain of chromosomes can occur during mitotic or meiotic division.
Antibodies: Immune-related proteins that can detect and bind to foreign substances in the blood such as pathogens.
Apoptosis: A series of molecular steps that is activated leading to cell death. Apoptosis can be activated when a cell fails checkpoints during the cell cycle; however, cancer cells have the ability to avoid apoptosis.
Autosomal: Refers to a pattern of inheritance in which an allele is located on an autosome (non sex chromosome).
Base pairs: Chemical bonding between nucleotides. In DNA, adenine (A) pairs with thymine (T) and cytosine (C) pairs with guanine (G); in RNA, adenine (A) always pairs with uracil (U).
Carbohydrate: Molecules composed of carbon and hydrogen atoms that can be broken down to supply energy.
Carrier: An individual who has a heterozygous genotype that is typically associated with a disease.
Cell cycle: A cycle the cell undergoes with checkpoints between phases to ensure that DNA replication and cell division occur properly.
Cell surface antigen: A protein that is found on a red blood cell’s surface.
Centromere: A structural feature that is defined as the “center” of a chromosome and that creates two different arm lengths. This term also refers to the region of attachment for microtubules during mitosis and meiosis.
Chromatin: DNA wrapped around histone complexes. During cell division, chromatin becomes a condensed chromosome.
Chromosome: DNA molecule that is wrapped around protein complexes, including histones.
Codominance: The effects of both alleles in a genotype can be seen in the phenotype.
Codons: A sequence that comprises three DNA nucleotides that together code for a protein.
Complex diseases: A category of diseases that are polygenic and are also influenced by environment and lifestyle factors.
Cytoplasm: The “jelly-like” matrix inside of the cell that contains many organelles and other cellular molecules.
Deleterious: A mutation that increases an organism’s susceptibility to disease.
Deoxyribonucleic acid (DNA): A molecule that carries the hereditary information passed down from parents to offspring. DNA can be described as a “double helix”’ shape. It includes two chains of nucleotides held together by hydrogen bonds with a sugar phosphate backbone.
Diploid: Refers to an organism or cell with two sets of chromosomes.
DNA methylation: Methyl groups bind DNA, which modifies the transcriptional activity of a gene by turning it “on” or “off.”
DNA polymerase: Enzyme that adds nucleotides to existing nucleic acid strands during DNA replication. These enzymes can be distinguished by their processivity (e.g., DNA replication).
DNA replication: Cellular process in which DNA is copied and doubled.
DNA sequence: The order of nucleotide bases. A DNA sequence can be short, long, or representative of entire chromosomes or organismal genomes.
Dominant: Refers to an allele for which one copy is sufficient to be visible in the phenotype.
Elongation: The assembly of new DNA from template strands with the help of DNA polymerases.
Enzymes: Proteins responsible for catalyzing (accelerating) various biochemical reactions in cells.
Epigenetic profile: The methylation pattern throughout a genome—that is, which genes (and other genomic sites) are methylated and unmethylated.
Epigenetics: Changes in gene expression that do not result in a change of the underlying DNA sequence. These changes typically involve DNA methylation and histone modifications. These changes are reversible and can also be inherited by the next generation.
Euchromatin: Loosely coiled chromosomes found within the nucleus that are accessible for regulatory processing of DNA.
Eukaryote: Single-celled or multicelled organism characterized by a distinct nucleus, with each organelle surrounded by its own membrane.
Exon: Protein-coding segment of a gene.
Gametes: Haploid cells referred to as an egg and sperm that will fuse together during sexual reproduction to form a diploid organism.
Gene: Segment of DNA that contains protein-coding information and various regulatory (e.g., promoter) and noncoding (e.g., introns) regions.
Genetic recombination: A cellular process that occurs during meiosis I in which homologous chromosomes pair up and sister chromatids on different chromosomes physically swap genetic information.
Genome: All the genetic information of an organism.
Genotype: The combination of two alleles that code for or are associated with the same gene.
Genotyping: A molecular procedure that is performed to test for the presence of certain alleles or to discover new ones.
Germ cells: Specialized cells that form gametes (egg and sperm cells).
Haploid: Cell or organism with one set of chromosomes (n = 23).
Helicase: A protein that breaks the hydrogen bonds that hold double-stranded DNA together.
Heterozygous: Genotype that consists of two different alleles.
Histones: Proteins that DNA wraps around to assist with DNA organization within the nucleus.
Homologous chromosomes: A matching pair of chromosomes wherein one chromosome is maternally inherited and the other is paternally inherited.
Homozygous: Genotype that consists of two identical alleles.
Incomplete dominance: Heterozygous genotype that produces a phenotype that is a blend of both alleles.
Initiation: The recruitment of proteins to separate DNA strands and begin DNA replication.
Interphase: Preparatory period of the cell cycle when increased metabolic demand allows for DNA replication and doubling of the cell prior to cell division.
Introns: Segment of DNA that does not code for proteins.
Karyotyping: The microscopic procedure wherein the number of chromosomes in a cell is determined.
Lagging strand: DNA template strand that is opposite to the leading strand during DNA replication. This strand is created in several disconnected sections and other enzymes fill in the missing nucleotide gaps between these sections.
Leading strand: DNA template strand in which replication proceeds continuously.
Lipids: Fatty acid molecules that serve various purposes in the cell, including energy storage, cell signaling, and structure.
Meiosis: The process that gametes undergo to divide. The end of meiosis results in four haploid daughter cells.
Mendelian genetics: A classification given to phenotypic traits that are controlled by a single gene.
Messenger RNA (mRNA): RNA molecule that is transcribed from DNA. Its tri-nucleotide codons are “read” by a ribosome to build a protein.
Microarray technology: A genotyping procedure that utilizes a microarray chip, which is a collection of thousands of short nucleotide sequences attached to a solid surface that can probe genomic DNA.
Microbiome: The collective genomes of the community of microorganisms that humans have living inside of their bodies.
Mitochondrial DNA (mtDNA): Circular DNA segment found in mitochondria that is inherited maternally.
Mitochondrion: Specialized cellular organelle that is the site for energy production. It also has its own genome (mtDNA).
Mitosis: The process that somatic cells undergo to divide. The end of mitosis results in two diploid daughter cells.
Molecular anthropologists: Individuals who use molecular techniques (primarily genetics) to compare ancient and modern populations and to study living populations of humans and nonhuman primates.
Molecular geneticists: Biologists that study the structure and function of genes.
Mutation: A nucleotide sequence variation from the template DNA strand that can occur during replication. Mutations can also happen during recombination.
Next-generation sequencing: A genotyping technology that involves producing millions of nucleotide sequences (from a single DNA sample) that are then read with a sequencing machine. It can be used for analyzing entire genomes or specific regions and requires extensive program-based applications.
Nuclear envelope: A double-layered membrane that encircles the nucleus.
Nucleic acid: A complex structure (like DNA or RNA) that carries genetic information about a living organism.
Nucleotide: The basic structural component of nucleic acids, which includes DNA (A, T, C, and G) and RNA (A, U, C, and G).
Nucleus: Double-membrane cellular organelle that helps protect DNA and also regulates nuclear activities.
Organelle: A structure within a cell that performs specialized tasks that are essential for the cell. There are different types of organelles, each with its own function.
Pathogenic: A genetic mutation (i.e., allele) that has a harmful phenotypic disease-causing effect.
Pedigree: A diagram of family relationships that indicates which members may have or carry certain genetic and/or phenotypic traits.
Penetrance: The proportion of how often the possession of an allele results in an expected phenotype. Some alleles are more penetrant than others.
Phenotype: The physical appearance of a given trait.
Phospholipid bilayer: Two layers of lipids that form a barrier due to the properties of a hydrophilic (water-loving) head and a hydrophobic (water-repelling) tail.
Polygenic trait: A phenotype that is controlled by two or more genes.
Polymerase chain reaction (PCR): A molecular biology procedure that can make copies of genomic DNA segments. A small amount of DNA is used as a starting template and is then used to make millions of copies.
Prokaryote: A single-celled organism characterized by the lack of a nucleus and membrane-enclosed organelles.
Promoter: The region of a gene that initiates transcription. Transcription factors can bind and DNA methylation may occur at a promoter site, which can modify the transcriptional activities of a gene.
Protein: Chain of amino acids that folds into a three-dimensional structure that allows a cell to function in a variety of ways.
Protein synthesis: A multi-step process by which amino acids are strung together by RNA machinery read from a DNA template.
Recessive: Refers to an allele whose effect is not normally seen unless two copies are present in an individual’s genotype.
Ribonucleic acid (RNA): Single-stranded nucleic acid molecule.There are different RNAs found within cells and they perform a variety of functions, such as cell signaling and involvement in protein synthesis.
Ribosomal RNA (rRNA): A ribosome-bound molecule that is used to correctly assemble amino acids into proteins.
Ribosome: An organelle in the cell found in the cytoplasm or endoplasmic reticulum. It is responsible for reading mRNA and protein assemblage.
RNA polymerase: An enzyme that catalyzes the process of making RNA from a DNA template.
Sanger-sequencing: A process that involves the usage of fluorescently labeled nucleotides to visualize DNA (PCR fragments) at the nucleotide level.
Semi-conservative replication: DNA replication in which new DNA is replicated from an existing DNA template strand.
Sequencing: A molecular laboratory procedure that produces the order of nucleotide bases (i.e., sequences).
Sister chromatids: During DNA replication, sister chromatids are produced on the chromosome. In cell division, sister chromatids are pulled apart so that two cells can be formed. In meiosis, sister chromatids are also the sites of genetic recombination.
Somatic cells: Diploid cells that comprise body tissues and undergo mitosis for maintenance and repair of tissues.
Splicing: The process by which mature mRNAs are produced. Introns are removed (spliced) and exons are joined together.
Sugar phosphate backbone: A biochemical structural component of DNA. The “backbone” consists of deoxyribose sugars and phosphate molecules.
Telomere: A compound structure located at the ends of chromosomes to help protect the chromosomes from degradation after every round of cell division.
Termination: The halt of DNA replication activity that occurs when a DNA sequence “stop” codon is encountered.
Tissue: A cluster of cells that are morphologically similar and perform the same task.
Transcription: The process by which DNA nucleotides (within a gene) are copied, which results in a messenger RNA molecule.
Transcription factors: Proteins that bind to regulatory regions of genes (e.g., promoter) and increase or decrease the amount of transcriptional activity of a gene, including turning them “on” or “off.”
Transfer RNA (tRNA): RNA molecule involved in translation. Transfer RNA transports amino acids from the cell’s cytoplasm to a ribosome.
Translation: The process by which messenger RNA codons are read and amino acids are “chained together” to form proteins.
X-linked: Refers to a pattern of inheritance where the allele is located on the X or Y chromosome.
About the Author
Hayley Mann, M.A.
Binghamton University, hmann3@binghamton.edu
Hayley Mann received her bachelor’s degree in Genetics from the University of California, Davis, and continued her graduate studies in Biological and Molecular Anthropology at the California State University, Sacramento. She is currently a Ph.D. candidate at Binghamton University, where her dissertation focus is on studying genetic variation of Pacific Islanders (Republic of Vanuatu) and also changes in health as the result of colonization. Hayley also works in clinical molecular carrier screening and specializes in DNA-sequencing procedures.
For Further Exploration
National Human Genome Research Institute
NOVA. 2018. Gene Sequencing Speeds Diagnosis of Deadly Newborn Diseases. NOVA, March 7, 2018. Accessed January 31, 2023. https://www.pbs.org/wgbh/nova/next/body/newborn-gene-sequencing/.
Zimmer, Carl. N.d. “Carl Zimmer’s Game of Genomes.” STATnews. Accessed January 31, 2023. https://www.statnews.com/feature/game-of-genomes/season-one/.
Illumina. 2016. “Illumina Sequencing by Synthesis.” YouTube.com, October 5, 2016. Accessed January 31, 2023. https://www.youtube.com/watch?v=fCd6B5HRaZ8.
References
Aartsma-Rus, Annemieke, Ieke B. Ginjaar, and Kate Bushby. 2016. “The Importance of Genetic Diagnosis for Duchenne Muscular Dystrophy.” Journal of Medical Genetics 53 (3): 145–151.
Acuna-Hidalgo, Rocio, Joris A. Veltman, and Alexander Hoischen. 2016. “New Insights into the Generation and Role of De Novo Mutations in Health and Disease.” Genome Biology 17 (241): 1–19.
Albert, Benjamin, Susanna Tomassetti, Yvonne Gloor, Daniel Dilg, Stefano Mattarocci, Slawomir Kubik, Lukas Hafner, and David Shore. 2019. "Sfp1 Regulates Transcriptional Networks Driving Cell Growth and Division through Multiple Promoter-Binding Modes." Genes & Development 33 (5–6): 288–293.
Almathen, Faisal, Haitham Elbir, Hussain Bahbahani, Joram Mwacharo, and Olivier Hanotte. 2018. “Polymorphisms in Mc1r and Asip Genes Are Associated with Coat Color Variation in the Arabian Camel.” Journal of Heredity 109 (6): 700–706.
Ballester, Leomar Y., Rajyalakshmi Luthra, Rashmi Kanagal-Shamanna, and Rajesh R. Singh. 2016. “Advances in Clinical Next-Generation Sequencing: Target Enrichment and Sequencing Technologies.” Expert Review of Molecular Diagnostics 16 (3): 357–372.
Baranovskiy, Andrey G., Vincent N. Duong, Nigar D. Babayeva, Yinbo Zhang, Youri I. Pavlov, Karen S. Anderson, and Tahir H. Tahirov. 2018. “Activity and Fidelity of Human DNA Polymerase Alpha Depend on Primer Structure.” Journal of Biological Chemistry 293 (18): 6824–6843.
Brezina, Paulina R., Raymond Anchan, and William G. Kearns. 2016. “Preimplantation Genetic Testing for Aneuploidy: What Technology Should You Use and What Are the Differences?” Journal of Assisted Reproduction and Genetics 33 (7): 823–832.
Bultman, Scott J. 2017. “Interplay Between Diet, Gut Microbiota, Epigenetic Events, and Colorectal Cancer." Molecular Nutrition & Food Research 61 (1):1–12.
Cutting, Garry R. 2015. “Cystic Fibrosis Genetics: From Molecular Understanding to Clinical Application.” Nature Reviews Genetics 16 (1): 45–56.
D'Alessandro, Giuseppina., and Fabrizio d'Adda di Fagagna. 2017. “Transcription and DNA Damage: Holding Hands or Crossing Swords?” Journal of Molecular Biology 429 (21): 3215–3229.
De Craene, Johan-Owen, Dimitri L. Bertazzi, Séverine Bar, and Sylvie Friant. 2017. “Phosphoinositides, Major Actors in Membrane Trafficking and Lipid Signaling Pathways.” International Journal of Molecular Sciences 18 (3): 1–20.
Deng, Lian, and Shuhua Xu. 2018. “Adaptation of Human Skin Color in Various Populations.” Hereditas 155 (1): 1–12.
Dever, Thomas E., Terri G. Kinzy, and Graham D. Pavitt. 2016. “Mechanism and Regulation of Protein Synthesis in Saccharomyces Cerevisiae.” Genetics 203 (1): 65–107.
Eme, Laura, Anja Spang, Jonathan Lombard, Courtney W. Stairs, and Thijs J. G. Ettema. 2017. “Archaea and the Origin of Eukaryotes.” Nature Reviews Microbiology 15 (12): 711–723.
Gomez-Carballa, Alberto, Jacobo Pardo-Seco, Stefania Brandini, Alessandro Achilli, Ugo A. Perego, Michael D. Coble, Toni M. Diegoli, et al. 2018. “The Peopling of South America and the Trans-Andean Gene Flow of the First Settlers.” Genome Research 28 (6): 767–779.
Gvozdenov, Zlata, Janhavi Kolhe, and Brian C. Freeman. 2019. “The Nuclear and DNA-Associated Molecular Chaperone Network.” Cold Spring Harbor Perspectives in Biology 11 (10): a034009.
Harkins, Kelly M., and Anne C. Stone. 2015. “Ancient Pathogen Genomics: Insights into Timing and Adaptation.” Journal of Human Evolution 79: 137–149.
Jackson, Maria, Leah Marks, Gerhard H. W. May, and Joanna B. Wilson. 2018. “The Genetic Basis of Disease.” Essays in Biochemistry 62 (5): 643–723.
Lenormand, Thomas, Jan Engelstadter, Susan E. Johnston, Erik Wijnker, and Christopher R. Haag. 2016. “Evolutionary Mysteries in Meiosis.” Philosophical Transactions of the Royal Society B 371: 1–14.
Levy, Shawn E., and Richard M. Myers. 2016. “Advancements in Next-Generation Sequencing.” Annual Review of Genomics and Human Genetics 17: 95–115.
Lindo, John, Emilia Huerta-Sánchez, Shigeki Nakagome, Morten Rasmussen, Barbara Petzelt, Joycelynn Mitchell, Jerome S. Cybulski, et al. 2016. "A Time Transect of Exomes from a Native American Population Before and After European Contact." Nature Communications 7: 1–11. https://doi.org/10.1038/ncomms13175.
Lu, Mengfei, Cathryn M. Lewis, and Matthew Traylor. 2017. “Pharmacogenetic Testing through the Direct-to-Consumer Genetic Testing Company 23andme.” BMC Medical Genomics 10 (47): 1–8.
Ly, Lundi, Donovan Chan, Mahmoud Aarabi, Mylene Landry, Nathalie A. Behan, Amanda J. MacFarlane, and Jacquetta Trasler. 2017. “Intergenerational Impact of Paternal Lifetime Exposures to Both Folic Acid Deficiency and Supplementation on Reproductive Outcomes and Imprinted Gene Methylation.” Molecular Human Reproduction 23 (7): 461–477.
Ma, Wenxiu, Giancarlo Bonora, Joel B. Berletch, Xinxian Deng, William S. Noble, and Christine M. Disteche. 2018. “X-Chromosome Inactivation and Escape from X Inactivation in Mouse.” Methods in Molecular Biology 1861: 205–219.
Machiela, Mitchell J., Weiyin Zhou, Eric Karlins, Joshua N. Sampson, Neal D. Freedman, Qi Yang, Belynda Hicks, et al. 2016. “Female Chromosome X Mosaicism Is Age-Related and Preferentially Affects the Inactivated X Chromosome.” Nature Communications 7: 1–9. https://doi.org/10.1038/ncomms11843.
Mahdavi, Morteza, Mohammadreza Nassiri, Mohammad M. Kooshyar, Masoume Vakili-Azghandi, Amir Avan, Ryan Sandry, Suja Pillai, Alfred K. Lam, and Vinod Gopalan. 2019. “Hereditary Breast Cancer; Genetic Penetrance and Current Status with BRCA.” Journal of Cellular Physiology 234 (5): 5741–5750.
McDade, Thomas W., Calen P. Ryan, Meaghan J. Jones, Morgan K. Hoke, Judith Borja, Gregory E. Miller, Christopher W. Kuzawa, and Michael S. Kobor. 2019. “Genome-Wide Analysis of DNA Methylation in Relation to Socioeconomic Status During Development and Early Adulthood.” American Journal of Physical Anthropology 169 (1): 3–11.
Migeon, Barbara R. 2017. “Choosing the Active X: The Human Version of X Inactivation.” Trends in Genetics 33 (12): 899–909.
Myerowitz, Rachel. 1997. “Tay-Sachs Disease-Causing Mutations and Neutral Polymorphisms in the Hex A Gene.” Human Mutation 9 (3): 195–208.
Onufriev, Alexey V., and Helmut Schiessel. 2019. “The Nucleosome: From Structure to Function through Physics.” Current Opinion in Structural Biology 56: 119–130.
Quillen, Ellen E., Heather L. Norton, Esteban J. Parra, Frida Lona-Durazo, Khai C. Ang, Florin M. Illiescu, Laurel N. Pearson, et al. 2019. “Shades of Complexity: New Perspectives on the Evolution and Genetic Architecture of Human Skin.” American Journal of Physical Anthropology 168 (67): 4–26.
Raspelli, Erica, and Roberta Fraschini. 2019. “Spindle Pole Power in Health and Disease.” Current Genetics 65 (4): 851–855.
Ravinet, M., R. Faria, R. K. Butlin, J. Galindo, N. Bierne, M. Rafajlovic, M. A. F. Noor, B. Mehlig, and A. M. Westram. 2017. “Interpreting the Genomic Landscape of Speciation: A Road Map for Finding Barriers to Gene Flow.” Journal of Evolutionary Biology 30 (8): 1450–1477.
Regev, Aviv, Sarah A. Teichmann, Eric S. Lander, Ido Amit, Christophe Benoist, Ewan Birney, Bernd Bodenmiller, et al. 2017. “The Human Cell Atlas.” Elife 6e27041: 1–30. https://doi.org/10.7554.eLife.27041.
Roberts, Andrea L., Nicole Gladish, Evan Gatev, Meaghan J. Jones, Ying Chen, Julia L. MacIsaac, Shelley S. Tworoger, et al. 2018. “Exposure to Childhood Abuse Is Associated with Human Sperm DNA Methylation.” Translational Psychiatry 8 (194): 1–11.
Roger, Andrew J., Sergio A. Muñoz-Gómez, and Ryoma Kamikawa. 2017. “The Origin and Diversification of Mitochondria.” Current Biology 27 (21): R1177–R1192. https://www.sciencedirect.com/science/article/pii/S096098221731179X?via%3Dihub#!
Ségurel, Laure, and Céline Bon. 2017. “On the Evolution of Lactase Persistence in Humans.” Annual Review of Genomics and Human Genetics 18: 297–319.
Sheth, Bhavisha P., and Vrinda S. Thaker. 2017. “DNA Barcoding and Traditional Taxonomy: An Integrated Approach for Biodiversity Conservation.” Genome 60 (7): 618–628.
Skloot, Rebecca. 2010. The Immortal Life of Henrietta Lacks. New York: Crown Publishing Group.
Snedeker, Jonathan, Matthew Wooten, and Xin Chen. 2017. “The Inherent Asymmetry of DNA Replication.” Annual Review of Cell and Developmental Biology 33: 291–318.
Sullivan-Pyke, Chantae, and Anuja Dokras. 2018. “Preimplantation Genetic Screening and Preimplantation Genetic Diagnosis.” Obstetrics and Gynecology Clinics of North America 45 (1): 113–125.
Szostak, Jack W. 2017. “The Narrow Road to the Deep Past: In Search of the Chemistry of the Origin of Life.” Angewandte Chemie International Edition 56 (37): 11037–11043.
Tessema, Mathewos, Ulrich Lehmann, and Hans Kreipe. 2004. “Cell Cycle and No End.” Virchows Archiv European Journal of Pathology 444 (4): 313–323.
Tishkoff, Sarah A., Floyd A. Reed, Alessia Ranciaro, Benjamin F. Voight, Courtney C. Babbitt, Jesse S. Silverman, Kweli Powell, et al. 2007. “Convergent Adaptation of Human Lactase Persistence in Africa and Europe.” Nature Genetics 39 (1): 31–40.
Visootsak, Jeannie, and John M. Graham, Jr. 2006. “Klinefelter Syndrome and Other Sex Chromosomal Aneuploidies.” Orphanet Journal of Rare Diseases 1:42. https://doi.org/10.1186/1750-1172-1-42.
Wolfe, George C., dir. 2017. The Immortal Life of Henrietta Lacks. HBO Films, April 22, 2017. TV Movie.
Yamamoto, Fumi-ichiro, Henrik Clausen, Thayer White, John Marken, and Sen-itiroh Hakomori. 1990. “Molecular Genetic Basis of the Histo-Blood Group ABO System.” Nature 345 (6272): 229–233.
Zlotogora, Joël. 2003. “Penetrance and Expressivity in the Molecular Age.” Genetics in Medicine 5 (5): 347–352.
Zorina-Lichtenwalter, Katerina, Ryan N. Lichtenwalter, Dima V. Zaykin, Marc Parisien, Simon Gravel, Andrey Bortsov, and Luda Diatchenko. 2019. “A Study in Scarlet: MC1R as the Main Predictor of Red Hair and Exemplar of the Flip-Flop Effect.” Human Molecular Genetics 28 (12): 2093-2106.
Zwart, Haeh. 2018. “In the Beginning Was the Genome: Genomics and the Bi-Textuality of Human Existence.” New Bioethics 24 (1): 26–43.
Joylin Namie, Ph.D., Truckee Meadows Community College
Learning Objectives
- Identify and describe the major developments in scientific thought that led to the discovery of evolutionary processes.
- Explain how natural selection works and results in evolutionary change over time.
- Explain what is meant by the “Modern Synthesis” and its impacts on evolutionary thought.
- Discuss the teaching of human evolution in the U.S. and abroad.
The Beginnings of Evolutionary Thinking
Throughout our evolutionary history, humans have developed an understanding of the natural world as they interacted with and extracted resources from it. To survive, our earliest ancestors possessed an understanding of the physical environment, including weather patterns, animal behavior, edible and medicinal plants, locations of water, and seasonal cycles. Many ancient cultures, including those of the Americas (Dunbar-Ortiz 2014), Mesopotamia, and Egypt, left writings, hieroglyphics, and stories passed down through oral tradition detailing their understanding of the natural environment, human and zoological anatomy, botany, and medical practices (Moore 1993).
There are also over 2,000 years of organized thought and writing regarding evolution, including contributions from Greek, Roman, and Islamic scholars. Three examples of note are included here. The Greek philosopher Aristotle (384–322 BCE) studied the natural world, publishing several volumes on animals based on systematic observations, rather than attributing what he observed to divine intervention, as his contemporaries were doing (Figure 2.1). Aristotle’s system for the biological classification of nearly 500 species of animals was based on his own observations and dissections, interviews with specialists such as beekeepers and fishermen, and accounts of travelers. His nine book History of Animals, published in the 4th century BC (n.d.), was one of the first zoological taxonomies ever created. Aristotle’s primary contribution to the classification of biological species was to recognize that natural groups are based on structure, physiology, mode of reproduction, and behavior (Moore 1993, 39).
Aristotle’s History of Animals also placed animals in a hierarchy, ranking animals above plants due to what he claimed were their abilities to sense the world around them and to move. He also graded animals according to their modes of reproduction. Those giving birth to live young were placed above those who laid eggs. Warm-blooded animals ranked above invertebrates. This concept of “higher” and “lower” organisms was expanded upon by scholars in the Medieval period to form the Scala Naturae (Latin for “ladder of being”). This “Great Chain of Being,” depicting a hierarchy of beings with God at the top and minerals at the bottom (Figure 2.2), was thought by medieval Christians to have been decreed by God; in this Great Chain, humans were placed closer to God than other species. Aristotle’s works were rediscovered by Islamic scholars in the ninth century and translated into Arabic, Syriac, Persian, and later into Latin, becoming part of university curriculum in 13th-century Europe (Lindberg 1992), allowing Aristotle’s works and ideas to influence other thinkers for 2,000 years.
Science also owes a debt to many Arabic scholars. One such Islamic scholar and writer, who built upon the ideas of Aristotle, was Abū ʿUthman ʿAmr ibn Baḥr al-Kinānī al-Baṣrī / al-Jāḥiẓ, known as Al-Jahiz (776–868 CE), who authored over 200 books (El-Zaher 2018; Figure 2.3). His most well-known work was the seven-volume Kitab al-Hayawan or Book of Animals, in which he described over 350 species in zoological detail. Importantly, Al-Jahiz introduced the idea and mechanisms of biological evolution 1,000 years before Darwin proposed the concept of natural selection in 1859 (Love 2020). For instance, Al-Jahiz wrote about the struggle for existence, the transformation of species over time, and environmental factors that influence the process, all ideas that were later espoused by western European scientists in the 19th century. Al-Jahiz wrote:
Animals engage in a struggle for existing, and for resources, to avoid being eaten, and to breed. Environmental factors influence organisms to develop new characteristics to ensure survival, thus transforming them into new species. Animals that survive to breed can pass on their successful characteristics to their offspring. [Masood 2009]
Another important early Islamic scientist is Ibn al-Haytham (965–1040), a 10th-century Islamic scholar who contributed a great deal to our understanding of optics and how human vision works (Figure 2.4; Lindberg 1992, 177–180). Born in what is now Iraq, al-Haytham was a scholar of many disciplines, including mathematics, physics, mechanics, astronomy, philosophy, and medicine. He authored some 200 books in his lifetime and was an expert on Aristotle’s natural philosophy, logic, and metaphysics. In relation to evolution, al-Haytham’s methodology of investigation—specifically, using experiments to verify theory—is similar to what later became known as the modern scientific method. He is most famous for discovering the laws of reflection and refraction over 1,000 years ago and inventing the camera obscura, which was incredibly important for the eventual development of photography. His work is credited for its influence on astronomy, mathematics, and optics, inspiring Galileo, Johannes Kepler, and Sir Isaac Newton (Tasci 2020). As an inspirational scientific figure, al-Haytham was celebrated in 2016 by UNESCO as a trailblazer and the founder of modern optics (Figure 2.5). An International Year of Light was named in his honor and a scholarly conference on his contributions was held to coincide with the 1,000th anniversary of the publication of his Kitāb al-Manāẓir (Book of Optics; UNESCO.org 2015).
The writings of these Islamic scholars as well as similar scientific texts from other cultures were unknown to or unacknowledged by Western scientists until recently. Fortunately, many science teachers are now incorporating this content into their classes (Love 2020).
Western European Evolutionary Thought
Although there have been many different scientific traditions throughout world history, a new global discourse around science emerged in Western Europe in the 19th century. Europeans pointed to the continuing expansion of their colonial power, as well as their military and technological success, as evidence of the efficacy of Western science, which came to dominate on a global scale (Elshakry 2010). The movement toward a global science centered in Western Europe began with formulation of the Scientific Method.
The Scientific Method was first codified by Francis Bacon (1561–1626), an English politician who was likely influenced by the methods of inquiry established by Ibn al-Haytham centuries prior (Tbakhi and Amr 2007). Bacon has been called the founder of empiricism for proposing a system for weighing the truthfulness of knowledge based solely on inductive reasoning and careful observations of natural phenomena. Ironically, he died as a result of trying to scientifically observe the effects of cold on the putrefaction of meat. On a journey out of London, he purchased a chicken and stuffed it with snow for observation, catching a chill in the process. One week later, he died of bronchitis (Urbach, Quinton, and Lea 2023).
The second important development with regard to evolution was the concept of a species. John Ray (1627–1705), an English parson and naturalist, was the first person to publish a biological definition of species in his Historia Plantarum (History of Plants), a three volume work published in 1686, 1688, and 1704. Ray defined a species as a group of morphologically similar organisms arising from a common ancestor. However, we now define a species as a group of similar organisms capable of producing fertile offspring. In keeping with the scientific method, Ray classified plants according to similarities and differences that emerged from observation. He claimed that any seed from the same plant was the same species, even if it had slightly different traits.
The modern period of biological classification began with the work of Carl von Linne (“Carolus Linnaeus”) (1707–1778), a Swedish scientist who laid the foundations for the modern scheme of taxonomy used today. He established the system of binomial nomenclature, in which a species of animal or plant receives a name consisting of two terms: the first term identifies the genus to which it belongs and the second term identifies the species. His original Systema Naturae, published in 1736, went through several editions. By the tenth edition in 1758, mammals incorporated primates, including apes and humans, and the term Homo sapiens was introduced to signify the latter (Paterlini 2007).
Georges-Louis Leclerc, Comte de Buffon (1707–1788), was a prominent French naturalist whose work influenced prominent scientists in the second half of the 18th century. Buffon's idea that species change over time became a cornerstone of modern evolutionary theory. His technique of comparing similar structures across different species, called comparative anatomy, is still in use today in the study of evolution. He published 36 volumes of Histoire Naturelle during his lifetime and heavily influenced two prominent French thinkers who were to have significant impacts on our understanding of evolution, Georges Cuvier and Jean-Baptiste Lamarck.
Georges Cuvier (1769–1832) was a paleontologist and comparative anatomist (Figure 2.6). One of his first major contributions to the field of evolution was proof that some species had become extinct through detailed and comprehensive analyses of large fossil quadrupeds (Moore 1993, 111). The idea of extinction was not new, but it was challenging to demonstrate if a fossil species was truly extinct or still had living relatives elsewhere. It was also challenging in that it ran counter to religious beliefs of the time. The Bible’s Book of Genesis was interpreted as saying that all species had been created by God in the seven days it took to create the world and that all created species have survived to this day. Extinction was interpreted as implying imperfection, suggesting God’s work was flawed. Also, given that the Earth was calculated to have been created in 4004 B.C.E., based on biblical genealogies, there would not have been enough time for species to disappear (Moore 1993, 112).
Cuvier was so knowledgeable in this field that he became famous for his ability to reconstruct what an extinct animal looked like from fragmentary remains. He demonstrated that fossil mammoths differed from similar living creatures, such as elephants. His many examples of fossils telling the stories of animals that lived and then disappeared were taken as incontrovertible proof of extinctions (PBS 2001). Where Cuvier went awry was his hypothesis of how extinction worked and its causes. As part of his study of comparative anatomy, Cuvier made observations of stratified layers of rock, or sediment, each containing different species. From this, he drew conclusions that species were “fixed” and did not evolve, but then went extinct, and that different assemblages of fossils occurred at different times in the past, as evidenced by the sedimentary layers (Moore 1993, 118). Cuvier explained this through a theory of catastrophism, which stated that successive catastrophic deluges (akin to Biblical floods) swept over parts of the Earth periodically, exterminating all life. When the waters receded from a particular region, lifeforms from unaffected regions would repopulate the areas that were destroyed, giving rise to a new layer of species that looked different from the layer below it. This theory implied that species were fixed in place and did not evolve and that the Earth was young. In fact, Cuvier postulated that the last catastrophe was a deluge he believed occurred five to six thousand years ago, paving the way for the advent of humans (Moore 1993, 118). Cuvier’s catastrophism became part of an ongoing and vociferous debate between two schools of geology. The catastrophists believed the present state of the earth was the consequence of a series of violent catastrophes of short duration, while the uniformitarians thought it was the result of slow acting geological forces that continue to shape the earth.
James Hutton (1726–1797) was one prominent proponent of uniformitarianism. Based on evidence he found at sites in his native Scotland, Hutton argued that the Earth was much older than previously thought. Examining the geology of Siccar Point, a cliff site on the eastern coast of Scotland (Figure 2.7), Hutton concluded that the intersection of the vertical and horizontal rocks represented a gap in time of many millions of years, during which the lower rocks had been deformed and eroded before the upper layers were deposited on top. From this, Hutton argued sediments are deposited primarily in the oceans, where they become strata, or layers of sedimentary rock. Volcanic action uplifts these strata to form mountains, which are then subject to erosion from rain, rivers, and wind, returning sediment to the oceans (Moore 1993, 121). Hutton’s Theory of the Earth (1788) demanded vast periods of time (known as “deep time”) for such slow-working forces to shape the earth. At the time, he was heavily criticized for this view, as it contradicted the biblical version of the history of creation.
Another Scotsman, who was to become a highly influential geologist and a close friend of Darwin, was Charles Lyell (1797–1875). Lyell was originally a lawyer who began his studies of Geology at Oxford under the tutelage of catastrophist William Buckland, from whom he diverged when Buckland tried to find physical evidence of Noah’s flood from the Christian Bible. Lyell was instead intent on establishing geology as a science based on observation. Building upon Hutton’s ideas (published 50 years earlier), Lyell traveled throughout Europe, documenting evidence of uniformitarianism. During his travels, he cataloged evidence of sea level rise and fall and of volcanoes positioned atop much older rocks. He also found evidence of valleys formed through erosion, mountains resulting from earthquakes, and volcanic eruptions that had been witnessed or documented in the past (University of California Berkeley Museum of Paleontology n.d.). Lyell also espoused the principle that “rocks and strata (layers of rock) increase in age the further down they are in a geological sequence. Barring obvious upheavals or other evidence of disturbance, the same principle must apply to any fossils contained within the rock. The lower down in a sequence of rocks a fossil is, the older it is likely to be (Wood 2005, 12).”
Lyell published the first edition of his three-volume Principles of Geology in 1830–1833 (Figure 2.8). It established geology as a science, underwent constant revisions as new scientific evidence was discovered, and was published in 12 editions during Lyell’s lifetime. In it, he espoused the key concept of uniformitarianism—that "the present is the key to the past.” What this meant was that geological remains from the distant past can be explained by reference to geological processes now in operation and directly observable.
Jean-Baptiste Lamarck (1744–1829) was the first Western scientist to propose a mechanism explaining why and how traits changed in species over time, as well as to recognize the importance of the physical environment in acting on and shaping physical characteristics. Lamarck’s view of how and why species changed through time, known as the “Theory of Inheritance of Acquired Characteristics,” was first presented in the introductory lecture to students in his invertebrate zoology class at the Museum of Natural History in Paris in 1802 (Burkhardt 2013). It was based on the idea that as animals adapted to their environments through the use and disuse of characteristics, their adaptations were passed on to their offspring through reproduction (Figure 2.9). Lamarck was right about the environment having an influence on characteristics of species, as well as about variations being passed on through reproduction. He simply had the mechanism wrong.
Lamarck’s theory involved a three-step process. Step one involves an animal experiencing a radical change in its environment. Step two is the animal (either individual or species) responding with a new kind of behavior. Step three is how the behavioral change results in morphological (meaning physical) changes to the animal that are successfully passed on to subsequent generations (Ward 2018, 8). Lamarck’s most famous example was the proposition that giraffes actively stretched their necks to reach leaves on tall trees to eat. Over their lifetimes, the continuation of this habit resulted in gradual lengthening of the neck. These longer necks were then passed on to their offspring. Lamarck's theory was disproved when evolutionary biologist August Weismann published the results of an experiment involving mice (Figure 2.10). Weismann amputated the tails of 68 mice and then successively bred five generations of them, removing the tails of all offspring in each generation, eventually producing 901 mice, all of whom had perfectly healthy long tails in spite of having parents whose tails were missing (Weismann 1889).
How giraffes actually ended up with long necks is a different story. In an environment where the food supply is higher off the ground, and perhaps less available to competing species, giraffes who happened to have slightly longer necks (due to random individual variation and genetic mutation) would be more likely to survive. These giraffes would then be able to reproduce, passing along the slight variation in neck length that would allow their offspring to do the same. Over time, individuals with longer necks would be overrepresented in the population, and neck lengths overall would increase among giraffes. Unfortunately, Lamarck’s ideas challenged the scientific establishment of the time and were rejected. He was discredited and harassed “to the point of loss of money, reputation, and then health” (Ward 2018, 9).
The final piece in the evolutionary puzzle leading up to the theory of natural selection was put forth by Thomas Malthus (1766–1834), who published An Essay on Population in 1798. Malthus lived in England during the time of the Industrial Revolution. It was a time of great poverty and misery when many people migrated from the countryside to squalid, disease-ridden cities to work extremely long hours in dangerous conditions in factories, coal mines, and other industrial workplaces. Birth rates were high and starvation and disease were rampant. Malthus struggled to explain why. His answer was basically the idea of carrying capacity, an ecological concept still in use today. Malthus suggested the rate of population growth exceeded the rate of increase of the human food supply. In other words, people were outgrowing the available food crops. He also suggested that populations of animals and plants were naturally constrained by the food supply, resulting in reductions in population in times of scarcity, “restraining them within the prescribed bounds” (Moore 1993, 147). But, despite significant challenges, some individuals always survived. This was the key to later understandings of evolutionary change in species over time.
The Journey to Natural Selection
In Western European thought, the individual most closely associated with evolution is Charles Darwin (1809–1882). However, as one can see from the individuals and ideas presented in the prior section, he was not the first person to explore evolution and how it might work. In fact, Darwin built upon and synthesized many of the ideas—from geology to biology, ecology, and economy—discussed above. He was simply in the right place at the right time. If he had not worked out his ideas when he did, someone else would have. As a matter of fact, as noted below, someone else did, forcing Darwin to publicly reveal his theory.
Darwin’s journey to the discovery of natural selection began during a childhood spent being curious, experimenting, and collecting natural specimens. When Darwin was 12 years old, his nickname was “Gas” because of the foul-smelling chemistry experiments he and his older brother, Erasmus, performed late into the evenings in their makeshift laboratory in the garden of their parent’s home (Costa 2017). Darwin was also a lifelong collector of biological and geological specimens, most famously beetles, at times going to great lengths in pursuit of a new specimen, as one of his personal letters relates,
I will give a proof of my zeal: one day, on tearing off some old bark, I saw two rare beetles, and seized one in each hand; then I saw a third and new kind, which I could not bear to lose, so that I popped the one which I held in my right hand into my mouth. Alas! it ejected some intensely acrid fluid, which burnt my tongue so that I was forced to spit the beetle out, which was lost, as was the third one. [Darwin 2001 (1897), 50].
Darwin continued his observations and experiments during his formal education, culminating in his graduation from Cambridge in 1831, at which point he was invited to become a gentleman naturalist for a British Royal Navy surveying mission of the globe aboard the H.M.S. Beagle. It is worth noting that Darwin was only 22 years old and the captain’s third choice for the position (Costa 2017), but he proved extremely curious and methodical. The mission departed in December of 1831 and returned five years later (Figure 2.11). During this time, Darwin produced copious notebooks, observations, drawings, and reflections on the natural phenomena he encountered and the experiments he performed.
Discussing all of Darwin’s work aboard the Beagle is beyond the scope of this chapter, but his primary interests were in cataloging new varieties of plant and animal life and examining the geology of the places the ship made landfall. Part of Darwin’s success with regard to both ventures was due to his extreme seasickness, which began before the ship even left Plymouth Harbor. It never let up, encouraging Darwin to go ashore at every available opportunity. “In fact, of the nearly five years of the voyage, Darwin was actually on board the ship for just a year and a half altogether” (Costa 2017, 18).
During the voyage, the young Darwin tried to make sense of what he saw through the lens of the scientific paradigms he held when he left England, but he continually made observations that challenged these paradigms. For example, while the Beagle crewmen were charting the coast of Argentina, Darwin conducted fieldwork on land. There he observed species that were new to him, like armadillos. He also collected fossils, including those of extinct armadillos. Meaning, he had found both extant and extinct members of the same species in the same geographic location, which challenged the theory of catastrophism put forth by Cuvier, who argued that each variant of an animal, living or extinct, was its own distinct species (Moore 1993, 144). Darwin also observed geographic variation in the same species all along the east coast of South America, from Brazil to the southern tip of Argentina. He noted that some species were found in multiple localities and differed from place to place. Those living closer to each other exhibited only slight variations, while those living further apart might be cataloged as entirely different species if one did not know better.
He made similar observations in the Galapagos Islands located off the northwest coast of Ecuador, with regard to giant tortoises and finches (Figure 2.12). A local resident of the islands explained to Darwin that each island had its own variety of tortoise and that locals could discern which island a tortoise came from simply by looking at it. Darwin noted other such examples in both plants and animals, meaning geographic variation was occurring on separate, neighboring islands.
Prevailing views of time argued that variations in living species, and even the fossil armadillos and the living armadillos, were the result of separate creation events. According to this view, each variation, no matter how slight, was a different species. Challenging these ideas would mean challenging not only catastrophism, but the Fixity of Species and other well-accepted ideas of the time. Darwin was aware that he was a young, unestablished naturalist. He was also aware of the ruin that befell Lamarck when his theories were rejected. Lastly, Lyell, who was a good friend of Darwin’s, rejected evolution altogether. It is no wonder that Darwin published a great deal about the geological and fossil data he collected when he returned from the voyage, but not his early hypotheses about evolution.
Upon Darwin’s return to England, it took another twenty years of data collection and experimentation before he was ready to share his conclusions about evolution. Much of this work was conducted at Down House, his home of forty years, where he performed all sorts of experiments that laid the groundwork for his ideas about evolution. Darwin’s home was his laboratory, and he engaged the help of his children, neighbors, friends, and servants in collecting, dissecting, and experimenting. At one point in the 1850s, sheets of moistened paper covered with frogs eggs lined the hallways of the house, while flocks of sixteen different pigeon breeds cooed outside, glass jars filled with salt water and floating seeds filled the cellar, and the smell of dissected pigeon skeletons pervaded the air inside the house. There were also ongoing experiments in the yard, including piles of dissected flowers, beekeeping, and fenced-off plots of land where seedlings were under study. Darwin was a keen experimental scientist, observer, and a prolific writer and presenter of scientific papers. He regarded his work as “one long argument” that never really ended. In fact, Darwin published ten books after On the Origin of Species, addressing such far-ranging topics as animal behavior, orchids, and domestication, among others (Costa 2017).
Darwin may not have published Origins in 1859 had it not been for receiving a paper in June of 1858 from Alfred Russel Wallace, an English naturalist working in Malaysia, espousing the same ideas. Wallace had sent the paper to Darwin asking if it was worthy of publication and requesting he forward it to Lyell and the English botanist, Joseph Hooker. Darwin wrote to Lyell and Hooker about Wallace’s paper, entitled On the Tendency of Varieties to Depart Indefinitely from the Original Type. In recognition that both Wallace and Darwin had arrived at the same discovery, a “joint” paper composed of four parts (none of them actually coauthored) was read to the Linnaean Society by Lyell, then secretary of the Society, on July 1, 1858, and published on August 20. Darwin published On the Origin of Species 15 months later. (The original composite paper read before the Linnaean society is available to read for free from the Alfred Russell Wallace Website, on the 1858 Darwin-Wallace paper page.)
The Mechanism of Natural Selection
On the subject of natural selection and how it works, let’s hear from Darwin himself from the original publication of On the Origin of Species (1859):
A struggle for existence inevitably follows from the high rate at which all organic beings tend to increase. Every being, which during its natural lifetime produces several eggs or seeds, must suffer destruction during some period of its life, and during some season or occasional year, otherwise, on the principle of geometric increase, its numbers would quickly become so inordinately great that no country could support the product. Hence, as more individuals are produced than can possibly survive, there must in every case be a struggle for existence…It is the doctrine of Malthus applied with manifold force to the whole animal and vegetable kingdoms…There is no exception to the rule that every organic being naturally increases at so high a rate, that if not destroyed, the earth would soon be covered by the progeny of a single pair. …Owing to this struggle for life, any variation, however slight and from whatever cause proceeding, if it be in any degree profitable to an individual of any species…will tend to the preservation of that individual, and will be inherited by its offspring. The offspring, also, will thus have a better chance of surviving, for, of the many individuals of any species which are periodically born, but a small number can survive. I have called this principle, by which slight variation, if useful, is preserved, by the term Natural Selection. [Darwin 1859, 61–62]
Let us take a moment here to explore the mechanism of natural selection in more detail. Before we begin, it is important to recognize that Darwin defined evolution as descent with modification, by which he meant that species share a common ancestor yet change over time, giving rise to new species. Descent with modification refers to the fact that offspring from two parents look different from each of their parents, and from each other, meaning they descend with slight differences (“modifications”). If you have ever observed a litter of puppies or a field of flowers and stopped to examine each individual closely, you have seen that each differs from the next, and none look exactly like their parents. These variations are random, not specific, and may or may not be present in the following generations.
Darwin struggled to explain why some slight differences were preserved over time, while others were not. He turned to what he knew of animal breeding (artificial selection) for an explanation (Richards 1998). Darwin bred different breeds of pigeons at Down House, carefully documenting phenotypic differences across generations, including slight anatomical variations he observed through dissection. He also grew and crossbred species of flowers and dissected those too. Darwin was also very fond of hunting and of hunting dogs. In an early draft of his theory on speciation, he used greyhounds as an example of adaptation and selection, “noting how its every bone and muscle, instinct and habit, were fitted to run down hare (rabbits) (University of Cambridge n.d.).” In each case of plant and animal breeding Darwin observed, he noted that humans were selecting variants in each generation that had characteristics humans desired (i.e., sweetness of fruits, colors of flowers, fur type and color of animals). Breeders then continually bred plants and animals with the desired variants, over and over again. These small changes added up over time to create new species of plants and breeds of animals. Darwin also noted that artificial selection does not necessarily render plants or animals better adapted to their original environments. The characteristics humans desire often result in plants less likely to survive in the wild and animals more likely to suffer from certain behavioral or health problems. One has only to examine high rates of hip dysplasia in several modern breeds of dogs to observe what Darwin was referring to.
From his studies of artificial selection, Darwin drew the conclusion that nature also acts upon variations among members of the same species. Instead of human intervention, the forces of nature, such as heat, cold, predation, disease, and now climate change, determine which offspring, with which variants, survive and reproduce. These individuals then pass down these favorable variants to their own offspring. In this way, nature selects for traits that are beneficial within a particular environment and selects against traits that are disadvantageous within a particular environment. Over many generations, populations of a species become more and more adapted (or, in evolutionary terms, “fit”) for their specific environments. Darwin named this process natural selection.
This theory explained the variations in tortoises Darwin had observed years earlier in the Galapagos Islands (see Figure 2.12). Tortoises who lived on larger islands with lush vegetation to feed on were larger than those on smaller islands. They also had shorter necks and dome-shaped shells as their food was close to the ground. Tortoises on smaller, drier islands fed on cacti, which grew much taller. These tortoises had longer necks, longer front legs, and saddle-shaped shells, which allowed them to successfully stretch to reach the edible cactus pads that grew on the tops of the plants. How did these observable differences in the two tortoise populations emerge? Darwin would argue that, over time, small, random variations in the tortoises were differentially selected for by the distinct natural environments on different islands.
In addition to the biogeographical evidence Darwin offered from his research aboard the Beagle, as well as the evidence he documented from the artificial selection of plants and animals, he also relied, where possible, on fossil evidence. One example, mentioned above, were the fossil findings of extinct armadillos in Argentina in the same locations as living armadillos. Unfortunately, as Darwin himself noted, the geological record was incomplete, most often missing the transitional fossil forms that bridge extinct and living species. That issue has since been resolved with scientific research in geochronology and paleontology, among other fields. It is now well-established that life is far more ancient than was believed in Darwin’s time and that these ancient forms of life were the ancestors to all life on this planet (Kutschera and Niklas 2004).
What Darwin was Missing
Although the theory of evolution by natural selection gained traction in scientific circles in the decades after Darwin’s publication of Origins, he was never able to discover the mechanisms that caused variation within members of the same species or the means by which traits were inherited. This began later in 1892 with the publication of The Germ-Plasm: A Theory of Heredity by August Weismann, the same Weismann of the mouse tail experiment presented earlier in this chapter. In his book, Weissman proposed a theory of germ-plasm, which was a precursor to the later discovery and understanding of DNA. Weismann specialized in cytology, a branch of biology devoted to understanding the function of plant and animal cells. Germ-plasm, he proposed, was a substance in the germ cells (what we would call gametes, or sex cells, today) that carried hereditary information. He predicted that an offspring inherits half of its germ-plasm from each of its parents, and claimed that other cells (e.g. somatic, or body, cells) could not transmit genetic information from parents to offspring. This thereby erased the possibility that acquired traits (which he argued resided in somatic cells) could be inherited (Zou 2015). This contribution to evolutionary theory was an important step toward understanding genetic inheritance, but a connection between genetics and evolution was still lacking.
A series of lectures by a deceased Augustinian monk named Gregor Mendel (1822–1884), originally published in 1865, changed that perspective (Moore 1993, 285). Although Darwin was unknown to Mendel, he began a series of experiments with pea plants shortly after the publication of Darwin’s Origins, aiming to add to evolutionary understandings of heredity. As Mendel bred different generations of pea plants that had differences in seed shape and color, pod shape and color, flower position, and stem length, he documented consistent expression of some variations over others in subsequent generations. He meticulously documented the statistics of each crossing of plants and the percentages of phenotypes that resulted, eventually discovering the concept of dominance and recessiveness of characteristics. He also documented that there is no blending of inherited characteristics. For example, pea pod colors in the offspring of two parent plants, one with yellow pods and one with green, were either yellow or green, not yellowish green. Mendel also discovered that characteristics are inherited and expressed independently of each other, meaning the color of the pea pod was not necessarily expressed in conjunction with the pod being wrinkled or smooth. The recognition of the importance of Mendel’s work began with its rediscovery by Hugo de Vries and Carl Correns, both of whom were working on hypotheses regarding heredity in plants and had arrived at conclusions similar to Mendel’s. Both published papers supporting Mendel’s conclusions in 1900 (Moore 1993). Research into the inheritance of characteristics continued through the next three decades, and by the close of the 1930s, no major scientific questions remained regarding the transmission of heredity through genes, although what genes did and what chemicals they were made of were still under investigation. (chapter 3)
The Modern Synthesis refers to the merging of Mendelian genetics with Darwinian evolution that took place between 1930 and 1950. The basic principles of the synthetic theory were influenced by scientists working in many different fields, including genetics, zoology, biology, paleontology, botany, and statistics. Although there were differences of opinion among them, evolution came to be defined as changes in allele frequencies within populations. Genetic mutations, changes in the genetic code that are the original source of variation in every living thing, were believed to be random, the sources of phenotypic variation, and transmitted through sexual reproduction. These assertions were supported by a growing body of field and laboratory research, as well as new work in mathematics in the field of population genetics that defined evolution as numerical changes in gene frequencies within an interbreeding population from one generation to the next (Corning 2020). These changes in gene frequencies were argued to be the result of natural selection, mutation, migration (gene flow), and genetic drift, or random chance. Empirical research and mathematics demonstrated that very small selective forces acting over a relatively long time were able to generate substantial evolutionary change, including speciation (Plutynski 2009). Thus, the Modern Synthesis encompassed both microevolution, which refers to changes in gene frequencies between generations within a population, and macroevolution, longer-term changes in a population that can eventually result in speciation, wherein individuals from different populations are no longer able to successfully interbreed and produce viable offspring.
Genetics and the Origin of Species, published in 1937 by Theodosius Dobzhansky (Figure 2.13), was a cornerstone of the modern synthesis, applying genetics to the study of natural selection in wild populations, appealing to both geneticists and field biologists. Dobzhansky was interested in speciation, particularly in finding out what kept one species distinct from another and how speciation occurred. His research involved fruit flies, the species Drosophila pseudoobscura. At the time he began in the 1920s, biologists assumed all members of the same species had nearly identical genes. Dobzhansky traveled from Canada to Mexico capturing wild members of D.pseudoobscura, discovering that different populations had different combinations of alleles (forms of a gene) that distinguished them from other populations, even though they were all members of the same species. What, then, led to the creation of new species? Dobzhansky realized it was sexual selection. Members of the same species are more likely to live among their own kind and to recognize, and prefer, them as mates. Over time, as a result of random mutations, natural selection in a given environment, and genetic drift, meaning random changes in allele frequencies, members of the same population accumulate mutations distinct to their own gene pool, eventually becoming genetically distinct from other populations. What this means is that they have become a new species.
From these studies, Dobzhansky and others developed the Bateson-Dobzhansky-Muller model, also known as Dobzhansky-Muller model (Figure 2.14). It is a model of the evolution of genetic incompatibility. Combining genetics with natural selection, the model is important in understanding the role of reproductive isolation during speciation and the role of natural selection in bringing it about. Due to sexual selection (mate preference), populations can become reproductively isolated from one another. Eventually, novel mutations may arise and be selected for in one or both populations, rendering members of each genetically incompatible with the other, resulting in two distinct species.
Evolution and Natural Selection Observable Today (Transform into Special Topic/Dig Deeper)
Although this chapter primarily focuses on the past, it is important to remember that natural selection and evolution are still ongoing processes. Climate change, deforestation, urbanization, and other human impacts on the planet are influencing evolution among many contemporary species of plants and animals. One such example occurs among crested anoles (Anolis cristatellus), small lizards of the Caribbean jungle that are increasingly making their home in cities (Figure 2.15).
As urban sprawl continues across the planet, shrinking the availability of wilderness habitat, many wild species have come to make their homes in cities. “Urbanization has dramatically transformed landscapes around the world—changing how animals interact with nature, creating "heat islands" with higher temperatures, and hurting local biodiversity. Yet many organisms survive and even thrive in these urban environments, taking advantage of new habitats created by humans (National Science Foundation 2023). A recent example of lizards in Puerto Rico demonstrates evolution happening quickly in both behavior and genes that has come about as a result of the pressures of urban life (Winchell et al. 2023). Crested anoles, who once lived only in forests, now scurry around towns and cities throughout the Caribbean. As a result of having to sprint across large open spaces, like hot streets and parking lots, they have developed longer limbs. City-living lizards also now sport longer toe pads with special scales that allow them to cling to smooth surfaces, like windows and walls (as well as the plastic patio furniture pictured in Figure 2.15), rather than to the rough surfaces of bark and rock that their forest-living relatives climb. These adaptations enhance their ability to escape predators and survive in cities.
Researchers were curious to see if these changes were the result of genetic changes in urban populations, so they captured 96 male lizards in three Puerto Rican regions and compared their genomes to each other and to forest specimens in each location. They found that members of the three city-living populations were genetically distinct from each other, as well as from forest populations in their respective regions. In total, 33 genes in the urban lizards’ genomes were different from their forest-living counterparts and were linked to urbanization. These changes are estimated to have occurred just within the last 30 to 80 generations, suggesting that selective pressures related to survival in urban environments is strong. As study coauthor Kristen Winchell put it, “We are watching evolution as it is unfolding” (National Public Radio 2023). (If you are interested in hearing more about the study, see “How Lizards Adapt to Urban Living,” an episode of Science Sessions, a free podcast from the Proceedings of the National Academy of Sciences (PNAS 2023) featuring Dr. Winchell and her work.)
Misconceptions About Evolution Through Natural Selection
After many years of teaching about evolution and natural selection, it continues to surprise me how many misconceptions exist about how the process works. If you do a web image search for “human evolution,” the following image is likely to appear (Figure 2.16).
What is wrong with this picture? First, it implies that humans evolved from chimpanzees, which is incorrect. Although, as primates, we share a common ancestor very far back in time, we split from other primates, including our closest relatives, the nonhuman apes, several million years ago. This image also suggests that evolution is gradual and progressive; that it is intentional and directional; and that there is an end to it—a stopping point. As you will be learning, evolution takes place in fits and starts, depending on the physical environment, changes in climate, food supply, predation, reproductive success, and other factors. It is also not intentional, in the sense that there is no predetermined end; in fact, if environmental conditions change, species can evolve in different directions or even go extinct. Evolution also does not necessarily progress in the same direction over time. One example is the eel-like creature Qikiqtania wakei that lived 375 million years ago. It was originally a fish that evolved to walk on land, then evolved to live back in the water. Early tetrapods like Qikiqtania were likely spending more and more time out of the water during this period. The arrangement of bones and joints in their fins was starting to resemble arms and legs, which would have allowed them to prop themselves up in shallow water and survive on mudflats. Qikiqtania’s skeletal morphology, however, suggests that it then evolved from having rudimentary fingers and toes back to fins that allowed them to again swim in open water (Stewart et al. 2022).
There is also the misperception that natural selection can create entirely new anatomical structures out of thin air in response to changes in environmental pressures. For example, when asked if they can think of ways in which modern humans are continuing to evolve biologically, students often postulate that, as a result of climate change, humans might rapidly develop gills, webbed hands and feet, and learn to breathe underwater in response to rising sea levels. Unfortunately, natural selection can only act on slight variations in anatomy that are already present, and we have no rudimentary physiological system for breathing underwater. Given that natural selection can only act upon existing variation, humans have evolved in such a way that many parts of our bodies are prone to injury. Our knees are one example. The anterior cruciate ligament (ACL) in our knees is “vulnerable to tearing in humans because our upright bipedal posture forces it to endure much more strain than it is designed to” (Lents 2018, 23). When our ancestors made the transition from quadrupedalism to upright walking, we shifted from four bent legs to two straight legs, relying more on our bones than our muscles to support our weight. This is functional for normal walking and running in a straight line, but sudden shifts in direction and momentum, combined with the sizes and weights of modern humans, result in tears in an ACL that is simply not strong enough to bear the stress. If evolution had the capability to engineer a knee from scratch, it would look quite different, and any ligaments involved would likely be larger, stronger, and more flexible. For an interesting look at what anatomically modern humans might look like if we had evolved to withstand the stresses our bodies undergo in our present environment, see “This is what the perfect body looks like - according to science,” which was proposed by biological anthropologist Alice Roberts (Harrison 2018).
Another misperception about evolution is that some species are “more evolved” than others. Every species currently alive on the planet today is the result of millennia of natural selection that has rendered current members of that species well-adapted to their respective environments. Humans are no more “evolved” than fruit flies or yeast. What sets us apart are our cultural and technological abilities, which have allowed us to successfully survive in a wide variety of physical environments, many of which are now becoming too hot, too wet, or too dry to sustain human life without a great deal of technological intervention (IPCC 2022).
There is also some confusion about what “fitness” actually means and a failure to grasp that it changes as environmental conditions change. Evolutionary “fitness” is different from physical fitness. “Fitness” in evolutionary terms refers to an individual’s ability to survive and reproduce viable offspring who also survive and reproduce. Evolutionary fitness and reproductive success are highly dependent on specific environmental conditions, which can shift over time, greatly affecting the relative fitness of individuals in a population. Recent research on the impacts of climate change on dragonflies will serve to illustrate the point (Figure 2.17).
Pictured here is a male dragonfly, who, you will notice, has distinctive black markings on its wings. This is due to melanization. Males control breeding, and those with more ornamentation tend to attract more mates and to successfully ward off male competitors. Higher levels of melanization, however, have negative consequences for males in warming climates. The black markings absorb heat, elevating body temperatures, which can cause overheating, reduce male fighting ability, and even lead to death (Moore et al. 2021). Females are not as adversely affected because they spend more time in shaded areas, while males are more often flying in sunlit areas, fending off rivals. However, as highly melanized males become less viable, wing coloration is undergoing selection in males. In other words, what constitutes being “fit” for males has changed, favoring those who have fewer of the black markings and, therefore, are less negatively impacted by warming temperatures. Note that natural selection acts on individuals, “selecting” those who happen to be fit for particular environmental conditions at a particular point in time. Evolution, though, happens at the level of the population. If the climate continues to warm, populations of dragonflies who inhabit warming areas will increasingly exhibit less ornamentation in males.
Lastly, natural selection can only act on characteristics that influence reproductive success. Deleterious traits that have nothing to do with one’s ability to reproduce and successfully rear offspring to reproductive age will continue to be passed on. For example, the author of this chapter is a natural redhead, and redheads are predisposed genetically to a number of conditions that can negatively affect health (Colliss Harvey 2015), but some of these conditions are not diagnosed until later in life. One example is Parkinson’s disease (Chen et al. 2017), which is a degenerative neurological disorder. The average age of diagnosis of Parkinson’s is 60 years of age, meaning redheads may encounter such a diagnosis well past childbearing age, having already passed on the genetic predisposition. Thus, Parkinson’s disease cannot be selected out from the redhead family tree.
Dig Deeper: Teaching Evolution Around the World (Put in Chapter 17 or refer such a subject to chap 17)
Evolution is recognized as a central organizing principle for all scientific disciplines and accepted without controversy among scientists and educated people around the world. The United States has historically been the exception (Lerner 2000). In some parts of the U.S. the teaching of evolution to K-12 students continues to evoke controversy, related to politics and religion. The problem is compounded by the degree of control individual states and local school boards exercise over curriculum in the nation’s public schools (Lerner 2000).
The debate over teaching evolution in schools in the United States first came to a head in 1925 after several states attempted to legislate a ban against its teaching. The state of Tennessee eventually did pass such a ban, and the American Civil Liberties Union (ACLU) offered to defend any science teacher who agreed to break the law. John Scopes, who taught in a small, rural Tennessee school, continued to teach evolution, resulting in the “Scopes Monkey Trial,” one of the most famous media trials in American history (Figure 2.18). The entire nation listened to its broadcast live on the radio and read about it daily in hundreds of newspapers. Scopes was defended by Clarence Darrow, the most famous lawyer in the country at the time; Scopes was eventually convicted, though the conviction was later overturned. The pro-Evolution movement benefited greatly from Darrow’s questioning of those on the anti-Evolution side, whose responses were perceived negatively by well-educated listeners in northern cities. The teaching of evolution was also bolstered by a Supreme Court decision in 1947 overriding states’ rights to make decisions on church-state issues and by the launching of Sputnik, the first Soviet (Russian) satellite, in 1957, making science education a national priority. During the 1960s to 1970s, creationism and intelligent design began to take hold as courts ruled in favor of “academic freedom” (Pew Research Center 2019). Since that time, states, and even local school boards, have pushed for the removal of evolution from science curriculum and textbooks or for teaching evolution on equal footing with concepts such as creationism and intelligent design (Masci 2019).
Partly in response to chronically low scores by students in the United States on international measures of ability in science, math, and reading (Desilver 2017), development of the Next Generation Science Standards (NGSS) for K-12 education began in 2011 and were implemented in 2013 in public schools across the nation (nextgenscience.org). The standards were developed collaboratively by The National Research Council (NRC; a branch of the National Academy of Sciences), the National Science Teachers Association, the American Association for the Advancement of Science, and Achieve, an independent, nonpartisan, nonprofit education-reform organization dedicated to working with states to raise academic standards and graduation rates. Twenty-six states also partnered together in the development of the standards. However, due to a number of issues, including funding and support for teachers, adoption of the standards has varied by state. States were also allowed to alter the curriculum. One of the main issues in several states with the original curriculum was the teaching of evolution (Pew Research Center 2014). One prominent aspect to note is that the NGSS do not specify that human evolution be taught, and high school standards do not require teaching about all of the forces of evolution, only mutation and natural selection.
The situation regarding teaching evolution has changed greatly in recent years. A 2007 survey found that only one in three public high school biology teachers in the United States presented evolution consistently with the recommendations of the nation’s leading scientific authorities, and 13% of these teachers emphasized creationism as a valid scientific alternative to modern evolutionary biology (Plutzer, Branch, and Reid 2020). A repeat of the survey in 2019 demonstrated marked improvement in the amount of time teachers devoted to teaching evolution, as well as more teacher training and preparedness to teach evolution. Such improvements were attributed to the need to meet the Next Generation Science Standards, as well as continuing outreach by the National Science Teaching Association, the National Association of Biology Teachers, and the National Academy of Sciences in producing classroom resources and providing professional development opportunities to advance the inclusion of evolution in the nation’s classrooms.
Public acceptance of evolution has also substantially improved in recent years (Miller et al. 2022). National samples of American adults have been asked at regular intervals since 1985 to agree or disagree with the following statement: “Human beings, as we know them today, developed from earlier species of animals (Miller et al. 2022)” During the last decade, the percentage of U.S. adults agreeing with this statement increased from 40% to 54%—a majority for the first time. This level of acceptance of evolution in the United States is atypically low for a developed nation. In a study of the acceptance of evolution in 34 developed nations in 2005, only Turkey—at 27%—scored lower than the United States (Miller, Scott and Okamoto 2006). There are also distinct differences among members of the U.S. population in terms of acceptance of evolution, with 68% of those ages 18–24, 58% of those with college degrees, and 65% of those who have taken four or more college-level science courses the most accepting of evolution. An increasing number of parents also report changing their unfavorable views of evolution due to helping their children with science homework and science fair projects.
Approaches to teaching evolution vary across the globe, with considerable differences within and between nations. One example is the United Kingdom, home of Darwin and Wallace. There, evolution is not introduced until ages 14–16, which is considered quite late by some educators. And, although evolution is taught in biology classes, it is addressed as a separate topic, rather than integrated into the curriculum as a foundational concept. As in the United States, there is also considerable variation between public and private schools, and between religious and secular institutions, in their treatment of the topic and the inclusion of alternative viewpoints, such as creationism (Harmon 2011). There are similar differences across the European Union, including within different populations in member countries.
There are also differences in the teaching of evolution across many predominantly Muslim nations. Salman Hameed, Professor of Integrated Science and Humanities at Hampshire College and Director of the Center for the Study of Science in Muslim Societies (SSiMS), whose research focuses on the acceptance of evolution among Muslims, has uncovered a great deal of variation among Muslims regarding beliefs about evolution. He points out that there is no central position within Islam regarding evolution, leaving it up to governments, textbook authors, and other entities to decide whether, and how, to address evolution in education. Saudi Arabia, Oman, Algeria, Morocco, and Lebanon all ban the teaching of evolution on religious grounds. Other Islamic nations, including Egypt, Malaysia, Syria, and Turkey, include evolutionary concepts like natural selection in their science curriculum but refrain from discussing human evolution (Asghar, Hameed, and Farahani 2014).
Impeding acceptance of evolution in science classes around the world is the adoption of textbooks from the E.U., the U.K., and the U.S. that include examples that are not culturally relevant to local populations (Harmon 2011). One classic example Hameed cites is that of the peppered moths in England whose predominant color pattern evolved from mostly white to mostly black due to pollutants darkening the tree bark of their habitat during the Industrial Revolution. Historical examples like this have little relevance for 21st-century students who grew up in non-Western countries and know little of England’s history or of the species that live there. It also privileges Western science over local science, to which many individuals in former European colonies and territories object (Jones 2017). Hameed suggests customizing textbooks to include local fossils and species whenever possible (Harmon 2011).
Are We Still Evolving?
After reading this chapter, many students are curious to know if humans are still evolving. The answer is yes. As a species, we continue to respond to selective pressures biologically and culturally. This final section will focus on three contemporary examples of human evolution. Before beginning, let’s review the conditions necessary for natural selection to operate on a trait. First, the trait must be heritable, meaning it is transmitted genetically from generation to generation. There must also be variation of the trait within the population and the trait must influence reproductive success. Three examples of traits that meet these criteria are immunity to the Human Immunodeficiency Virus (HIV), height, and wisdom teeth (Andrews, Kalinowski, and Leonard 2011).
AIDS is a potentially fatal infectious disease caused by HIV, a zoonosis believed to be derived from Simian Immunodeficiency Viruses (SIVs) found in chimpanzees and monkeys and most likely transmitted to humans through the butchering of infected animals (Sharp and Hahn 2011). In total, 40 million people have died from AIDS-related illnesses since the start of the global epidemic in the 1980s. There were 38.4 million people around the world living with AIDS as of 2021, including 1.5 million new cases and 650,000 deaths in that year alone (UNAIDS 2021). A disease causing this level of morbidity and mortality represents a major selective pressure, especially given that infection can occur before birth (Goulder et al. 2016), thereby affecting future reproductive success.
The majority of people in the world are highly susceptible to HIV infection, but some are not. These latter individuals are homozygous for a rare, recessive allele at the CCR5 locus that makes them immune to HIV. Heterozygotes who inherit a single copy of this allele are more resistant to infection and, when infected, the disease takes longer to progress in the event that they are infected. The mechanism by which the allele prevents infection involves a 32-base pair deletion in the DNA sequence of the CCR5 gene, creating a nonfunctioning receptor on the surface of the cell that prevents HIV from infecting the cell. The allele is inherited as a simple Mendelian trait, and there is variation in its prevalence, ranging as high as 14% of the population in northern Europe and Russia (Novembre, Galvani, and Slatkin 2005; see Figure 2.19). What is interesting about the allele’s geographic distribution is that it does not map onto parts of the world with the highest rates of HIV infection (Figure 2.20), suggesting that AIDS was not the original selective pressure favoring this allele (see Figures 2.19 and 2.20).
Given its current geographic distribution, the bubonic plague, which ravaged Europe repeatedly from the 14th to the 19th centuries (Pamuk 2007), was initially proposed as the selective agent. Subsequent research suggests smallpox, which killed up to 400,000 people annually in 18th-century Europe (Hays 2005), was more likely the selective pressure (Novembre, Galvani, and Slatkin 2005). Given the mortality rates for smallpox (Crosby 2003), an allele that conferred immunity was highly advantageous, as it is for those faced with the threat of HIV infection today.
Height is another example of a trait experiencing selective pressure. If you have ever toured a historical site, you have likely hit your head on a doorframe or become claustrophobic trying to squeeze down a narrow hallway under a lower-than-average ceiling. It is not your imagination. Humans have gotten taller in recent centuries. In fact, the average height of people in industrialized nations has increased approximately 10 centimeters (about four inches) in the past 150 years. This increase has been attributed to improvements in nutrition, sanitation, and access to medical care (Hatton 2014). But this is only part of the story.
Height is highly heritable. Studies indicate 80% of variation in height within populations is due to genetics, with 697 different genetic variances identified as having an effect on adult stature (Devuyst 2014). Multiple studies also demonstrate a positive relationship between height and reproductive success for men (Andrews, Kalinowsky, and Leonard 2011). This is primarily due to sexual selection and nonrandom mating, namely women’s preferences for taller men, which may explain why height is one characteristic men often lie about on dating websites (Guadagno, Okdie, and Kruse 2012). Sexual selection also plays out with regard to economic success in Western cultures, with taller men more likely to be in higher-level positions that pay well. Research demonstrates an inch of height is worth an additional $789 per year in salary, meaning a man who is six feet tall will earn on average $5,525 more per year than an identical man who is five foot five purely due to heightism bias (Gladwell 2007). Over the course of a career, this can add up to hundreds of thousands of dollars, likely allowing taller men to attract more potential mates, increasing their reproductive success.
Wisdom teeth are also undergoing evolutionary pressure. Have you or anyone in your family had their wisdom teeth removed? While it can be a painful and expensive process, it is a common experience in Western nations. It begs the question as to why there is no longer room in our mouths for all of our teeth? Biological anthropologist Daniel Lieberman offers several reasons, including that modern humans are growing faster and maturing earlier, which could be leading to impaction if skeletal growth takes place faster than dental growth. He also argues that the soft diets many modern humans consume generate insufficient strain to stimulate enough growth in our jaws to accommodate all of our teeth. Lastly, as the human brain has expanded over the past hundreds of thousands of years, it is taking up more space in the skull, causing the jaw to shrink, leaving no room for third molars (Lieberman 2011).
Conversely, do you know anyone whose wisdom teeth never came in? That is fairly common in some populations, suggesting evolutionary pressure favoring the absence of wisdom teeth has been culturally influenced. The oldest fossil evidence of skulls missing third molars was found in China and is 300,000 to 400,000 years old, suggesting the earliest mutation selecting against the eruption of wisdom teeth arose in Asia (Main 2013). Since that time, jaws have continued to decrease in size to the point they often cannot accommodate third molars, which can become impacted, painful, and even infected, a condition physical anthropologist Alan Main argues might have interfered with the ability to survive and reproduce in ancestral populations (Main 2013). As we have learned, a mutation that positively influences reproductive success—such as being born without the trait to develop wisdom teeth—would likely be selected for over time. Evidence in modern humans suggests that this is the case, with 40% of modern Asians and 45% of Native Alaskans and Greenlanders (populations descended from Asian populations) lacking wisdom teeth. The percentage among those of European descent ranges from 10 to 25% and for African Americans is 11% (Main 2013). Later chapters of this textbook emphasize that directional selection progresses along a particular path until the environment changes and a trait is no longer advantageous. In the case of wisdom teeth, the ability of modern dentistry to preempt impaction through surgery may, in fact, be what is preventing natural selection from doing away with wisdom teeth altogether.
Key Developments in Evolutionary Thought
4th century BCE |
Aristotle (384–322 BCE) |
“Founder of Biology.” Publishes History of Animals, a biological classification system of over 500 animals based on structure, physiology, reproduction, and behavior. Also creates the “Great Chain of Being,” ranking species and placing humans closest to God. |
8th–9th century CE |
Al-Jahiz (776–868 CE) |
Writes seven-volume Book of Animals, which includes animal classifications and food chains. Introduces concept of biological evolution and its mechanisms. |
1011–1021 |
Ibn al-Haythem (965–1040 CE) |
“Father of Modern Optics.” Uses experimental science to catalog how vision works and discovers laws of reflection and refraction. Publishes Book of Optics and invents camera obscura, the foundation for modern photography. |
1620 |
Francis Bacon (1561–1626) |
“Father of Empiricism.” Publishes The Novum Annum, formulating the scientific method based on observation and inductive reasoning. |
1686 |
John Ray (1627–1705) |
First to publish a biological definition of species in History of Plants. |
1749 |
Comte de Buffon (1707–1788) |
Publishes Histoire Naturelle, comparing anatomical structures across species using methods still in use today. Inspires Lamarck and Cuvier. |
1758 |
Carl von Linne (Carolus Linnaeus) (1707–1778) |
Introduces system of binomial nomenclature. Publishes Systema Naturae, the tenth edition of which introduces the designation Homo sapiens for humans. |
1788 |
James Hutton (1726–1797) |
“Father of Geology.” Publishes Theory of the Earth; introduces idea of Deep Time; explains how features of the earth were formed through the actions of rain, wind, rivers, and volcanic eruptions. |
1798 |
Thomas Malthus (1766–1834) |
Economist and “Father of Statistics.” Publishes An Essay on Population; introduces concept of carrying capacity; explains how populations outstrip the food supply, leaving some individuals to die off; inspires Darwin’s idea of “natural selection.” |
1809 |
Jean-Baptiste Lamarck (1744–1829) |
Publishes theory of the Inheritance of acquired characteristics; is the first Western scientist to propose a mechanism explaining how traits change in species over time and to recognize the importance of the physical environment in acting on species and their survival. |
1810 |
Georges Cuvier (1769–1832) |
Paleontologist/comparative anatomist; proved species went extinct; proposed the Theory of Catastrophism. |
1830 |
Charles Lyell (1797–1875) |
Establishes geology as a science. Publishes first edition of The Principles of Geology (1830–33); issuing 12 total editions in his lifetime, each updated according to new scientific data. |
1858 |
Alfred Russel Wallace (1823–1913) |
Sends scientific paper to Darwin titled “On the Tendency of Varieties to Depart Indefinitely from the Original Type,” essentially espousing the concept of natural selection; a reading of the papers by both Wallace and Darwin to the Linnaean Society is conducted by Lyell. |
1859 |
Charles Darwin (1809–1882) |
Publishes On the Origin of Species by Means of Natural Selection (1859). |
1865 |
Gregor Mendel (1822–1884) |
Publishes Experiments in Plant Hybridization (1865), outlining the fundamentals of genetic inheritance. |
1889 |
August Weismann (1834–1914) |
Publishes Essays Upon Heredity (1889), disproving the inheritance of acquired characteristics. Publishes The Germ Plasm (1892), postulating an early idea of inheritance through sexual reproduction. |
1937 |
Theodosius Dobzhansky (1900–1975) |
One of the founders of the Modern Synthesis of biology and genetics. Publishes Genetics and the Origin of Species (1937). Documents a genetic model of speciation through reproductive isolation. |
Review Questions
- Summarize the major scientific developments that led to the formulation of the theory of natural selection.
- Explain how natural selection operates and how it leads to evolution in populations.
- Explain the importance of genetics to an understanding of human evolution.
- Have you observed current examples of evolution taking place where you live? In which species? Which forces of evolution are involved?
Key Terms
Allele: A nonidentical DNA sequence found in the same gene location on a homologous chromosome, or gene copy, that codes for the same trait but produces a different phenotype.
Artificial selection: The identification by humans of desirable traits in plants and animals, and the subsequent steps taken to enhance and perpetuate those traits in future generations.
Binomial nomenclature: A system of classification in which a species of animal or plant receives a name consisting of two terms: the first identifies the genus to which it belongs, and the second identifies the species.
Carrying capacity: The number of living organisms, including animals, crops, and humans, that a geographic area can support without environmental degradation.
Catastrophism: The theory that the Earth’s geology has largely been shaped by sudden, short-lived, violent events, possibly worldwide in scope. Compare to uniformitarianism.
Comparative anatomy: Georges-Louis Leclerc’s technique of comparing similar anatomical structures across different species.
Creationism: The belief that the universe and all living organisms originate from specific acts of divine creation, as in the Biblical account, rather than by natural processes such as evolution.
Empiricism: The idea that all learning and knowledge derives from experience and observation. It became prominent in the 17th and 18th centuries in western Europe due to the rise of experimental science.
Evolution: In a biological sense, this term refers to cumulative inherited change in a population of organisms through time. More specifically, evolution is defined as a change in allele (gene) frequencies from one generation to the next among members of an interbreeding population.
Extant: Still in existence; surviving.
Extinct: Said of a species, family, or other group of animals or plants that has no living members; no longer in existence.
Fixity of Species: The idea that a species, once created, remains unchanged over time.
Gene: A sequence of DNA that provides coding information for the construction of proteins.
Genetic drift: Random changes in allele frequencies within a population from one generation to the next.
Gene flow: The introduction of new genetic material into a population through interbreeding between two distinct populations.
Gene pool: The entire collection of genetic material in a breeding community that can be passed from one generation to the next.
Genotype: The genotype of an organism is its complete set of genetic material—its unique sequence of DNA. Genotype also refers to the alleles or variants an individual carries in a particular gene or genetic location.
Hybrid: Offspring of parents that differ in genetically determined traits.
Intelligent design: A pseudoscientific set of beliefs based on the notion that life on earth is so complex that it cannot be explained by the scientific theory of evolution and therefore must have been designed by a supernatural entity.
Macroevolution: Large and often-complex changes in biological populations, such as species formation.
Microevolution: Changes in the frequency of a gene or allele in an interbreeding population.
Modern synthesis: The mid–20th century merging of Mendelian genetics with Darwinian evolution that resulted in a unified theory of evolution.
Natural selection: The natural process by which the survival and reproductive success of individuals or groups within an interbreeding population that are best adjusted to their environment leads to the perpetuation of genetic qualities best suited to that particular environment at that point in time.
Phenotype: The detectable or visible expression of an organism’s genotype.
Scientific method: A method of procedure that has characterized natural science since the 17th century, consisting of systematic observation, measurement, experimentation, and the formulation, testing, and modification of hypotheses.
Speciation: The process by which new genetically distinct species evolve from the main population, usually through geographic isolation or other barriers to gene flow.
Species: A group of living organisms consisting of similar individuals capable of exchanging genes or interbreeding. The species is the principal natural taxonomic unit, ranking below a genus and denoted by a Latin binomial (e.g., Homo sapiens).
Uniformitarianism: The theory that changes in the earth's crust during geologic history have resulted from the action of continuous and uniform processes—such as wind, precipitation, evaporation, condensation, erosion, and volcanic action—that continue to act in the present. Compare to catastrophism.
About the Author
Joylin Namie, Ph.D.
Truckee Meadows Community College, jnamie@tmcc.edu
Joylin Namie is Professor of Anthropology at Truckee Meadows Community College, where she teaches courses in biological and cultural anthropology. Her current research interest is in (un)sustainable tourism in desert environments, particularly in the country of Jordan and the U.S. state of Nevada. She was awarded a fellowship to Jordan from the Council of American Overseas Research Centers (CAORC) in 2020 to explore this topic, including visiting Petra and other important tourism destinations in Jordan. Dr. Namie’s favorite things in life are teaching, traveling, and spending time with her dog, Charley.
For Further Exploration
Costa, James T. 2017. Darwin’s Backyard: How Small Experiments Led to a Big Theory. New York: W.W. Norton.
Darwin, Charles. 1905. The Voyage of the Beagle. (Originally published in 1839 as Journal and Remarks). [Author’s note: Several editions exist with different publishers, including illustrated editions, paperback editions, and e-books.]
Moore, John A. 1993. Science as a Way of Knowing: The Foundations of Modern Biology. Cambridge, MA: Harvard University Press.
References
Al-Haytham, Ibn. 1011-1021. Kitāb al-Manāẓir (Book of Optics). Cairo, Egypt.
Al-Jahiz. 776–868 CE. Kitab al-Hayawan (Book of Animals).
Andrews, Tessa M., Steven T. Kalinowski, and Mary J. Leonard. 2011. “Are Humans Evolving? A Classroom Discussion to Change Students’ Misconceptions Regarding Natural Selection.” Evolution: Education and Outreach 4 (3): 456–466.
Aristotle. 384-322 BCE. History of Animals.
Asghar, Anila, Salman Hameed, and Najme Kashani Farahani. 2014. “Evolution in Biology Textbooks: A Comparative Analysis of Five Muslim Countries.” Religion & Education 41 (1). Accessed February 12, 2023. https://www.tandfonline.com/doi/abs/10.1080/15507394.2014.855081.
Associated Press. January 10, 2023. “Forest Lizards Have Genetically Morphed To Survive Life In The City, Researchers Say.” National Public Radio (NPR). Retrieved February 19, 2023 from https://www.npr.org/2023/01/10/1148150056/forest-lizards-genetically-morphed-cities.
Burkhardt, Richard W. 2013. “Lamarck, Evolution, and the Inheritance of Acquired Characters.” Genetics 194 (4): 793–805.
Chen, Xiqun, Danielle Feng, Michael A. Schwartzchild, and Xiang Gao. 2017. “Red Hair, MC1R Variants, and Risk for Parkinson’s Disease—A Meta-Analysis.” Annals of Clinical and Translational Neurology 4 (3): 212–216.https://doi.org/10.1002/acn3.381.
Colliss Harvey, Jacky. 2015. Red: A History of the Redhead. New York: Black Dog & Levanthal.
Corning, Peter A. 2020. “Beyond the Modern Synthesis: A Framework for a More Inclusive Biological Synthesis.” Progress in Biophysics and Molecular Biology 153: 5–12.
Costa, James T. 2017. Darwin’s Backyard: How Small Experiments Led to a Big Theory. New York: W. W. Norton.
Crosby, Alfred W., Jr. 2003. The Columbian Exchange: Biological and Cultural Consequences of 1492. 30th Anniversary Edition. Westport, CT: Praeger.
Darwin, Charles. 1859. On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life. First Edition. London: John Murray.
Darwin, Francis, ed. 2001[1897]. The Life & Letters of Charles Darwin, vol. 2. Honolulu: University Press of the Pacific.
Desilver, Drew. 2017. “U.S. Students’ Academic Achievement Still Lags That of Their Peers in Many Other Countries.” Pew Research Center, February 15. Accessed May 25, 2022. https://www.pewresearch.org/fact-tank/2017/02/15/u-s-students-internationally-math-science/.
Devuyst, Olivier. 2014. “High Time for Human Height.” Peritoneal Dialysis International 34 (7):685–686.
Dobzhansky, Theodosius. 1937. Genetics and the Origin of Species. Columbia University Biological Series (Volume 11). New York: Columbia University Press.
Dunbar-Ortiz, Roxanne. 2014. An Indigenous Peoples’ History of the United States. Boston: Beacon.
El-Zaher, Sumaya. 2018. “The Father of the Theory of Evolution: Al-Jahiz and His Book of Animals.” MVSLIM.com, October 9. Accessed August 27, 2022. https://mvslim.com/the-father-of-the-theory-of-evolution-al-jahiz-and-his-book-of-animals/.
Elshakry, Marwa. 2010. “When Science Became Western: Historiographical Reflections.” ISIS 101 (1). Accessed November 20, 2022. https://www.journals.uchicago.edu/doi/10.1086/652691.
Gladwell, Malcolm. 2007. Blink: The Power of Thinking without Thinking. New York: Back Bay
Guadagno, Rosanna E., Bradley M. Okdie, and Sara A. Kruse. 2012. “Dating Deception: Gender, Online Dating, and Exaggerated Self-Presentation.” Computers in Human Behavior 28 (2): 642–647.
Harmon, Katherine. 2011. “Evolution Abroad: Creationism Evolves in Science Classrooms around the Globe.” Scientific American, March 3. Accessed May 25, 2022. https://www.scientificamerican.com/article/evolution-education-abroad/.
Harrison, Ellie. 2018. “This is What the Perfect Body Looks Like - According to Science.” Radiotimes.com. Accessed June 14, 2023. https://www.radiotimes.com/tv/documentaries/this-is-what-the-perfect-body-looks-like-according-to-science/.
Hatton, Tim. 2014. “Why Did Humans Grow Four Inches in 100 Years? It Wasn’t Just Diet.” The Conversation, May 1. https://theconversation.com/why-did-humans-grow-four-inches-In-100-years-it-wasnt-just-diet-25919.
Hays, J. N. 2005. Epidemics and Pandemics: Their Impacts on Human History. Santa Barbara, CA: ABC-CLIO.
Hutton, James. 1788. Theory of the Earth. Transactions of the Royal Society of Edinburgh, Volume 1. Scotland: Royal Society of Edinburgh.
IPCC. 2022. Climate Change 2022: Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, edited by H.-O. Pörtner, D.C. Roberts, M. Tignor, E.S. Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem, and B. Rama. Cambridge: Cambridge University Press. https://www.ipcc.ch/report/ar6/wg2/.
Jones, Stephen. 2017. “Religion, Science, and Evolutionary Theory”[an interview with Salman Hameed]. Podcast, The Religious Studies Project, January 30. Accessed May 29, 2023. https://www.religiousstudiesproject.com/podcast/religion-science-and-evolutionary-theory/.
Kutschera, Ulrich, and Karl J. Niklas. 2004. “The Modern Theory of Biological Evolution: An Expanded Synthesis.” Naturwissenschaften 91: 255–276.
Leclerc, Georges-Louis, Comte de Buffon. 1749-1804. Histoire Naturelle. Volumes 1-36. Paris: Imprimerie Royale (Royal Printing House).
Lents, Nathan H. 2018. Human Errors: A Panorama of Our Glitches, from Pointless Bones to Broken Genes. Boston: Houghton Mifflin Harcourt.
Lerner, Lawrence S. 2000. Good Science, Bad Science: Teaching Evolution in the States. Washington, DC: Thomas B. Fordham Foundation.
Lieberman, Daniel E. 2011. The Evolution of the Human Head. Cambridge, MA: Harvard University Press.
Lindberg, David C. 1992. The Beginnings of Western Science: The European Scientific Tradition in Philosophical, Religious, and Institutional Context, 600 B.C. to A.D. 1450. Chicago: The University of Chicago Press.
Linnaeus, Carl. 1736. Systema Naturae. First Edition. Stockholm: Laurentius Salvius.
Love, Shayla. 2020. “A Thousand Years before Darwin, Islamic Scholars Were Writing about Evolution.” VICE World News, October 5. Accessed August 27, 2022. https://www.vice.com/en/article/ep4ykn/a-thousand-years-before-darwin-islamic-scholars-were-writing-about-natural-selection.
Main, Douglas. 2013. “Ancient Mutation Explains Missing Wisdom Teeth.” Live Science, March 13. https://www.livescience.com/27529-missing-wisdom-teeth.html.
Malthus, Thomas Robert. 1798. An Essay on the Principle of Population. London: J. Johnson.
Masci, David. 2019. “For Darwin Day, 6 Facts About the Evolution Debate.” Pew Research Center. Accessed June 14, 2023. https://www.pewresearch.org/short-reads/2019/02/11/darwin-day/.
Masood, Ehsan. 2009. “Islam’s Evolutionary Legacy.” The Guardian, March 1. Accessed February 13, 2023. https://www.theguardian.com/commentisfree/belief/2009/feb/27/islam-religion-evolution-science.
Miller, Jon D., Eugenie C. Scott, Mark S. Ackerman, Belen Laspra, Glenn Branch, Carmelo Polino, and Jordan S. Huffaker. 2022. “Public Acceptance of Evolution in the United States, 1985–2020.” Public Understanding of Science 31 (2): 223–238.
Miller, Jon D., Eugenie C. Scott, and Shinji Okamoto. 2006, August 11. “Public Acceptance of Evolution.” Science 313 (5788): 765-766. DOI: 10.1126/science.1126746.
Moore, John A. 1993. Science as a Way of Knowing: The Foundations of Modern Biology. Cambridge, MA: Harvard University Press.
Moore, Michael P., Kaitlyn Hersch, Chanont Sricharoen, Sarah Lee, Caitlin Reice, Paul Rice, Sophie Kronick, Kim A. Medley, and Kasey D. Fowler-Finn. 2021. “Sex-Specific Ornament Evolution Is a Consistent Feature of Climatic Adaptation across Space and Time in Dragonflies.” PNAS 118 (28): e2101458118. https://doi.org/10.1073/pnas.210145818.
National Public Radio. 2023. “Forest Lizards Have Genetically Morphed to Survive Life in the City, Researchers Say.” National Public Radio (NPR), January 10. Accessed February 19, 2023. https://www.npr.org/2023/01/10/1148150056/forest-lizards-genetically-morphed-cities.
National Science Foundation. 2023. “Urban Lizards Share Genomic Markers Not Found in Forest-Dwellers.” National Science Foundation, February 7. Accessed May 25, 2023. https://beta.nsf.gov/news/urban-lizards-share-genomic-markers-not-found.
Novembre, John, Alison P. Galvani, and Montgomery Slatkin. 2005. “The Geographic Spread of the CCR5 Δ32 HIV-Resistance Allele.” PLoS Biology 3 (11): e339.
Pamuk, Şevket. 2007. “The Black Death and the Origins of the ‘Great Divergence’ across Europe, 1300–1600.” European Review of Economic History 11 (3): 289–317.
Paterlini, Marta. 2007. “There Shall Be Order: The Legacy of Linnaeus in the Age of Molecular Biology.” EMBO Reports 8 (9): 814–816. https://doi.org/10.1038/sj.embor.7401061.
Pew Research Center. 2019. “Darwin in America: The Evolution Debate in the United States.” Pew Research Center, February 6, 2019. Accessed June 13, 2023. https://www.pewresearch.org/religion/2019/02/06/darwin-in-america-2/.
Pew Research Center. 2014. “Fighting Over Darwin, State by State.” Pew Research Center, February 4, 2009; updated February 3, 2014. Accessed May 25, 2022. https://www.pewresearch.org/religion/2009/02/04/fighting-over-darwin-state-by-state/.
Plutynski, Anya. 2009. “The Modern Synthesis.” Routledge Encyclopedia of Philosophy, November 15 (updated December 14, 2009). Accessed November 27, 2022. https://philsci-archive.pitt.edu/15335/.
Plutzer, Eric, Glenn Branch, and Ann Reid. 2020. “Teaching Evolution in the U.S. Public Schools: A Continuing Challenge.” Evolution: Education and Outreach13: Article 14. https://evolution-outreach.biomedcentral.com/articles/10.1186/s12052-020-00126-8.
Proceedings of the National Academy of Sciences (PNAS). 2023. “How Lizards Adapt to Urban Living.” Science Sessions Podcast, February 13, . Accessed February 19, 2023. https://www.pnas.org/post/podcast/lizards-adapt-urban-living.
Public Broadcasting System. 2001. “Georges Cuvier.” WGBH Evolution Library. Accessed May 26, 2022. https://www.pbs.org/wgbh/evolution/library/02/1/l_021_01.html# :~:text=With%20elegant%20studies%20of%20the,as%20incontrovertible%20proof%20of %20extinctions.
Ray, John. 1686-1704. Historia plantarum. London: Clark. Volume 1 (1686), Volume II (1688), Volume III (1704).
Richards, Richard A. 1998. “Darwin, Domestic Breeding, and Artificial Selection.” Endeavour 22 (3): 106–109.
Solloch, Ute V., Kathrin Lang, Vinzenz Lange, and Irena Böhme. 2017. “Frequencies of Gene Variant CCR5-Δ32 in 87 Countries Based on Next-Generation Sequencing of 1.3 Million Individuals Sampled from 3 National DKMS Donor Centers.” Human Immunology 78 (11–12): 701-717.
Stewart, Thomas A., Justin B. Lemberg, Ailis Daly, Edward B. Daeschler, and Neil H. Shubin. 2022. “A New Epistostegalian from the Late Devonian of the Canadian Arctic.” Nature 608 (7923): 563–568.
Tasci, Ufuk Necat. 2020. “How a 10-Century Muslim Physicist Discovered How Humans See.” TRT World, May 25. Accessed May 18, 2022. https://www.trtworld.com/magazine/how-a-10th-century-muslim-physicist-discovered-how-humans-see-36620.
Tbakhi, Abdelghani, and Samir S. Amr. 2007. “Ibn Al-Haytham: Father of Modern Optics.” Annals of Saudi Medicine 27 (6): 464–467. UNAIDS. 2021. “Global HIV & AIDS Statistics—Fact Sheet.” https://www.unaids.org/en/resources/fact-sheet.
UNESCO.org. 2015. “International Year of Light: Ibn al Haytham, Pioneer of Modern Optics celebrated at UNESCO.” UNESCO, September 8. Accessed May 18, 2022. https://www.unesco.org/en/articles/international-year-light-ibn-al-haytham-pioneer-modern-optics-celebrated-unesco.
University of California Berkeley Museum of Paleontology. N.d. “The History of Evolutionary Thought—Uniformitarianism: Charles Lyell.” Understanding Evolution website. Accessed February 13, 2023. https://evolution.berkeley.edu/the-history-of-evolutionary-thought/1800s/uniformitarianism-charles-lyell/.
University of Cambridge. N.d. “Darwin and Dogs.” The Darwin Correspondence Project website. Accessed February 17, 2023. https://www.darwinproject.ac.uk/commentary/curious/darwin-and-dogs.
Urbach, Peter Michael, Anthony M. Quinton, and Kathleen Marguerite Lea. “Francis Bacon: British Author, Philosopher, and Statesman.” Encyclopaedia Britannica. Last updated May 12, 2023. https://www.britannica.com/biography/Francis-Bacon-Viscount-Saint-Alban.
Ward, Peter. 2018. Lamarck’s Revenge: How Epigenetics Is Revolutionizing Our Understanding of Evolution’s Past and Present. New York: Bloomsbury.
Weismann, August. 1892. Das Keimplasma: Eine Theorie der Vererbung (The Germ Plasm: a Theory of Inheritance). Jena (Germany): Fischer.
Weismann, August. 1889. Translations. Essays upon Heredity. Oxford: Clarendon. Accessed November 27, 2022. E-copy available at https://www.esp.org/books/weismann/essays/facsimile/.
Winchell, Kristen M., Shane C. Campbell-Staton, Jonathan B. Losos, and Anthony Geneva. 2023. “Genome-Wide Parallelism Underlies Contemporary Adaptation In Urban Lizards.” PNAS 120 (3): e2216789120. https://doi.org/10.1073/pnas.2216789120.
Wood, Bernard. 2005. Human Evolution: A Very Short Introduction. Oxford: Oxford University Press.
Zou, Yawen. 2015. "The Germ-Plasm: a Theory of Heredity (1893), by August Weismann." Embryo Project Encyclopedia, January 26. Accessed February 18, 2023. https://embryo.asu.edu/handle/10776/8284.
Ashley Kendell, Ph.D., California State University, Chico
Alex Perrone, M.A., M.S.N, R.N., P.H.N., Butte Community College
Colleen Milligan, Ph.D., California State University, Chico
This chapter is a revision from "Chapter 15: Bioarchaeology and Forensic Anthropology” by Ashley Kendell, Alex Peronne, and Colleen Milligan. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Content Warning and Disclaimer: This chapter includes images of human remains as well as discussions centered on human skeletal analyses. All images are derived from casts, sketches, nonhuman skeletal material, as well as non-Indigenous skeletal materials curated within the CSU, Chico Human Identification Lab, and the Hartnett-Fulginiti donated skeletal collection.
Learning Objectives
- Define forensic anthropology as a subfield of biological anthropology.
- Describe the seven steps carried out during skeletal analysis.
- Outline the four major components of the biological profile.
- Contrast the four categories of trauma.
- Explain how to identify the different taphonomic agents that alter bone.
- Discuss ethical considerations for forensic anthropology.
Forensic anthropology is a subfield of biological anthropology and an applied area of anthropology. Forensic anthropologists use skeletal analysis to gain information about humans in the present or recent past, then they apply this information within a medicolegal context. This means that forensic anthropologists specifically conduct their analysis on recently deceased individuals (typically within the last 50 years) as part of investigations by law enforcement. Forensic anthropologists can assist law enforcement agencies in several different ways, including aiding in the identification of human remains whether they are complete, fragmentary, burned, scattered, or decomposed. Additionally, forensic anthropologists can help determine what happened to the deceased at or around the time of death as well as what processes acted on the body after death (e.g., whether the remains were scattered by animals, whether they were buried in the ground, or whether they remained on the surface as the soft tissue decomposed).
Many times, because of their expertise in identifying human skeletal remains, forensic anthropologists are called to help with outdoor search-and-recovery efforts, such as locating remains scattered across the surface or carefully excavating and documenting buried remains. In other cases, forensic anthropologists recover remains after natural disasters or accidents, such as fire scenes, and can help identify whether each bone belongs to a human or an animal. Forensic anthropology spans a wide scope of contexts involving the law, including incidences of mass disasters, genocide, and war crimes.
A point that can be somewhat confusing for students is that although the term forensic is included in this subfield of biological anthropology, there are many forensic techniques that are not included in the subfield. Almost exclusively, forensic anthropology deals with skeletal analysis. While this can include the comparison of antemortem (before death) and postmortem (after death) radiographs to identify whether remains belong to a specific person, or using photographic superimposition of the cranium, it does not include analyses beyond the skeleton. For example, blood-spatter analysis, DNA analysis, fingerprints, and material evidence collection do not fall under the scope of forensic anthropology.
So, what can forensic anthropologists glean from bones alone? Forensic anthropologists can address a number of questions about a human individual based on their skeletal remains. Some of those questions are as follows: How old was the person? Was the person biologically male or female? How tall was the person? What happened to the person at or around their time of death? Were they sick? The information from the skeletal analysis can then be matched with missing persons records, medical records, or dental records, aiding law enforcement agencies with identifications and investigations.
Skeletal Analysis
Forensic anthropology relies on skeletal analysis to reveal information about the deceased. The methodology and approaches outlined below are specific to the United States. Forensic anthropological methods differ depending on the country conducting an investigation. In the United States, there are typically seven steps or questions to the process:
- Is it bone?
- Is it human?
- Is it modern or archeological?
- How many individuals are present or what is the minimum number of individuals (MNI)?
- Who is it?
- Is there evidence of trauma before or around the time of death?
- What happened to the remains after death?
Is It Bone?
One of the most important steps in any skeletal analysis starts with determining whether or not material suspected to be bone is in fact bone. Though it goes without saying that a forensic anthropologist would only carry out analysis on bone, this step is not always straightforward. Whole bones are relatively easy to identify, but determining whether or not something is bone becomes more challenging once it becomes fragmentary. As an example, in high heat such as that seen on fire scenes, bone can break into pieces. During a house fire with fatalities, firefighters watered down the burning home. After the fire was extinguished, the sheetrock (used to construct the walls of the home) was drenched and crumbled. The crumbled sheetrock was similar in color and form to burned, fragmented bone, therefore mistakable for human remains (Figure 15.1). Forensic anthropologists on scene were able to separate the bones from the construction material, helping to confirm the presence of bone and hence the presence of individual victims of the fire. In this case, forensic anthropologists were able to recognize the anatomical and layered structure of bone and were able to distinguish it from the uniform and unlayered structure of sheetrock.
As demonstrated by the example above, both the macrostructure (visible with the naked eye) and microstructure (visible with a microscope) of bone are helpful in bone identification. Bones are organs in the body made up of connective tissue. The connective tissue is hardened by a mineral deposition, which is why bone is rigid in comparison to other connective tissues such as cartilage (Tersigni-Tarrant and Langley 2017, 82–83; White and Folkens 2005, 31). In a living body, the mineralized tissue does not make up the only component of bone—there are also blood, bone marrow, cartilage, and other types of tissues. However, in dry bone, two distinct layers of the bone are the most helpful for identification. The outer layer is made up of densely arranged osseous (bone) tissue called compact (cortical) bone. The inner layer is composed of much more loosely organized, porous bone tissue whose appearance resembles that of a sponge, hence the name spongy (trabecular) bone. Knowing that most bone contains both layers helps with the macroscopic identification of bone (Figures 15.2, 15.3). For example, a piece of coconut shell might look a lot like a fragment of a human skull bone. However, closer inspection will demonstrate that coconut shell only has one very dense layer, while bone has both the compact and spongy layers.
The microscopic identification of bone relies on knowledge of osteons, or bone cells (Figure 15.4). Under magnification, bone cells are visible in the outer, compact layer of bone. The bone cells are arranged in a concentric pattern around blood vessels for blood supply. The specific shape of the cells can help differentiate, for example, a small piece of PVC (white plastic) pipe from a human bone fragment (Figure 15.5).
Is It Human?
Once it has been determined that an object is bone, the next logical step is to identify whether the bone belongs to a human or an animal. Forensic anthropologists are faced with this question in everyday practice because human versus nonhuman bone identification is one of the most frequent requests they receive from law enforcement agencies.
There are many different ways to distinguish human versus nonhuman bone. The morphology (the shape/form) of human bone is a good place for students to start. Identifying the 206 bones in the adult human skeleton and each bone’s distinguishing features (muscle attachment sites, openings and grooves for nerves and blood vessels, etc.) is fundamental to skeletal analysis.
Nevertheless, there are many animal bones and human bones that look similar. For example, the declawed skeleton of a bear paw looks a lot like a human hand, pig molars appear similar to human molars, and some smaller animal bones might be mistaken for those of an infant. To add to the confusion, fragmentary bone may be even more difficult to identify as human or nonhuman. However, several major differences between human and nonhuman vertebrate bone help distinguish the two.
Forensic anthropologists pay special attention to the density of the outer, compact layer of bone in both the cranium and in the long bones. Human cranial bone has three distinctive layers. The spongy bone is sandwiched between the outer (ectocranial) and inner (endocranial) compact layers. In most other mammals, the distinction between the spongy and compact layers is not always so definite. Secondly, the compact layer in nonhuman mammal long bones can be much thicker than observed in human bone. Due to the increased density of the compact layer, nonhuman bone tends to be heavier than human bone (Figure 15.6).
The size of a bone can also help determine whether it belongs to a human. Adult human bones are larger than subadult or infant bones. However, another major difference between human adult bones and those of a young individual or infant human can be attributed to development and growth of the epiphyses (ends of the bone). The epiphyses of human subadult bones are not fused to the shaft (Figure 15.7). Therefore, if a bone is small and it is suspected to belong to a human subadult or infant, the epiphyses would not be fused. Many small animal bones appear very similar in form compared to adult human bones, but they are much too small to belong to an adult human. Yet they can be eliminated as subadult or infant bones if the epiphyses are fused to the shaft.
Is It Modern or Archaeological?
Forensic anthropologists work with modern cases that fall within the scope of law enforcement investigations. Accordingly, it is important to determine whether discovered human remains are archaeological or forensic in nature. Human remains that are historic are considered archeaological. The scientific study of human remains from archaeological sites is called bioarchaeology.
Dig Deeper: Bioarchaeology
For readers who are interested in the sister subfield of bioarchaeology, which studies human remains and material culture from the past, please refer to chapter 8 of Bioarchaeology: Interpreting Human Behavior from Skeletal Remains, in TRACES: An Open Invitation to Archaeology (Blatt, Michael, and Bright forthcoming).
A forensic anthropologist should begin their analysis by reviewing the context in which the remains were discovered. This will help them understand a great deal about the remains, including determining whether they are archaeological or forensic in nature as well as considering legal and ethical issues associated with the collection, analysis, and storage of human remains (see “Ethics and Human Rights” section of this chapter for more information).
The “context” refers to the relationship the remains have to the immediate area in which they were found. This includes the specific place where the remains were found, the soil or other organic matter immediately surrounding the remains, and any other objects or artifacts in close proximity to the body. For example, imagine that a set of remains has been located during a house renovation. The remains are discovered below the foundation. Do the remains belong to a murder victim? Or was the house built on top of an ancient burial ground? Observing information from the surroundings can help determine whether the remains are archaeological or modern. How long ago was the foundation of the house erected? Are there artifacts in close proximity to the body, such as clothing or stone tools? These are questions about the surroundings that will help determine the relative age of the remains.
Clues directly from the skeleton may also indicate whether the remains are archaeological or modern. For example, tooth fillings can suggest that the individual was alive recently (Figure 15.8). In fact, filling material has changed over the decades, so the specific type of material used to fix a cavity can be matched with specific time periods. Gold was used in dental work in the past, but more recently composite (a mixture of plastic and fine glass) fillings have become more common.
How Many Individuals Are Present?
What Is MNI?
Another assessment that an anthropologist can perform is the calculation of the number of individuals in a mixed burial assemblage. Because not all burials consist of a single individual, it is important to burial assemblage be able to estimate the number of individuals in a forensic context. Quantification of the number of individuals in a burial assemblage can be done through the application of a number of methods, including the following: the Minimum Number of Individuals (MNI), the Most Likely Number of Individuals (MLNI), and the Lincoln Index (LI). The most commonly used method in biological anthropology, and the focus of this section, is determination of the MNI.
The MNI presents “the minimum estimate for the number of individuals that contributed to the sample” (Adams and Konigsberg 2008, 243). Many methods of calculating MNI were originally developed within the field of zooarchaeology for use on calculating the number of individuals in faunal or animal assemblages (Adams and Konigsberg 2008, 241). What MNI calculations provide is a lowest possible count for the total number of individuals contributing to a skeletal assemblage. Traditional methods of calculating MNI include separating a skeletal assemblage into categories according to the individual bone and the side the bone comes from and then taking the highest count per category and assigning that as the minimum number (Figure 15.9).
Why Calculate MNI?
In a forensic context, the determination of MNI is most applicable in cases of mass graves, commingled burials, and mass fatality incidents. The term commingled is applied to any burial assemblage in which individual skeletons are not separated into separate burials. As an example, the authors of this chapter have observed commingling of remains resulting from mass fatality wildfire events. Commingled remains may also be encountered in events such as a plane or vehicle crash. It is important to remember that in any forensic context, MNI should be referenced and an MNI of one should be substantiated by the fact that there was no repetition of elements associated with the case.
Constructing the Biological Profile
Who Is It?
“Who is it?” is one of the first questions that law enforcement officers ask when they are faced with a set of skeletal remains. To answer this question, forensic anthropologists construct a biological profile (White and Folkens 2005, 405). A biological profile is an individual’s identifying characteristics, or biological information, which include the following: biological sex, age at death, stature, population affinity, skeletal variation, and evidence of trauma and pathology.
Forensic anthropologists typically construct a biological profile to help positively identify a deceased person. The following section will lay out each component of the biological profile and briefly review standard methodology used for each.
Assessing Biological Sex
Assessment of biological sex is often one of the first things considered when establishing a biological profile because several other parts, such as age and stature estimations, rely on an assessment of biological sex to make the calculations more accurate.
Assessment of biological sex focuses on differences in both morphological (form or structure) and metric (measured) traits in individuals. When assessing morphological traits, the skull and the pelvis are the most commonly referenced areas of the skeleton. These differences are related to sexual dimorphism usually varying in the amount of robusticity seen between males and females. Robusticity deals with strength and size; it is frequently used as a term to describe a large size or thickness. In general, males will show a greater degree of robusticity than females. For example, the length and width of the mastoid process, a bony projection located behind the opening for the ear, is typically larger in males. The mastoid process is an attachment point for muscles of the neck, and this bony projection tends to be wider and longer in males. In general, cranial features tend to be more robust in males (Figure 15.10).
When considering the pelvis, the features associated with the ability to give birth help distinguish females from males. During puberty, estrogen causes a widening of the female pelvis to allow for the passage of a baby. Several studies have identified specific features or bony landmarks associated with the widening of the hips, and this section will discuss one such method. The Phenice Method (Phenice 1969) is traditionally the most common reference used to assess morphological characteristics associated with sex. The Phenice Method specifically looks at the presence or absence of (1) a ventral arc, (2) the presence or absence of a subpubic concavity, and (3) the width of the medial aspect of the ischiopubic ramus (Figure 15.11). When present, the ventral arc, a ridge of bone located on the ventral surface of the pubic bone, is indicative of female remains. Likewise the presence of a subpubic concavity and a narrow medial aspect of the ischiopubic ramus is associated with a female sex estimation. Assessments of these features, as well as those of the skull (when both the pelvis and skull are present), are combined for an overall estimation of sex.
Metric analyses are also used in the estimation of sex. Measurements taken from every region of the body can contribute to estimating sex through statistical approaches that assign a predictive value of sex. These approaches can include multiple measurements from several skeletal elements in what is called multivariate (multiple variables) statistics. Other approaches consider a single measurement, such as the diameter of the head of the femur, of a specific element in a univariate (single variable) analysis (Berg 2017, 152–156).
It is important to note that, although forensic anthropologists usually begin assessment of biological profile with biological sex, there is one major instance in which this is not appropriate. The case of two individuals found in California, on July 8, 1979, is one example that demonstrates the effect age has on the estimation of sex. The identities of the two individuals were unknown; therefore, law enforcement sent them to a lab for identification. A skeletal analysis determined that the remains represented one adolescent male and one adolescent female, both younger than 18 years of age. This information did not match with any known missing children at the time.
In 2015, the cold case was reanalyzed, and DNA samples were extracted. The results indicated that the remains were actually those of two girls who went missing in 1978. The girls were 15 years old and 14 years old at the time of death. It is clear that the 1979 results were incorrect, but this mistake also provides the opportunity to discuss the limitations of assessing sex from a subadult skeleton.
Assessing sex from the human skeleton is based on biological and genetic traits associated with females and males. These traits are linked to differences in sexual dimorphism and reproductive characteristics between females and males. The link to reproductive characteristics means that most indicators of biological sex do not fully manifest in prepubescent individuals, making estimations of sex unreliable in younger individuals (SWGANTH 2010b). This was the case in the example of the 14-year-old girl. When examined in 1979, her remains were misidentified as male because she had not yet fully developed female pelvic traits.
Sex vs. Gender
Biological sex is a different concept than gender. While biological anthropologists can estimate sex from the skeleton, estimating an individual’s gender would require a greater context because gender is defined culturally rather than biologically. Take, for example, an individual who identifies as transgender. This individual has a gender identity that is different from their biological sex. The gender identity of any individual depends on factors related to self-identification, situation or context, and cultural factors. While in the U.S. we have historically thought of sex and gender as binary concepts (male or female), many cultures throughout the world recognize several possible gender identities. In this sense, gender is seen as a continuous or fluid variable rather than a fixed one.
Historically, forensic anthropologists have used a binary construct to categorize human skeletal remains as either male or female (with the accompanying categories of probable male, probable female, and indeterminate). In the case of transgender and gender nonconforming individuals, the binary approach to sex assessment may delay or hinder identification efforts (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020; Tallman, Kincer, and Plemons 2021). As such, many forensic anthropologists have begun to address the inherent problems associated with a binary approach to sex identification and to explore ways of assessing social identity and self-identified gender using skeletal remains and forensic context.
For the duration of this section, the term transgender refers to individuals whose gender identity differs from the sex assigned at birth (Schall, Rogers, and Deschamps-Braly 2020:2). Transgender individuals transition from one gender binary to another, such as male-to-female (MTF) or female-to-male (FTM). While many of the gender-affirming procedures available to trans and gender-nonconforming individuals are focused on soft tissue modifications (e.g., breast augmentation, genital reconstruction, hormone therapies, etc.), there are a number of gender-affirmation surgeries that do leave a permanent record on the skeleton. Generally speaking, FTM transgender people are reported to undergo fewer surgical procedures than do MTF transgender people (Buchanan 2014). The discussion below focuses on Facial Feminization Surgery (FFS), which leaves a permanent record on the human skeleton that may be used to help make an identification.
FFS refers to a combination of procedures focused on sexually dimorphic features of the face, with the intent of transforming typically male facial features into more feminine forms. Facial Feminization Surgery procedures were developed by Dr. Douglas Ousterhout, a San Francisco based cranio-maxillofacial surgeon, in the mid-1980s (Schall, Rogers, and Deschamps-Braly 2020:2). FFS can include one or a combination of the following: hairline lowering, forehead reduction and contouring, brow lift, reduction rhinoplasty, cheek enhancement, lift lift, lip filling, chin contouring, jaw contouring, and/or tracheal shave (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020:2). Of the procedures outlined previously, four are known to directly affect the facial skeleton: forehead contouring, rhinoplasty, chin contouring, and jaw contouring (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020:2).
Because FFS procedures have been widely documented in the medical (and more recently the forensic anthropological) literature, there are a number of indicators that a forensic anthropologist can use to make more informed evaluations of gender, including evidence of bone remodeling in sexually dimorphic regions of the skull (e.g., forehead, chin, jawline), as well as the presence of plates, pins, or other surgical hardware that may be evidence of FFS (Buchanan 2014; Schall, Rogers, and Deschamps-Braly 2020; Tallman, Kincer, and Plemons 2021). Additionally, some forensic anthropologists suggest cautiously integrating contextual information from the scene, such as personal effects, material evidence, and recovery scene information, into their evaluation of an individual’s social identity (Beatrice and Soler 2016; Birkby, Fenton, and Anderson 2008; Soler and Beatrice 2018; Soler et al. 2019; Tallman, Kincer, and Plemons 2021; Winburn, Schoff, and Warren 2016). The ultimate goal of many skeletal analyses is to make a positive identification on a set of unidentified remains.
Assessment of Population Affinity
Population affinity is another component of the biological profile. We use the term population affinity to refer to the variation seen among modern populations—variation that is both genetic and environmentally driven. The word affinity refers to similarities or relationships between individuals. As forensic anthropologists, we compare an unknown individual to multiple reference groups and look for the degree of similarity in observable traits with those groups. As noted previously, population affinity can aid law enforcement in their identification of missing persons or unknown skeletal remains.
Within the field of anthropology, the estimation of population affinity has a contentious history, and early attempts at classification were largely based on the erroneous assumption that an individual’s phenotype (outward appearance) was correlated with their innate intelligence and abilities (see Chapter 13 for a more in-depth discussion of the history of the race concept). The use of the term race is deeply embedded in the social context of the United States. In any other organism/living thing, groups divided according to the biological race concept would be defined as a separate subspecies. The major issue with applying the biological race concept to humans is that there are not enough differences between any two populations to separate on a genetic basis. In other words, biological races do not exist in human populations. However, the concept of race has been perpetuated and upheld by sociocultural constructs of race.
The conundrum for forensic anthropologists is the fact that while races do not exist on a biological level, we still socially recognize and categorize individuals based on their phenotype. Clearly, our phenotype is an important factor in not only how we are viewed by others but also how we identify ourselves. It is also a commonly reported variable. Often labeled as “race,” we are asked to report how we self-identify on school applications, government identification, surveys, census reports, and so forth. It follows then that when a person is reported missing, the information commonly collected by law enforcement and sometimes entered into a missing person’s database includes their age, biological sex, stature, and “race.” Therefore, the more information a forensic anthropologist can provide regarding the individual’s physical characteristics, the more he or she can help to narrow the search.
As an exercise, create a list of all of the women you know who are between the ages of 18 and 24 and approximately 5’ 4” to 5’ 9” tall. You probably have several dozen people on the list. Now, consider how many females you know who are between the ages of 18 and 24, are approximately 5’ 4” to 5’ 9” tall, and are Vietnamese. Your list is going to be significantly shorter. That’s how missing persons searches go as well. The more information you can provide regarding a decedent’s phenotype, the fewer possible matches law enforcement are left to investigate. This is why population affinity has historically been included as a part of the biological profile.
(Put at the beginning of this serction)In an effort to combat the erroneous assumptions tied to the race concept, forensic anthropologists have attempted to reframe this component of the biological profile. The term race is no longer used in casework and teaching. Historically, the word ancestry is and was deemed a more appropriate way to describe an individual’s phenotype. However, in more recent years, forensic anthropologists have begun using the term population affinity, recognizing that we are basing our analysis on the similarities we see based on the reference samples we have available (Winburn and Algee-Hewitt 2021). An important note here is that it is possible to hinder identifications and harm individuals when tools like estimations of population affinity are misapplied, misinterpreted, or misused. For this reason, the field of forensic anthropology has ongoing conversations about the appropriateness of this analysis in the biological profile (Bethard and DiGangi 2020; Stull et al. 2021).
Traditionally, population affinity was accomplished through a visual inspection of morphological variants of the skull (morphoscopics). These methods focused on elements of the facial skeleton, including the nose, eyes, and cheek bones. However, in an effort to reduce subjectivity, nonmetric cranial traits are now assessed within a statistical framework to help anthropologists better interpret their distribution among living populations (Hefner and Linde 2018). Based on the observable traits, a macromorphoscopic analysis will allow the practitioner to create a statistical prediction of geographic origin. In essence, forensic anthropologists are using human variation in the estimation of geographic origin, by referencing documented frequencies of nonmetric skeletal indicators or macromorphoscopic traits.
Population affinity is also assessed through metric analyses. The computer program Fordisc is an anthropological tool used to estimate different components of the biological profile, including ancestry, sex, and stature. When using Fordisc, skeletal measurements are input into the computer software, and the program employs multivariate statistical classification methods, including discriminant function analysis, to generate a statistical prediction for the geographic origin of unknown remains based on the comparison of the unknown to the reference samples in the software program. Fordisc also calculates the likelihood of the prediction being correct, as well as how typical the metric data is for the assigned group.
Estimating Age-at-Death
Estimating age-at-death from the skeleton relies on the measurement of two basic physiological processes: (1) growth and development and (2) degeneration (or aging). From fetal development on, our bones and teeth grow and change at a predictable rate. This provides for relatively accurate age estimates. After our bones and teeth cease to grow and develop, they begin to undergo structural changes, or degeneration, associated with aging. This does not happen at such predictable rates and, therefore, results in less accurate or larger age-range estimations.
During growth and development stages, two primary methods used for estimations of age of subadults (those under the age of 18) are epiphyseal union and dental development. Epiphyseal union (or epiphyseal fusion) refers to the appearance and closure of the epiphyseal plates between the primary centers of growth in a bone and the subsequent centers of growth (see Figure 15.7). Prior to complete union, the cartilaginous area between the primary and secondary centers of growth is also referred to as the growth plates (Schaefer, Black, and Scheuer 2009). Different areas of the skeleton have documented differences in the appearance and closure of epiphyses, making this a reliable method for aging subadult remains (SWGANTH 2013).
As an example of its utility in the identification process, epiphyseal development was used to identify two subadult victims of a fatal fire in Flint, Michigan, in February 2010. The remains represented two young girls, ages three and four. Due to the intensity of the fire, the subadult victims were differentiated from each other through the appearance of the patella, the kneecap. The patella is a bone that develops within the tendon of the quadriceps muscle at the knee joint. The patella begins to form around three to four years of age (Cunningham, Scheuer, and Black 2016, 407–409). In the example above, radiographs of the knees showed the presence of a patella in the four-year-old girl and the absence of a clearly discernible patella in the three-year-old.
Dental development begins during fetal stages of growth and continues until the complete formation and eruption of the adult third molars (if present). The first set of teeth to appear are called deciduous or baby teeth. Individuals develop a total of 20 deciduous teeth, including incisors, canines, and molars. These are generally replaced by adult dentition as an individual grows (Figure 15.12). A total of 32 teeth are represented in the adult dental arcade, including incisors, canines, premolars, and molars. When dental development is used for age estimations, researchers use both tooth-formation patterns and eruption schedules as determining evidence. For example, the crown of the tooth forms first followed by the formation of the tooth root. During development, an individual can exhibit a partially formed crown or a complete crown with a partially formed root. The teeth generally begin the eruption process once the crown of the tooth is complete. The developmental stages of dentition are one of the most reliable and consistent aging methods for subadults (Langley, Gooding, and Tersigni-Tarrant 2017, 176–177).
Degenerative changes in the skeleton typically begin after 18 years of age, with more prominent changes developing after an individual reaches middle adulthood (commonly defined as after 35 years of age in osteology). These changes are most easily seen around joint surfaces of the pelvis, the cranial vault, and the ribs. In this chapter, we focus on the pubic symphysis surfaces of the pelvis and the sternal ends of the ribs, which show metamorphic changes from young adulthood to older adulthood. The pubic symphysis is a joint that unites the left and right halves of the pelvis. The surface of the pubic symphysis changes during adulthood, beginning as a surface with pronounced ridges (called billowing) and flattening with a more distinct rim to the pubic symphysis as an individual ages. As with all metamorphic age changes, older adults tend to develop lipping around the joint surfaces as well as a breakdown of the joint surfaces. The most commonly used method for aging adult skeletons from the pubic symphysis is the Suchey-Brooks method (Brooks and Suchey 1990; Katz and Suchey 1986). This method divides the changes seen with the pubic symphysis into six phases based on macroscopic age-related changes to the surface. Figure 15.13 provides a visual of the degenerative changes that typically occur on the pubic symphysis.
The sternal end of the ribs, the anterior end of the rib that connects via cartilage to the sternum, is also used in age estimations of adults. This method, first developed by M. Y. İşcan and colleagues, considers both the change in shape of the sternal end as well as the quality of the bone (İşcan, Loth, and Wright 1984; İşcan, Loth, and Wright 1985). The sternal end first develops a billowing appearance in young adulthood. The bone typically develops a wider and deeper cupped end as an individual ages. Older adults tend to exhibit bony extensions of the sternal end rim as attaching cartilage ossifies. Figure 15.14 provides a visual of the degenerative changes that typically occur in sternal rib ends.
Estimating Stature
Stature, or height, is one of the most prominently recorded components of the biological profile. Our height is recorded from infancy through adulthood. Doctor’s appointments, driver's license applications, and sports rosters all typically involve a measure of stature for an individual. As such, it is also a component of the biological profile nearly every individual will have on record. Bioarchaeologists and forensic anthropologists use stature estimation methods to provide a range within which an individual’s biological height would fall. Biological height is a person’s true anatomical height. However, the range created through these estimations is often compared to reported stature, which is typically self-reported and based on an approximation of an individual’s true height (Ousley 1995).
In June 2015, two men were shot and killed in Granite Bay, California, in a double homicide. Investigators were able to locate surveillance camera footage from a gas station where the two victims were spotted in a car with another individual believed to be the perpetrator in the case. The suspect, sitting behind the victims in the car, hung his right arm out of the window as the car drove away. The search for the perpetrator was eventually narrowed down to two suspects. One suspect was 5’ 8” while the other suspect was 6’ 4”, representing almost a foot difference in height reported stature between the two. Forensic anthropologists were given the dimensions of the car (for proportionality of the arm) and were asked to calculate the stature of the suspect in the car from measurements of the suspect’s forearm hanging from the window. Approximate lengths of the bones of the forearm were established from the video footage and used to create a predicted stature range. Stature estimations from skeletal remains typically look at the correlation between the measurements of any individual bone and the overall measurement of body height. In the case above, the length of the right forearm pointed to the taller of the two suspects who was subsequently arrested for the homicide.
Certain bones, such as the long bones of the leg, contribute more to our overall height than others and can be used with mathematical equations known as regression equations. Regression methods examine the relationship between variables such as height and bone length and use the correlation between the variables to create a prediction interval (or range) for estimated stature. This method for calculating stature is the most commonly used method (SWGANTH 2012). Figure 15.15 shows the measurement of the bicondylar length of the femur for stature estimations.
Identification Using Individualizing Characteristics
One of the most frequently requested analyses within the forensic anthropology laboratory is assistance with the identification of unidentified remains. While all components of a biological profile, as discussed above, can assist law enforcement officers and medical examiners to narrow down the list of potential identifications, a biological profile will not lead to a positive identification. The term positive identification refers to a scientifically validated method of identifying previously unidentified remains. Presumptive identifications, however, are not scientifically validated; rather, they are based on circumstances or scene context. For example, if a decedent is found in a locked home with no evidence of forced entry but the body is no longer visually identifiable, it may be presumed that the remains belong to the homeowner. Hence, a presumptive identification.
The medicolegal system ultimately requires that a positive identification be made in such circumstances, and a presumptive identification is often a good way to narrow down the pool of possibilities. Biological profile information also assists with making a presumptive identification based on an individual’s phenotype in life (e.g., what they looked like). As an example, a forensic anthropologist may establish the following components of a biological profile: white male, between the ages of 35 and 50, approximately 5’ 7” to 5’ 11.” While this seems like a rather specific description of an individual, you can imagine that this description fits dozens, if not hundreds, of people in an urban area. Therefore, law enforcement can use the biological profile information to narrow their pool of possible identifications to include only white males who fit the age and height outlined above. Once a possible match is found, the decedent can be identified using a method of positive identification.
Positive identifications are based on what we refer to as individualizing traits or characteristics, which are traits that are unique at the individual level. For example, brown hair is not an individualizing trait as brown is the most common hair color in the U.S. But, a specific pattern of dental restorations or surgical implants can be individualizing, because it is unlikely that you will have an exact match on either of these traits when comparing two individuals.
A number of positive methods are available to forensic anthropologists, and for the remainder of this section we will discuss the following methods: comparative medical and dental radiography and identification of surgical implants.
Comparative medical and dental radiography is used to find consistency of traits when comparing antemortem records (medical and dental records taken during life) with images taken postmortem (after death). Comparative medical radiography focuses primarily on features associated with the skeletal system, including trabecular pattern (internal structure of bone that is honeycomb in appearance), bone shape or cortical density (compact outer layer of bone), and evidence of past trauma, skeletal pathology, or skeletal anomalies. Other individualizing traits include the shape of various bones or their features, such as the frontal sinuses (Figure 15.16).
Comparative dental radiography focuses on the number, shape, location, and orientation of dentition and dental restorations in antemortem and postmortem images. While there is not a minimum number of matching traits that need to be identified for an identification to be made, the antemortem and postmortem records should have enough skeletal or dental consistencies to conclude that the records did in fact come from the same individual (SWGANTH 2010a). Consideration should also be given to population-level frequencies of specific skeletal and dental traits. If a trait is particularly common within a given population, it may not be a good trait to utilize for positive identification.
Surgical implants or devices can also be used for identification purposes (Figure 15.17). These implements are sometimes recovered with human remains. One of the ways forensic anthropologists can use surgical implants to assist in decedent identification is by providing a thorough analysis of the implant and noting any identifying information such as serial numbers, manufacturer symbols, and so forth. This information can then sometimes be tracked directly to the manufacturer or the place of surgical intervention, which may be used to identify unknown remains (SWGANTH 2010a).
Special Topic: Trans Doe Task Force
The Trans Doe Task Force (TDTF) is a Trans-led nonprofit organization that investigates cases involving LGBTQ+ missing and murdered persons. The organization specifically focuses on transgender and gender-variant cases, providing connections between law enforcement agencies, medical examiner offices, forensic anthropologists, and forensic genetic genealogists to increase the chances of identification. Additionally, the TDTF curates a data repository of missing, murdered, and unclaimed LGBTQ+ individuals, and they continuously try innovative approaches to identify these individuals, whose lived gender identity may not match their biological sex.
For more information visit transdoetaskforce.org
Trauma Analysis
Types of Trauma
Within the field of anthropology, trauma is defined as an injury to living tissue caused by an extrinsic force or mechanism (Lovell 1997:139). Forensic anthropologists can assist a forensic pathologist by providing an interpretation of the course of events that led to skeletal trauma. Typically, traumatic injury to bone is classified into one of four categories, defined by the trauma mechanism. A trauma mechanism refers to the force that produced the skeletal modification and can be classified as (1) sharp force, (2) blunt force, (3) projectile, or (4) thermal (burning). Each type of trauma, and the characteristic pattern(s) associated with that particular categorization, will be discussed below.
First, let’s consider sharp-force trauma, which is caused by a tool that is edged, pointed, or beveled—for example, a knife, saw, or machete (SWGANTH 2011). The patterns of injury resulting from sharp-force trauma include linear incisions created by a sharp, straight edge; punctures; and chop marks (Figure 15.18; SWGANTH 2011). When observed under a microscope, an anthropologist can often determine what kind of tool created the bone trauma. For example, a power saw cut will be discernible from a manual saw cut.
Second, blunt-force trauma is defined as “a relatively low-velocity impact over a relatively large surface area” (Galloway 1999, 5). Blunt-force injuries can result from impacts from clubs, sticks, fists, and so forth. Blunt-force impacts typically leave an injury at the point of impact but can also lead to bending and deformation in other regions of the bone. Depressions, fractures, and deformation at and around the site of impact are all characteristics of blunt-force trauma (Figure 15.19). As with sharp-force trauma, an anthropologist attempts to interpret blunt-force injuries, providing information pertaining to the type of tool used, the direction of impact, the sequence of impacts, if more than one, and the amount of force applied.
Third, projectile trauma refers to high-velocity trauma, typically affecting a small surface area (Galloway 1999, 6). Projectile trauma results from fast-moving objects such as bullets or shrapnel. It is typically characterized by penetrating defects or embedded materials (Figure 15.20). When interpreting injuries resulting from projectile trauma, an anthropologist can often offer information pertaining to the type of weapon used (e.g., rifle vs. handgun), relative size of the bullet (but not the caliber of the bullet), the direction the projectile was traveling, and the sequence of injuries if there are multiple present.
Finally, thermal trauma is a bone alteration that results from bone exposure to extreme heat. Thermal trauma can result in cases of house or car fires, intentional disposal of a body in cases of homicidal violence, plane crashes, and so on. Thermal trauma is most often characterized by color changes to bone, ranging from yellow to black (charred) or white (calcined). Other bone alterations characteristic of thermal trauma include delamination (flaking or layering due to bone failure), shrinkage, fractures, and heat-specific burn patterning. When interpreting injuries resulting from thermal damage, an anthropologist can differentiate between thermal fractures and fractures that occurred before heat exposure, thereby contributing to the interpretation of burn patterning (e.g., was the individual bound or in a flexed position prior to the fire?).
While there are characteristic patterns associated with the four categories of bone trauma, it is also important to note that these bone alterations do not always occur independently of different trauma types. An individual’s skeleton may present with multiple different types of trauma, such as a projectile wound and thermal trauma. Therefore, it is important that the anthropologist recognize the different types of trauma and interpret them appropriately.
Timing of Injury
Another important component of any anthropological trauma analysis is the determination of the timing of injury (e.g., when did the injury occur). Timing of injury is traditionally split into one of three categories: antemortem (before death), perimortem (at or around the time of death), and postmortem (after death). This classification system differs slightly from the classification system used by the pathologist because it specifically references the qualities of bone tissue and bone response to external forces. Therefore, the perimortem interval (at or around the time of death) means that the bone is still fresh and has what is referred to as a green bone response, which can extend past death by several weeks or even months. For example, in cold or freezing temperatures a body can be preserved for extended periods of time, increasing the perimortem interval, while in desert climates decomposition is accelerated, thereby significantly decreasing the postmortem interval (Galloway 1999, 12). Antemortem injuries (occurring well before death and not related to the death incident) are typically characterized by some level of healing, in the form of a fracture callus or unification of fracture margins. Finally, postmortem injuries (occurring after death, while bone is no longer fresh) are characterized by jagged fracture margins, resulting from a loss of moisture content during the decomposition process (Galloway 1999, 16). In general, all bone traumas should be classified according to the timing of injury, if possible. This information will help the medical examiner or pathologist better understand the circumstances surrounding the decedent’s death, as well as events occurring during life and after the final disposition of the body.
The Role of the Forensic Anthropologist in Trauma Analysis
Within the medicolegal system, forensic anthropologists are often called upon by the medical examiner, forensic pathologist, or coroner to assist with an interpretation of trauma. The forensic anthropologist’s main focus in any trauma analysis is the underlying skeletal system—as well as, sometimes, cartilage. Analysis and interpretation of soft tissue injuries fall within the purview of the medical examiner or pathologist. It is also important to note that the main role of the forensic anthropologist is to provide information pertaining to skeletal injury to assist the medical examiner/pathologist in their final interpretation of injury. Forensic anthropologists do not hypothesize as to the cause of death of an individual. Instead, a forensic anthropologist’s report should include a description of the injury (e.g., trauma mechanism, number of injuries, location, timing of injury); documentation of the injury, which may be utilized in court testimony (e.g., photographs, radiographs, measurements); and, if applicable, a statement as to the condition of the body and state of decomposition, which may be useful for understanding the depositional context (e.g., how long has the body been exposed to the elements; was it moved or in its original location; are any of the alterations to bone due to environmental or faunal exposure instead of intentional human modification).
Taphonomy
What Happened to the Remains After Death?
The majority of the skeletal analysis process revolves around the identity of the deceased individual. However, there is one last, very important question that forensic anthropologists should ask: What happened to the remains after death? Generally speaking, processes that alter the bone after death are referred to as taphonomic changes (refer to Chapter 7 for a discussion regarding taphonomy and the fossil record).
The term taphonomy was originally used to refer to the processes through which organic remains mineralize, also known as fossilization. Within the context of biological anthropology, the term taphonomy is better defined as the study of what happens to human remains after death (Komar and Buikstra 2008). Initial factors affecting a body after death include processes such as decomposition and scavenging by animals. However, taphonomic processes encompass much more than the initial period after death. For example, plant root growth can leach minerals from bone, leaving a distinctive mark. Sunlight can bleach human remains, leaving exposed areas whiter than those that remained buried. Water can wear the surface of the bone until it becomes smooth.
Some taphonomic processes can help a forensic anthropologist estimate the relative amount of time that human remains have been exposed to the elements. For example, root growth through a bone would certainly indicate a body was buried for more than a few days. Forensic anthropologists must be very careful when attempting to estimate time since death based on taphonomic processes because environmental conditions can greatly influence the rate at which taphonomic processes progress. For example, in cold environments, tissue may decay slower than in warm, moist environments.
Forensic anthropologists must contend with taphonomic processes that affect the preservation of bones. For example, high acidity in the soil can break down human bone to the point of crumbling. In addition, when noting trauma, they must be very careful not to confuse postmortem (after death) bone damage with trauma.
Taphonomic Process |
Definition |
Rodent Gnawing |
When rodents, such as rats and mice, chew on bone, they leave sets of parallel grooves. The shallow grooves are etched by the rodent’s incisors. |
Carnivore Damage
|
Carnivores may leave destructive dental marks on bone. The tooth marks may be visible as pit marks or punctures from the canines, as well as extensive gnawing or chewing of the ends of the bones to retrieve marrow. |
Burned Bone |
Fire causes observable damage to bone. Temperature and the amount of time bone is heated affect the appearance of the bone. Very high temperatures can crack bone and result in white coloration. Color gradients are visible in between high and lower temperatures, with lower temperatures resulting in black coloration from charring. Cracking can also reveal information about the directionality of the burn. |
Root Etching |
Plant roots can etch the outer surface of bone, leaving grooves where the roots attached as they leached nutrients. During this process, the plant’s roots secrete acid that breaks down the surface of the bone.
|
Weathering
|
Many different environmental conditions affect bone. River transport can smooth the surface of the bone due to water abrasion. Sunlight can bleach the exposed surface of bone. Dry and wet environments or the mixture of both types of environments can cause cracking and exfoliation of the surface. Burial in different types of soil can cause discoloration, and exposure can cause degreasing. |
Cut Marks |
Humans may alter bone by cutting, scraping, or sawing it directly or in the process of removing tissue. The groove pattern—that is, the depth and width of the cuts—can help identify the tool used in the cutting process. |
Ethics and Human Rights
Working with human remains requires a great deal of consideration and respect for the dead. Forensic anthropologists have to think about the ethics of our use of human remains for scientific purposes. How do we conduct casework in the most respectable manner possible? While there are a wide range of ethical considerations to consider when contemplating a career in forensic anthropology, this chapter will focus on two major categories: working with human remains and acting as an expert within the medicolegal system.
Working with Human Remains
Forensic anthropologists work with human remains in a number of contexts, including casework, excavation, research, and teaching. When working with human remains, it is always important to use proper handling techniques. To prevent damage to skeletal remains, bones should be handled over padded surfaces. Skulls should never be picked up by placing fingers in the eye orbits, foramen magnum (hole at the base of the skull for entry of the spinal cord), or through the zygomatic arches (cheekbones). Human remains, whether related to casework, fieldwork, donated skeletal collections, or research, were once living human beings. It is important to always bear in mind that work with remains should be ingrained with respect for the individual and their relatives. In addition to fieldwork, casework, and teaching, anthropologists are often invited to work with remains that come from a bioarchaeological context or from a human rights violation. While this discussion of ethics is not comprehensive, two case examples will be provided below in which an anthropologist must consider the ethical standards outlined above.
Modern Human Rights Violations
Forensic anthropologists may also be called to participate in criminal investigations involving human rights violations. Anthropological investigations may include assistance with identifications, determination of the number of victims, and trauma analyses. In this role, forensic anthropologists play an integral part in promoting human rights, preventing future human rights violations, and providing the evidence necessary to prosecute those responsible for past events. A few ethical considerations for the forensic anthropologist involved in human rights violations include the use of appropriate standards of identification, presenting reliable and unbiased testimony, and maintaining preservation of evidence. For a more comprehensive history of forensic anthropological contributions to human rights violations investigations, see Ubelaker 2018.
Acting as an Expert in the Medicolegal System
In addition to the ethical considerations involved in working with human skeletal remains, forensic anthropologists must abide by ethical standards when they act as experts within the medicolegal system. The role of the forensic anthropologist within the medicolegal system is primarily to provide information to the medical examiner or coroner that will aid in the identification process or determination of cause and manner of death. Forensic anthropologists also may be called to testify in a court of law. In this capacity, forensic anthropologists should always abide by a series of ethical guidelines that pertain to their interpretation, presentation, and preservation of evidence used in criminal investigations. First and foremost, practitioners should never misrepresent their training or education. When appropriate, outside opinions and assistance in casework should be requested (e.g., consulting a radiologist for radiological examinations or odontologist for dental exams). The best interest of the decedent should always take precedence. All casework should be conducted in an unbiased way, and financial compensation should never be accepted as it can act as an incentive to take a biased stance regarding casework. All anthropological findings should be kept confidential, and release of information is best done by the medical examiner or coroner. Finally, while upholding personal ethical standards, forensic anthropologists are also expected to report any perceived ethical violations committed by their peers.
Ethical standards for the field of forensic anthropology are outlined by the Organization of Scientific Area Committees (OSAC) for Forensic Science, administered by the National Institute of Standards and Technology (NIST). OSAC and NIST recently began an initiative to develop standards that would strengthen the practice of forensic science both in the United States and internationally. OSAC’s main objective is to “strengthen the nation’s use of forensic science by facilitating the development of technically sound forensic science standards and by promoting the adoption of those standards by the forensic science community” (NIST n.d.). Additionally, OSAC promotes the establishment of best practices and other guidelines to ensure that forensic science findings and their presentation are reliable and reproducible (NIST 2023).
Special Topic: Native American Graves Protection and Repatriation Act (NAGPRA)
There is a long history in the United States of systematic disenfranchisement of Native American people, including lack of respect for tribal sovereignty. This includes the egregious treatment of Native American human remains. Over several centuries, thousands of Native American remains were removed from tribal lands and held at institutions in the United States, such as museums and universities.
In 1990, a landmark human rights federal law, the Native American Graves Protection and Repatriation Act (NAGPRA), spurred change in the professional standards and practice of biological anthropology and archaeology. NAGPRA established a legal avenue to provide protection for and repatriation of Native American remains, cultural items, and sacred objects removed from Federal or tribal lands to Native American lineal descendants, Indian tribes, and Native Hawaiian organizations. Human remains and associated artifacts, curated in museum collections and federally funded institutions, are subject to three primary provisions outlined by the NAGPRA statute: (1) protection for Native graves on federal and private land; (2) recognition of tribal authority on such lands; and (3) the requirement that all Native skeletal remains and associated artifacts be inventoried and culturally affiliated groups be consulted concerning decisions related to ownership and final disposition (Rose, Green, and Green 1996). NAGPRA legislation was enacted to ensure ethical consideration and treatment of Native remains and to improve dialogue between scientists and Native groups.
- For more information about NAGPRA, visit the Bureau of Reclamation NAGPRA website
- To read the text of the law, visit the US Congress NAGPRA law website.
- For further discussion of NAGPRA history, please see TRACES: An Open Invitation to Archaeology open textbook website
Becoming a Forensic Anthropologist
What does it take to be a forensic anthropologist? Forensic anthropologists are first and foremost anthropologists. While many forensic anthropologists have an undergraduate degree in anthropology, they may also major in biology, criminal justice, pre-law, pre-med, and many other related fields. Practicing forensic anthropologists typically have an advanced degree, either a Master’s or Doctoral degree in Anthropology. Additional training and experience in archaeology, the medico-legal system, rules of evidence, and expert witness testimony are also common. Practicing forensic anthropologists are also encouraged to be board-certified through the American Board of Forensic Anthropology (ABFA). Learn more about the field and educational opportunities on the ABFA website: https://www.theabfa.org/coursework.
Review Questions
- What is forensic anthropology? What are the seven primary steps involved in a skeletal analysis?
- What are the major components of a biological profile? Why are forensic anthropologists often-tasked with creating biological profiles for unknown individuals?
- What are the four major types of skeletal trauma?
- What is taphonomy, and why is an understanding of taphonomy often critical in forensic anthropology analyses?
- What are some of the ethical considerations faced by forensic anthropologists?
About the Authors
Ashley Kendell, Ph.D.
California State University, Chico, akendell@csuchico.edu
Dr. Ashley Kendell is currently an associate professor and forensic anthropologist at Chico State. Prior to beginning her position at Chico State, she was a visiting professor at the University of Montana and the forensic anthropologist for the state of Montana. Dr. Kendell obtained her doctorate from Michigan State University, and her research interests include skeletal trauma analysis and digitization and curation methods for digital osteological data. She is also a Registry Diplomate of the American Board of Medicolegal Death Investigators. Throughout her doctoral program, she worked as a medicolegal death investigator for the greater Lansing, Michigan, area and was involved in the investigation of over 200 forensic cases.
Alex Perrone, M.A., M.S.N, R.N., P.H.N.
Butte Community College, perroneal@butte.edu
Alex Perrone is a lecturer in anthropology at Butte Community College. She is also a Registered Nurse and a certified Public Health Nurse. She is a former Supervisor of the Human Identification Laboratory in the Department of Anthropology at California State University, Chico. Her research interests include bioarchaeology, paleopathology, forensic anthropology, skeletal biology, California prehistory, and public health. She has worked on bioarchaeological and archaeological projects in Antigua, California, Hawaii, Greece, and the UK, and was an archaeological technician for the USDA Forest Service. She assisted with training courses for local and federal law enforcement agencies and assisted law enforcement agencies with the recovery and analysis of human remains.
Colleen Milligan, Ph.D.
California State University, Chico, cfmilligan@csuchico.edu
Dr. Colleen Milligan is a biological and forensic anthropologist with research interests in bioarchaeology, skeletal biology, and forensic anthropology. She has been a Fellow with the Department of Homeland Security and has assisted in forensic anthropology casework and recoveries in the State of Michigan and California. She has also assisted in community outreach programs in forensic anthropology and forensic science, as well as recovery training courses for local, state, and federal law enforcement officers. She is a certified instructor through Peace Officers Standards and Training (POST). Dr. Milligan serves as the current co-director of the Chico State Human Identification Laboratory.
For Further Exploration
The American Board of Forensic Anthropology (ABFA)
The American Academy of Forensic Sciences (AAFS)
The Organization of Scientific Area Committees for Forensic Science (OSAC)
References
Adams, Bradley J., and Lyle W. Konigsberg, eds. 2008. Recovery, Analysis, and Identification of Commingled Remains. Totowa, NJ: Humana Press.
Beatrice, Jared S., and Angela Soler. 2016. “Skeletal Indicators of Stress: A Component of the Biocultural Profile of Undocumented Migrants in Southern Arizona.” Journal of Forensic Sciences 61 (5): 1164–1172.
Berg, Gregory E. 2017. “Sex Estimation of Unknown Human Skeletal Remains.” In Forensic Anthropology: A Comprehensive Introduction, Second Edition, edited by Natalie R. Langley and MariaTeresa A. Tersigni-Tarrant, 143–159. Boca Raton, FL: CRC Press.
Bethard, Jonathan D., and Elizabeth A. DiGangi. 2020. “Letter to the Editor—Moving Beyond a Lost Cause: Forensic Anthropology and Ancestry Estimates in the United States.” Journal of Forensic Sciences 65 (5): 1791–1792.
Birkby, Walter H., Todd W. Fenton, and Bruce E. Anderson. 2008. “Identifying Southwest Hispanics Using Nonmetric Traits and the Cultural Profile.” Journal of Forensic Sciences 53 (1): 29–33.
Blatt, Samantha, Amy Michael, and Lisa Bright. Forthcoming. “Bioarchaeology: Interpreting Human Behavior from Skeletal Remains.” In TRACES: An Open Invitation to Archaeology. https://textbooks.whatcom.edu/tracesarchaeology/.
Brooks, S., and J. M. Suchey. 1990. “Skeletal Age Determination Based on the Os Pubis: A Comparison of the Acsádi-Nemeskéri and Suchey-Brooks Methods.” Human Evolution 5 (3): 227–238.
Buchanan, Shelby. 2014. “Bone Modification in Male to Female Transgender Surgeries: Considerations for the Forensic Anthropologist.” MA thesis, Department of Geography and Anthropology, Louisiana State University, Baton Rouge.
Cunningham, Craig, Louise Scheuer, and Sue Black. 2016. Developmental Juvenile Osteology, Second Edition. London: Elsevier Academic Press.
Galloway, Alison, ed. 1999. Broken Bones: Anthropological Analysis of Blunt Force Trauma. Springfield, IL: Charles C. Thomas Publisher, LTD.
Hefner, Joseph T., and Kandus C. Linde. 2018. Atlas of Human Cranial Macromorphoscopic Traits. San Diego: Academic Press.
İşcan, M. Y., S. R. Loth, and R. K. Wright. 1984. “Age Estimation from the Rib by Phase Analysis: White Males.” Journal of Forensic Sciences 29 (4): 1094–1104.
İşcan, M. Y., S. R. Loth, and R. K. Wright. 1985. “Age Estimation from the Rib by Phase Analysis: White Females.” Journal of Forensic Sciences 30 (3): 853–863.Katz, Darryl, and Judy Myers Suchey. 1986. “Age Determination of the Male Os Pubis.” American Journal of Physical Anthropology 69 (4): 427–435.
Komar, Debra A., and Jane E. Buikstra. 2008. Forensic Anthropology: Contemporary Theory and Practice. New York: Oxford University Press.
Langley, Natalie R., Alice F. Gooding, and MariaTeresa Tersigni-Tarrant. 2017. “Age Estimation Methods.” In Forensic Anthropology: A Comprehensive Introduction, Second Edition, edited by Natalie R. Langley and MariaTeresa A. Tersigni-Tarrant, 175–191. Boca Raton, FL: CRC Press.
Lovell, Nancy C. 1997. “Trauma Analysis in Paleopathology.” Yearbook of Physical Anthropology 104 (S25): 139–170.
Native American Graves Protection and Repatriation Act (NAGPRA) 1990 (25 U.S. Code 3001 et seq.)
NIST (National Institute of Standards and Technology). N.d. “The Organization of Scientific Area Committees for Forensic Science.” Accessed April 18, 2023. https://www.nist.gov/topics/organization-scientific-area-committees-forensic-science.
Ousley, Stephen. 1995. “Should We Estimate Biological or Forensic Stature?” Journal of Forensic Sciences 40(5): 768–773.
Phenice, T. W. 1969. “A Newly Developed Visual Method of Sexing the Os Pubis.” American Journal of Physical Anthropology 30 (2): 297–302.
Rose, Jerome C., Thomas J. Green, and Victoria D. Green. 1996. “NAGPRA Is Forever: Osteology and the Repatriation of Skeletons.” Annual Review of Anthropology 25: 81–103.
Schaefer, Maureen, Sue Black, and Louise Scheuer. Juvenile Osteology: A Laboratory and Field Manual. 2009. San Diego: Elsevier Academic Press.
Schall, Jenna L., Tracy L. Rogers, and Jordan D. Deschamps-Braly. 2020. “Breaking the Binary: The Identification of Trans-women in Forensic Anthropology.” Forensic Science International 309: 110220. https://doi.org/10.1016/j.forsciint.2020.110220.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2010a. “Personal Identification.” Last modified June 3, 2010. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_personal_identification.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2010b. “Sex Assessment.” Last modified June 3, 2010. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_sex_assessment.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2011. “Trauma Analysis.” Last modified May 27, 2011. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_trauma.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2012. “Stature Estimation.” Last modified August 2, 2012. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_stature_estimation.pdf.
Scientific Working Group for Forensic Anthropology (SWGANTH). 2013. “Age Estimation.” Last modified January 22, 2013. https://www.nist.gov/sites/default/files/documents/2018/03/13/swganth_age_estimation.pdf.
Soler, Angela, and Jared S. Beatrice. 2018. “Expanding the Role of Forensic Anthropology in Humanitarian Crisis: An Example from the USA-Mexico Border. In Sociopolitics of Migrant Death and Repatriation: Perspectives from Forensic Science, edited by Krista E. Latham and Alyson J. O’Daniel, 115–128. New York: Springer.
Soler, Angela, Robin Reineke, Jared Beatrice, and Bruce E. Anderson. 2019. “Etched in Bone: Embodied Suffering in the Remains of Undocumented Migrants.” In The Border and Its Bodies: The Embodiment of Risk along the U.S.-México Line, edited by Thomas E. Sheridan and Randall H. McGuire, 173–207. Tucson: University of Arizona Press.
Stull, Kyra E., Eric J. Bartelink, Alexandra R. Klales, Gregory E. Berg, Michael W. Kenyhercz, Erica N. L’Abbé, Matthew C. Go, et al.. 2021. “Commentary on: Bethard JD, DiGangi EA. Letter to the Editor—Moving Beyond a Lost Cause: Forensic Anthropology and Ancestry Estimates in the United States. J Forensic Sci. 2020;65(5):1791–2. doi: 10.1111/1556-4029.14513.” Journal of Forensic Sciences 66 (1): 417–420.
Tallman, Sean D., Caroline D. Kincer, and Eric D. Plemons. 2022. “Centering Transgender Individuals in Forensic Anthropology and Expanding Binary Sex Estimation in Casework and Research.” Special issue, “Diversity and Inclusion,” Forensic Anthropology 5 (2): 161–180.
Tersigni-Tarrant, MariaTeresa A., and Natalie R. Langley. 2017. “Human Osteology.” In Forensic Anthropology: A Comprehensive Introduction, Second Edition, edited by Natalie R. Langley and MariaTeresa A. Tersigni-Tarrant, 81–109. Boca Raton, FL: CRC Press.
Ubelaker, Douglas H. 2018. “A History of Forensic Anthropology.” Special issue, “Centennial Anniversary Issue of AJPA,” American Journal of Physical Anthropology 165 (4): 915–923.
White, Tim D., and Pieter A. Folkens. 2005. The Human Bone Manual. Burlington, MA: Elsevier Academic Press.
Winburn, Allysha P., and Bridget Algee-Hewitt. 2021. “Evaluating Population Affinity Estimates in Forensic Anthropology: Insights from the Forensic Anthropology Database for Assessing Methods Accuracy (FADAMA).” Journal of Forensic Sciences 66 (4): 1210–1219.
Winburn, Allysha Powanda, Sarah Kiley Schoff, and Michael W. Warren. 2016. “Assemblages of the Dead: Interpreting the Biocultural and Taphonomic Signature of Afro- Cuban Palo Practice in Florida.” Journal of African Diaspora Archaeology and Heritage 5 (1): 1–37.
Joylin Namie, Ph.D., Truckee Meadows Community College
This chapter is a revision from "Chapter 16: Contemporary Topics: Human Biology and Health” by Joylin Namie. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Learning Objectives
- Describe the major transitions in patterns of disease that have occurred throughout human evolution.
- Describe what is meant by a “mismatch” between our evolved biology and contemporary lifestyles and how this is reflected in modern disease patterns.
- Explain how the human stress response can positively and negatively have an impact on health.
- Explain what a “syndemic” is and why the COVID-19 pandemic represents one.
- Describe the ways institutionalized racism and bias in the medical field contributed to different rates of exposure, differential treatment, morbidity, and mortality from COVID-19 for different ethnic groups in the United States.
When was the last time you needed to do research for an upcoming paper? I bet you started by looking for information online. How did you go about your search? Which websites looked promising? Which ones did not entice you to click past the home page? Once you found one you thought might be useful, how much time did you spend searching for information? At what point did you decide to leave that site and move on? I would wager money that you never once thought your behavior had anything to do with human evolution, but it does.
Although we may not often stop to think about it, our evolutionary past is reflected in many aspects of modern life. The ways we “forage” for information on the internet mimics the ways we once foraged for food during our several-million-year history as hunter-gatherers (Chin et al. 2015). Humans are visual hunters (Lieberman 2006). We practice optimal foraging strategy, meaning we make decisions based on energy return for investment (McElroy and Townsend 2009). When we search for information online, we locate a “patch,” in this case a website or research article, then quickly scan the contents to discern how much of it is useful to us. Like our hominin ancestors, we spend more time in “patches” with abundant resources and abandon sites quickly once we have exhausted the available goods. As with internet searches, our evolutionary past is also reflected in the kinds of landscapes we find appealing, the foods that taste good to us, why we break a sweat at the gym, and why we have to go to the gym at all (Bogin 1991; Dutton 2009; Lieberman 2015). Many of the health problems facing humans in the 21st century also have their beginnings in the millions of years we roamed the earth as foragers.
This chapter addresses contemporary health issues from an evolutionary perspective. It begins with a review of diet, activity patterns, and causes of morbidity (sickness) and mortality (death) among our preagricultural ancestors, which form the foundation for the ways our bodies function today. This is followed by a brief review of the health consequences of the transition to agriculture (see “Cultural Effects of Agriculture” in Chapter 12 for more detail), marking the first of several major epidemiological transitions (changes in disease patterns) experienced by humankind. It then addresses health conditions affecting modern, industrialized societies, including obesity, diabetes, heart disease, cancer, and the effect of stress on health. The environments in which we now live and the choices we make put a strain on biological systems that came about in response to selective pressures in our past, meaning the ways our bodies evolved are, in many ways, a mismatch for the conditions of modern life.
Preagricultural Humans
Diet
Humans may be the species with the longest list of nutritional requirements (Bogin 1991). This is due to the fact that we evolved in environments where there was a high diversity of edible species but low densities of any given species. Homo sapiens sapiens require 45–50 essential nutrients for growth, maintenance, and repair of cells and tissues. These include protein, carbohydrates, fats, vitamins, minerals, and water. As a species, we are (or were) physically active with high metabolic demands. We are also omnivorous and evolved to choose foods that are dense in essential nutrients. One of the ways we identified high-calorie resources in our evolutionary past was through taste, and it is no accident that humans find sweet, salty, fatty foods appealing.
The human predisposition toward sugar, salt, and fat is innate (Farb and Armelagos 1980). Receptors for sweetness are found in every one of our mouth’s 10,000 taste buds (Moss 2013). Preference for sweet makes sense in an ancestral environment where sweetness signaled high-value resources like ripe fruits. Likewise, “the long evolutionary path from sea-dwelling creatures to modern humans has given us salty body fluids, the exact salinity of which must be maintained” (Farb and Armelagos 1980), drawing us to salty-tasting things. Cravings for fat are also inborn, with some archaeological evidence suggesting that hominins collected animal bones for their fatty marrow, which contains two essential fatty acids necessary for brain development (Richards 2002), rather than for any meat remaining on the surface (Bogin 1991).
Bioarchaeological studies indicate Paleolithic peoples ate a wider variety of foods than many people eat today (Armelagos et al. 2005; Bogin 1991; Larsen 2014; Marciniak and Perry 2017). Foragers took in more protein, less fat, much more fiber, and far less sodium than modern humans typically do (Eaton, Konner, and Shostak 1988). Changes in tooth and intestinal morphology illustrate that animal products were an important part of human diets from the time of Homo erectus onward (Baltic and Boskovic 2015; Richards 2002; Wrangham 2009). Once cooking became established, it opened up a wider variety of both plant and animal resources to humans. However, the protein, carbohydrates, and fats preagricultural peoples ate were much different from those we eat today. Wild game lacked the antibiotics, growth hormones, and high levels of cholesterol and saturated fat associated with industrialized meat production today (Walker et al. 2005). Wild game was also protein dense, providing only 50% of energy as fat (Lucock et al. 2014). The ways meat is prepared and eaten today also have implications for disease.
Meats cooked well done over high heat and/or over an open flame, including hamburgers and barbecued meats, are highly carcinogenic due to compounds formed during the cooking process (Trafialek and Kolanowski 2014). Processed meats that have been preserved by smoking, curing, salting, or by adding chemical preservatives such as sodium nitrite (e.g., ham, bacon, pastrami, salami, and beef jerky) have been linked to cancers of the colon, lung, and prostate (Abid, Cross, and Sinha 2014; Figure 16.1). Nitrites/nitrates have additionally been linked to cancers of the ovaries, stomach, esophagus, bladder, pancreas, and thyroid (Abid, Cross, and Sinha 2014). In addition, studies analyzing the diets of 103,000 Americans for up to 16 years indicate that those who ate grilled, broiled, or roasted meats more than 15 times per month were 17% more likely to develop high blood pressure than those who ate meat fewer than four times per month, and participants who preferred their meats well done were 15% more likely to suffer from hypertension (high blood pressure) than those who ate their meats rare (Liu 2018). A previous study of the same cohort indicated “independent of consumption amount, open-flame and/or high-temperature cooking for both red meat and chicken is associated with an increased risk of type 2 diabetes among adults who consume animal flesh regularly” (Liu et al. 2018). Although meat has been argued to be crucial to cognitive and physical development among hominins (Wrangham 2009), there has been an evolutionary trade-off between the ability to preserve protein through cooking and the health risks of cooked meat and chemical preservatives.
Although carbohydrates represent half of the diet on average for both ancient foragers and modern humans, the types of carbohydrates consumed are very different. Ancient foragers ate fresh fruits, vegetables, grasses, legumes, and tubers, rather than the processed carbohydrates common in industrialized economies (Moss 2013). Their diets also lacked the refined white sugar and corn syrup found in many modern foods that contribute to the development of diabetes (Pontzer et al. 2012).
Physical Activity Patterns
How do we know how active our ancestors were? Hominin morphology and physiology provide us with clues. Adaptations to heat discussed in Chapter 14 evolved in response to the need for physical exertion in the heat of the day in equatorial Africa. Human adaptations for preventing hyperthermia (overheating) suggest an evolutionary history of regular, strenuous physical activity. Research with modern foraging populations also offers clues to ancient activity patterns. Although subject to sampling biases and marginal environments (Marlowe 2005), modern foragers provide the only direct observations of human behavior in the absence of agriculture (Lee 2013). From such studies, we know foragers cover greater distances in single-day foraging bouts than other living primates, and these treks require high levels of cardiovascular endurance (Raichlen and Alexander 2014). Recent research with the Hadza in Tanzania indicates they walk up to 11 kilometers (6.8 miles) daily while hunting and gathering (Pontzer et al. 2012), engaging in moderate-to-vigorous physical activity for over two hours each day—meeting the U.S. government’s weekly requirements for physical activity in just two days (Raichlen et al. 2016; Figure 16.2). The fact that humans were physically active in our evolutionary past is also supported by the fact that regular physical exercise has been shown to be protective against a variety of health conditions found in modern humans, including cardiovascular disease (Raichlen and Alexander 2014) and Alzheimer’s dementia (Mandsager, Harb, and Cremer 2018), even in the presence of brain pathologies indicative of cognitive decline (Buchman et al. 2019).
Infectious Disease
Population size and density remained low throughout the Paleolithic, by some estimates only 0.4 inhabitants per square kilometer (McClellan and Dorn 2006). This limited morbidity and mortality from infectious diseases, which sometimes require large populations to sustain epidemics. Our earliest ancestors had primarily two types of infections with which to contend (Armelagos 1990). The first were organisms that adapted to our prehominin ancestors and have been problems ever since. Examples include head lice, pinworms, and yaws. A second set of diseases were zoonoses, diseases that originate in animals and mutate into a form infectious to humans. One example is the Human Immunodeficiency Virus (HIV) that originated in nonhuman primates and was likely passed to humans through the butchering of hunted primates for food (Sharp and Hahn 2011). Zoonoses that could have infected ancient hunter-gatherers include tetanus and vector-borne diseases transmitted by flies, mosquitoes, fleas, midges, ticks, and the like. Many of these diseases are slow acting, chronic, or latent, meaning they can last for weeks, months, or even decades, causing low levels of sickness and allowing victims to infect others over long periods of time. Survival or cure does not result in lasting immunity, with survivors returning to the pool of potential victims. Such diseases often survive in animal reservoirs, reinfecting humans again and again (Wolfe et al. 2012). A study of bloodsucking insects preserved in samples of amber dating from 15 to 100 million years ago indicates that they carried microorganisms that today cause diseases such as river blindness, typhus, Lyme disease, and malaria (Poinar 2018). Such diseases may have been infecting humans throughout our history and may have had significant impacts on small foraging communities because they more often infected adults, who provided the food supply (Armelagos et al. 2005).
Health
Given their diets, levels of physical activity, and low population densities, nomadic preagricultural humans were likely in better health than many modern populations. This assertion is supported by comparative research conducted with modern foraging and industrialized populations. Measures of health taken from 20th-century foraging populations demonstrate excellent aerobic capacity, as measured by oxygen uptake during exertion, and low body-fat percentages, with triceps skinfold measurements half those of white Canadians and Americans. Serum cholesterol levels were also low, and markers for diabetes, hypertension, and cardiovascular disease were missing among them (Eaton, Konner, and Shostak 1988; Raichlen et al. 2016).
Health Consquences of the Transition to Agriculture and Animal Domestication
The shift from foraging to food production occurred relatively recently in our evolutionary history (Larsen 2014), and there are indications our biology has not yet caught up (Pritchard 2010). Beginning around 12,000 BCE in several parts of the globe, humans began to move toward a diet based on domesticated plants and animals (Armelagos et al. 2005). This involved manipulating the natural landscape to facilitate intensive food production, including the clearing of forest and construction of wells, irrigation canals, and ditches, exposing humans to water-borne illnesses and parasites and attracting mosquitos and other vectors of disease to human settlements. The heavy, repetitive physical labor of early agricultural production resulted in negative impacts on articular joints, including osteoarthritis (Larsen 2014). At the same time, nutritional diversity became restricted, focused on major cereal crops that continue to dominate agricultural production today, including corn, wheat, and rice (Jain 2012). This represented a major shift in diet from a wide variety of plant and animal foods to dependence on starchy carbohydrates, leading to increases in dental caries (cavities), reductions in stature and growth rates, and nutritional deficiencies (Larsen 2014). Domesticated animals added new foods to the human diet, including meat that was higher in fat and cholesterol than wild game as well as dairy products (Lucock et al. 2014). Agriculture provided the means to produce a storable surplus for the first time in human history, leading to the beginnings of economic inequality (see Chapter 12). Social hierarchies led to the unequal distribution of resources, concentrating infectious disease among the poor and malnourished (Zuckerman et al. 2014), a situation that continues to plague humanity today (Marmot 2005).
Sedentism and a rise in population density accompanied the move to agriculture, increasing the risk of infectious disease. Agriculture often provided enough calories, if not enough nutrition, to increase fertility. Although diets were worse and people unhealthier, populations continued to grow, even in the midst of high levels of child and maternal mortality and short life expectancies (Omran 2005). Hygiene became an issue as large settlements increased the problem of removing human waste and providing uncontaminated water (Armelagos et al. 2005). Domesticated animals provided reservoirs of zoonotic pathogens, which affected farmers more than foragers, as farmers were in closer proximity to their animals on a daily basis (Marciniak and Perry 2017). Many of these diseases became major killers of humankind, including influenza, tuberculosis, malaria, plague, syphilis, and smallpox, functioning as selective pressures in and of themselves (Cooling 2015). As these diseases encountered large human populations, they caused major epidemics that traveled along newly established routes for trade, warfare, and colonization.
Epidemiological Transitions
Changes in diet and physical-activity patterns, population densities, and exposure to zoonoses associated with agriculture resulted in an epidemiological transition, a shift in the causes of morbidity (sickness) and mortality (death) among humankind (Omran 2005). The first epidemiological transition from foraging to food production resulted in increases in dental caries (see Chapter 12), nutritional deficiencies, infectious disease, and skeletal conditions like osteoarthritis, as well as decreases in growth and height (Larsen 2014). A second epidemiological transition occurred following the Industrial Revolution in Western Europe and the United States when improved standards of living, hygiene, and nutrition minimized the effects of infectious disease, after which people began to experience higher rates of noncommunicable diseases, such as cancer, heart disease, and diabetes due to the changes in lifestyle, diet, and activity levels that are the subject of this chapter (Omran 2005). With the addition of immunizations and other public health initiatives, modified forms of this transition remain ongoing in many low- and middle-income countries (Zuckerman et al. 2014), with several now facing a “double burden” of disease, with poor, often rural, populations struggling with infectious diseases due to malnutrition, lack of sanitation, and access to health care, while more affluent citizens are victims of chronic illnesses. A third epidemiological transition is now underway as infectious diseases, including new, re-emergent, and multidrug-resistant diseases, have once again become major health concerns (Harper and Armelagos 2010; Zuckerman et al. 2014). These include COVID-19, Ebola, HIV/AIDS, tuberculosis, malaria, dengue, Lyme disease, and West Nile virus—all zoonoses that initially spread to humans through contact with animals.
Patterns of morbidity and mortality continue to shift across the globe. As with the first epidemiological transition resulting from the adoption of large-scale agriculture, such shifts can be the direct, if unintended, result of human interactions with the environment. For example, there has been a rise in chronic inflammatory diseases (CIDs) in developed countries (Versini et al. 2015). This includes increased rates of allergic conditions like asthma and autoimmune diseases like rheumatoid arthritis, multiple sclerosis, Crohn’s disease, and inflammatory bowel disease. This has coincided with the decrease in infectious disease associated with the second epidemiological transition, and the two are related. The “hygiene hypothesis” postulates the rise in CIDs is a result of limited exposure to nonlethal environmental pathogens in utero and early childhood (Zuckerman and Armelagos 2014). Modern human societies have become so sanitized that we are no longer exposed to microorganisms that stimulate the development of a healthy immune system (Versini et al. 2015). “In effect, the lifestyle changes—sanitary improvements, pasteurization, use of antibiotics, and improved hygiene—that contributed to the second transition may have produced a substantial trade-off, with developed nations exchanging a high burden of infectious disease for a higher burden of CIDs” (Zuckerman et al. 2014).
The third epidemiological transition, the re-emergence of infectious disease, reflects the continuing relationship between humans, animals, and pathogens. Over 60% of emerging infectious diseases (EIDs) since 1940 have been of zoonotic origin, with over 70% stemming from human contact with wildlife (Jones et al. 2008), including COVID-19. The crossover of COVID to humans is believed to have involved transmission from bats to an intermediate species then to humans, with infected humans then passing it to other humans in a wet market in Wuhan, China in late 2019 (Worobey et al. 2022). Two COVID variants, representing two distinct crossover events from animals to humans, were circulating in the market by February 2020. Similarly, the global bushmeat trade currently devastating Africa’s wildlife is a continuing source of Ebola infection (Asher 2017), as well as the original source of HIV and viruses related to leukemia and lymphoma among humans (Zuckerman et al. 2014). New strains of avian (bird) flu, some with mortality rates as high as 60% among humans (WHO n.d.), are transmitted to humans through poultry production and contact with wild birds (Davis 2005). Lastly, the use of antibiotics in commercial meat production is directly related to the rise of drug-resistant strains of previously controlled infectious diseases. An estimated 80% of antibiotics in the U.S. are used to promote growth and prevent infection in livestock, and drug-resistant bacteria from these animals are transmitted to humans through meat consumption (Ventola 2015).
A fourth epidemiological transition is currently underway in which some parts of the globe are suffering from a “triple burden” of infectious and chronic diseases combined with injuries and diseases related to intensifying globalization, urbanization, deforestation, and climate change (Karn and Sharma 2021). Massive flooding in Pakistan in 2022 (Figure 16.3) will serve to illustrate the concept. Following a severe heat wave in June 2022, Pakistan experienced extremely heavy seasonal monsoon rains, in some provinces 700% above normal. Combined with water flow from melting glaciers, this caused the worst flooding in the country’s history, putting one third of the nation under water (Sheerazi 2022). The heat wave, glacial melt, and extreme rainfall were all attributable to global climate change, inflicting destruction and disease on Pakistan, which produces less than 1% of total global carbon emissions (Government of Pakistan 2021).
As a direct result of the flooding, infrastructure, including roads, homes, and bridges, was destroyed, and 1,700 people died, nearly 13,000 were injured, and over 33 million were displaced. In addition to their initial injuries and trauma, displaced people lacked food, health care, safe water, and basic sanitation, leading to starvation and exposure to infectious diseases like malaria and dengue fever, as well as skin conditions like scabies, caused by mites. Pakistan also has a poverty rate of 30–40%, contributing to already-high rates of HIV, tuberculosis, and hepatitis. At the same time, the leading causes of death are heart disease, cancer, lower respiratory diseases, and stroke (CDC n.d), all chronic conditions.
These examples illustrate continuing interaction between humans, our evolved biology, and the physical and cultural environments in which we live. The remainder of this chapter will focus on selected diseases and the social, cultural, and environmental factors that contribute to their prevalence in modern, industrialized economies. We begin with the health condition that affects all of the others—obesity.
Obesity
According to the World Health Organization (2017), 1.9 billion of the world’s people are overweight and 650 million of these are obese. In the United States, 70% of Americans are overweight, and 40% of these meet the criteria for obesity. For the first time in human history, most of the world's population lives in countries where overweight and obesity kill more people than hunger (Figure 16.4). Improvements in public health and food production have allowed a greater number of people to live past childhood and to have enough to eat. This does not include everyone. Many people still struggle with poverty, hunger, and disease, even in the wealthiest of nations, including the United States. On a global scale, however, many people not only have enough food to survive but also to gain weight—enough extra weight to cause health problems.
Although studies show differences in daily energy expenditure between foraging and farming populations compared with industrialized peoples, the major contributor to obesity in Western populations is energy intake (Pontzer et al. 2012). Many people not only eat too much but too much of the wrong things. Biological anthropologist Leslie Lieberman (2006) argues that contemporary humans continue to rely on cues from foraging strategies of our evolutionary past that are now counterproductive in the obesogenic environments in which we now live.
Examine your own eating habits in the context of how humans once hunted and gathered. Humans once relied on visual cues to find food, often traveled long distances to obtain it, then transported it back to our home base. There they may have had to process it by hand to render it edible. Think of how much less energy it takes to find food now. If we have the financial resources, we can acquire big energy payoffs by simply sitting at home and using an app on our mobile phone to place an order for delivery. And, voila! High-calorie (if not highly nutritious) food arrives at our door. Should we venture out for food, search time is reduced by signage and advertising directing us toward high-density “patches” where food is available 24 hours a day. These include vending machines, gas stations, and fast-food outlets. Travel time is minimal and little human energy is used in the process (Lieberman 2006).
Foods are also prepackaged and prepared in ways that allow us to eat large quantities quickly. Think French fries or chicken nuggets, which we can easily eat with our hands while doing other things, like driving or watching television, rendering eating mindless and allowing us to take in food faster than our endocrine systems can tell us we are getting full. Modern “patches” offer low-fiber, calorie-dense resources, which allow us to eat larger quantities, a problem already encouraged by larger portion sizes (Lieberman 2006). Processed foods are also engineered to appeal to human preferences for sweet tastes and fatty, creamy textures (Moss 2013). Remember from earlier chapters that natural selection favored depth perception, color vision, grasping hands, and coordinated eye-hand movements as general primate traits. Advertising and packaging now use our color vision against us, attracting us to products that have little nutritional value but that play to our evolutionary predisposition toward variety. Remember those 50 different nutrients we require? That variety is now presented to us in the form of 55 different flavors of Oreo cookies (Cerón 2017), which we take out of the package and dip in milk using our hand-eye coordination and depth perception.
Even if we are ostensibly eating the same things our ancestors did, these foods are nothing alike. Take potatoes, for example. One medium-sized, plain, baked potato is a healthy food, especially if we eat the skin too. It contains 110 calories, 0 grams of fat, 26 grams of carbohydrates, and 3 grams of protein, plus 30% of the U.S. Recommended Daily Allowance (RDA) of vitamin C, 10% of vitamin B6, 15% of potassium, and no sodium (Potato USA). In contrast, a medium order of McDonald’s fries, which takes the potato and adds salt and fat, contains 340 calories, 16 grams of fat, 44 grams of carbohydrates, 4 grams of protein, and 230 mg of sodium (McDonalds). Potato chips take food processing to a whole new level, removing even more nutrition and adding a host of additional ingredients, including oils, preservatives, and artificial flavorings and colors (Moss 2013). Take Ruffles Loaded Bacon and Cheddar Potato Skins Potato Chips as an example (St. Pierre 2018). The number of ingredients increases from one to 11 to 35 as we move from the potato to the potato chip, moving further from nature with each step (Figure 16.5). It should be noted that the nutritional information for the potato chips is based on a serving size of 11 chips, an amount likely smaller than many people eat. Many sweet, fatty, salty foods like fries and chips are cheap, which is why many people choose to eat them (Moss 2013). The price of a medium-sized order of McDonald’s fries as of this writing is US$1.79, and the potato chips are $2.98 for an 8.5-ounce bag. A single potato prewrapped for microwaving is available in many supermarkets for US$1.99 but requires access to a microwave and eating utensils, making it less convenient.
Baked Potato |
French Fries [Medium order] |
Potato Chips [1 oz. serving of 11 chips] |
|
Calories | 110 | 340 | |
Calories from fat | 0 | 144 | |
Fat | 0g | 16g | |
Carbohydrates | 26g | 44g | |
Protein | 3g | 4g | |
Sodium | 0g | 230mg | |
Dietary fiber | 2g | 4g | |
Sugars | 1g | 0g | |
Cholesterol | 0g | 0g | |
Ingredients | Potato | Potatoes, vegetable oil (canola oil), soybean oil, hydrogenated soybean oil, natural beef flavor (wheat and milk derivatives), citric acid (preservative), dextrose, sodium acid pyrophosphate (main color), salt. | Potatoes, vegetable oil (sunflower, corn, and/or canola oil), bacon and chedder loaded potato skins seasoning (maltodextrin - made from corn) salt, cheddar cheese (milk, cheese cultures, salt enzymes), sour cream (cultured cream, skin milk), whey, dried onion, monosodium. |
Not only have we transformed the food supply and our eating in ways that are detrimental to our health, but these changes have been accompanied by reductions in physical activity. Sedentarism is built into contemporary lifestyles. Think of how much time you spent sitting down today. Some of it may have been in class or at work, some may have been driving a car or perhaps binge-watching your favorite show, playing a video game, or checking in on social media. An inactive lifestyle is almost dictated by the digital age (Lucock et al. 2014). Levels of physical activity in many countries are now so low that large portions of the population are completely sedentary, including one in five Americans (CDC 2022). For a species whose biology evolved in an environment where walking, lifting, and carrying were part of daily life, this is unhealthy and often leads to weight gain.
Biology and Genetics of Weight
Research indicates multiple genetic variants influence weight gain, and they are not spread evenly among human populations. Tuomo Rankinen and colleagues (2006) identified 127 genes associated with obesity, of which 22 contributed to weight gain. Claude Bouchard (2007) then identified five categories of obesity-promoting genotypes. These genotypes promote sedentarism, result in low metabolism, and lead to poor regulation of appetite, and a propensity to overeat. An example of the impact such genotypes can have in an environment of plenty is found among the population of the Micronesian island of Nauru. Historically, the island was geographically isolated and the food supply was unpredictable. These conditions favored genotypes that promoted the ability to rapidly build up and store fat in times of food availability. In Nauruans, there are two genetic variants favoring weight gain and insulin resistance, and both are associated with obesity and type 2 diabetes. One variant is also associated with hypertension. One of these variants is also found in Pima Indians, who live in parts of Arizona and Mexico. In the Pima, this variant is associated with a high body mass index (BMI) and type 2 diabetes, although it is not associated with the same outcomes in Japanese and British subjects (de Silva et al. 1999). The other variant was analyzed in Finnish and South Indian populations, neither of whom experienced the same outcome as Nauruans. This suggests these alleles may act as modifying genes for type 2 diabetes in some population groups (Baker et al. 1994). Unfortunately, Nauruans are one of those groups. Eventually, they became wealthy through phosphate mining on the island, gaining access to a calorie-rich Western diet of imported foods and developing a sedentary lifestyle. This resulted in rates of type 2 diabetes as high as 30–40% in Nauruans over the age of 15, which became the leading cause of death (Lucock et al. 2014), something Nauruans are taking seriously (Figure 16.6).
Factors other than biology influence which populations that carry a genetic predisposition to diabetes actually express it. The Pima Indians of Arizona, for example, were seriously impacted by U.S. government policies that affected water rights, forcing the population away from subsistence farming to dependence on government handouts and convenience food. This resulted in a significant loss of physical activity, malnutrition, and obesity. The Pima continue to experience hardship due to high rates of unemployment, poverty, and depression, sometimes made worse by alcoholism. In the absence of these pressures, the Pima were diabetes free for centuries prior, even though they relied on agriculture for subsistence, suggesting genetics alone is not responsible for high rates of obesity and diabetes in current Pima Indian populations (Smith-Morris 2004).
(Dig Deeper --> Colonization and Indigenous health?)
Obesity also has an epigenetic component. You learned about epigenetics in Chapter 3. With regard to obesity, epigenetics is counterintuitive in that mothers who do not take in enough calories during pregnancy often give birth to babies who grow up to be fat. What takes place is the fetus receives signals during pregnancy from its mother through the placenta and intrauterine environment about environmental conditions outside of the womb, in this case food insecurity. These signals encourage the turning on and off of genes related to metabolism. This alters the phenotype of the fetus so that if the child is born into an environment where food is plentiful, it will put on weight rapidly whenever possible, leading to obesity and related diseases later in life. If the child is a girl, her own eggs are formed in utero with the same genetic changes coded in, meaning she will pass along this same genetic predisposition to gain weight to her children. Hence, a biological propensity toward obesity can continue across generations (Worthman and Kuzara 2005). Epigenetic changes to genes that promote weight gain are argued to be partly responsible for the rapid rise in obesity and diabetes in developing countries gaining access to Western diets (Stearns, Nesse, and Haig 2008).
Obesity and overweight put a strain on several biological systems of the body, including the circulatory, endocrine, and skeletal systems, contributing to hypertension, heart disease, stroke, diabetes, and osteoarthritis (Figure 16.7). Obesity also elevates the risk of cancers of the breast, endometrium, kidney, colon, esophagus, stomach, pancreas, and gallbladder (National Institutes of Health 2017; Vucenik and Stains 2012). Diabetes—one of the fastest-growing health conditions around the globe (WHO 2016) and one tightly connected to obesity and overweight—is the focus of the following Special Topics box.
Special Topic: Diabetes
Diabetes mellitus is an endocrine disorder characterized by excessively high blood glucose levels (Martini et al. 2013). According to a report released by the World Health Organization, the number of people living with diabetes is growing in all regions of the world. Rates of diabetes have nearly doubled in the past three decades, largely due to increases in obesity and sugary diets (WHO 2016). One in 10 people around the world, 537 million people, now have diabetes, and three out of four live in low- and middle-income countries (IDF 2022). In the United States, 37 million people have diabetes (CDC 2020), where the disease is rising fastest among millennials (those ages 20–40) (BCBSA 2017), and one in every two adults with diabetes is undiagnosed (IDF 2022). Obesity and diabetes are linked: obesity causes a diet-related disease (diabetes) because of humans’ evolved metabolic homeostasis mechanism, which is poorly suited to contemporary energy environments (Lucock et al. 2014).
To function properly, cells need a steady fuel supply. Blood sugar (glucose) is the fuel for most cells in the body, and the body produces the hormone insulin to help move glucose into cells that need it (Figure 16.8). Foods that most readily supply glucose to your bloodstream are carbohydrates, especially starchy foods like potatoes or sweet, sugary foods like candy and soda. The body can also convert other types of foods, including protein-rich foods (e.g., lean meats) and fatty foods (e.g., vegetable oils and butter), into blood sugar in the liver via gluconeogenesis. Insulin’s main job is to tell your cells when to take up glucose. The cell also has to listen to the signal and mobilize the glucose transporters. This not only allows your cells to get the energy they need, but it also keeps blood sugar from building up to dangerously high levels when you are at rest.
This system has limits. Like the rest of our biology, it evolved during several million years when sugar was hard to come by and carbohydrates took the form of fresh foods with a low glycemic index (GI). Our ancestors were also active throughout the day, taking pressure off of the endocrine system. Now, sedentary lifestyles and processed-food diets cause many of us to take in more calories—and especially more carbohydrates—than our bodies can handle. There is only so much blood sugar your cells can absorb. Many modern populations are taxing those limits. After years of being asked by insulin to take in more glucose than they can use, cells eventually stop responding (McKee and McKee 2015). This is called type 2 diabetes or insulin resistance, which accounts for 90–95% of diabetes cases in the United States (Figure 16.9). type 1 diabetes is believed to be caused by an autoimmune response in which your immune system is attacking and destroying the insulin-producing cells in your pancreas (Figure 16.9). type 1 diabetes is a genetic condition that often shows up early in life, while type 2 is more lifestyle-related and develops over time.
Cardiovascular Disease
Cardiovascular disease (CVD)—which includes coronary heart disease, hypertension (high blood pressure), and stroke—is the leading cause of death globally, and heart disease remains the number one cause of death in the United States (American Heart Association 2018). Risk factors for cardiovascular disease include diet, obesity/overweight, diabetes, smoking and alcohol consumption, and physical inactivity.
The connections between these factors and heart disease may not seem obvious and will be addressed here beginning with diet. Diets high in saturated fat and cholesterol can lead to atherosclerosis, a condition in which fat and cholesterol form plaque inside the arteries, eventually building up and hardening to the point that blood flow is blocked. Too much salt in the diet leads to fluid retention, which increases blood volume and thereby blood pressure, taxing the heart. Obesity/overweight contribute to cardiovascular disease directly through increases in total blood volume, cardiac output, and cardiac workload. In other words, the heart has to work much harder if one is overweight (Akil and Ahmad 2011). Obesity also relates to CVD indirectly through elevation of blood pressure (hypertension) and diabetes. High levels of blood glucose from diabetes can damage blood vessels and the nerves that control the heart and blood vessels. Alcohol consumption can raise blood pressure and triglyceride levels, a type of fat found in the blood. Alcohol also adds extra calories, which may cause weight gain, especially around the abdomen, which is directly associated with risk of a heart attack (Akil and Ahmad 2011). Cigarette smoking also increases the risk of coronary heart disease. Nicotine increases blood pressure; in addition, cigarette smoke causes fatty buildup in the main artery in the neck and thickens blood, making it more likely to clot. It also decreases levels of HDL (“good”) cholesterol (American Heart Association 2018). Even secondhand smoke can have an adverse effect if exposure occurs on a regular basis. Chronic psychological stress also elevates the risk of heart disease (Dimsdale 2008). The repeated release of stress hormones like adrenaline elevates blood pressure and may eventually damage artery walls. The human stress response and its connections to health and disease are discussed in more detail below.
However, physical activity alters the likelihood of having heart disease, both directly and indirectly. Regular exercise of moderate to vigorous intensity strengthens the heart muscle and allows capillaries, tiny blood vessels in your body, to widen, improving blood flow. Regular exercise can also lower blood pressure and cholesterol levels and manage blood sugar levels, all of which reduce the risk of CVD.
Cancer
Cancer is the second-leading cause of death globally, causing one in every six deaths and killing nearly nine million people in 2015 (WHO 2018). Lifetime cancer risk in developed Western populations is now one in two, or 50% (Greaves 2015). Approximately one-third of deaths from cancer are due to behavioral and dietary factors, including high body mass index (BMI), low fruit and vegetable intake, lack of physical activity, and the use of tobacco and alcohol. Depending on the type of cancer and one’s own genetic inheritance, these factors can increase cancer risk from 2- to 100-fold (Greaves 2015). Cancer is the result of interactions between a person's genes and three categories of external agents: physical carcinogens (e.g., ultraviolet radiation), chemical carcinogens (e.g., tobacco smoke, asbestos), and biological carcinogens, such as infections from certain viruses, bacteria, or parasites (WHO 2018). Obesity is also a risk factor for cancer, including of the breast, endometrium, kidney, colon, esophagus, stomach, pancreas, and gallbladder (National Institutes of Health 2017; Vucenik and Stains 2012).
Cancer has been regarded as a relatively recent affliction for humans that became a problem after we were exposed to modern carcinogens and lived long enough to express the disease (David and Zimmerman 2010). Given the long history that humans share with many oncogenic (cancer-causing) parasites and viruses (Ewald 2018), and the recent discovery of cancer in the metatarsal bone of a 1.8-million-year-old hominin (Odes et al. 2016), this view is being challenged (See “Special Topic: Life Choices and Reproductive Cancers in Women”). The difficulties of identifying cancer in archaeological populations are many. Most cancer occurs in soft tissue, which rarely preserves, and fast-growing cancers would likely kill victims before leaving evidence in bone. It is also difficult to distinguish cancer from benign growths and inflammatory disease in ancient fossils, and there is often postmortem damage to fossil evidence from scavenging and erosion. However, using 3-D images, South African researchers recently diagnosed a type of cancer called osteosarcoma in a toe bone belonging to a human relative who died in Swartkrans Cave between 1.6 and 1.8 million years ago (Randolph-Quinney et al. 2016). This study provides the earliest evidence of cancer in hominins.
Special Topic: Life Choices and Reproductive Cancers in Women
Behavioral or “lifestyle” choices have an impact on cancer risk. Breast cancer is one example. It is the most common cancer in women worldwide, but incidence of new cases varies from 19.3 per 100,000 women in Eastern Africa to 89.7 per 100,000 women in Western Europe (WHO 2018). These differences are attributable to cultural changes among women in Western, industrialized countries that are a mismatch for our evolved reproductive biology. Age at menarche, the onset of menstrual periods, has dropped over the course of the last century from 16 to 12 years of age in the U.S. and Europe, with some girls getting their periods and developing breasts as young as eight years old (Greenspan and Deardorff 2014, Figure 16.10). A World Health Organization study involving 34 countries in Europe and North America suggests the primary reason for the increase in earlier puberty is obesity, with differences in BMI accounting for 40% of individual- and country-level variance (Currie et al. 2012). Early puberty in girls is associated with increased risk of breast cancer, ovarian cancer, diabetes, and high cholesterol in later life (Pierce and Hardy 2012).
At the same time that age at puberty is dropping for girls in Western nations, age at birth of the first child is later, at 26 years old (Mathews and Hamilton 2016). Women are also having fewer children, two on average (Gao 2015), with 15% of women choosing to remain childless (Livingston 2015). Rates of breastfeeding have risen in recent decades but drop to only 27% of infants once babies reach 12 months of age (CDC 2014). In contrast, data from modern foraging populations (Eaton et al. 1994) indicate that age at menarche is around 16 years old, age at birth of the first child is 19, breastfeeding on-demand continues for three years for each child, and the number of children averages six. These differences relate to elevated risk for reproductive cancers, including breast cancer, among women in developed countries.
Other than an established genetic risk (e.g., BRCA gene), the primary risk factor for breast cancer is exposure to estrogen. For women living in modern, industrialized economies, this exposure now often comes from women’s own ovaries rather than from external environmental sources (Stearns, Nesse, and Haig 2008). There is nothing biologically normal about regular monthly periods. Women in cultures without contraception are pregnant or breastfeeding for much of their reproductive lives, resulting in 100 or so menstrual cycles per lifetime. In contrast, Western women typically experience 400 or more (Strassmann 1997). This is partly due to early puberty. From menarche to the birth of a woman’s first child can be 14 years or longer in Western populations, after which breastfeeding, if undertaken at all, lasts for a few weeks or months. Oral contraceptives or other hormonal methods to control reproduction induce monthly periods. Age at menopause (the cessation of menstrual cycles) is 50–55 years old across human populations. For Western women, this translates into forty years of menstrual cycling. Each month the body prepares for a pregnancy that never occurs, experiencing cell divisions that put women at risk for cancers of the breast, endometrium, ovaries, and uterus (Strassmann 1999). Obesity adds to risk, as adipose (fat) tissues are the main source of estrogen biosynthesis. Thus, weight gain during the postmenopausal stage means higher exposure to estrogen and a greater risk of cancer (Ali 2014).
Women cannot return to our evolutionary past, and there are important social and economic reasons for delaying pregnancy and having fewer children. These include achieving educational and career goals, greater earning power, a reduction in the gender pay gap, more enduring marriages, and a decrease in the number of women needing public assistance (Sonfield et al. 2013). There are also cultural means by which we might reduce the risk of reproductive cancers. These include reformulating hormonal contraceptives with enough estrogen to maintain bone density but reducing the number of menstrual periods over the reproductive lifespan (Stearns, Nesse, and Haig 2008). Reducing fat intake may also lower estrogen levels. High-fat diets contribute to breast tumor development, while high fiber diets are beneficial in decreasing intestinal resorption of estrogenic hormones. Exercise also appears protective. Studies of former college athletes demonstrate risks of breast, uterine, and ovarian cancers later in life two to five times lower than those of nonathletes (Eaton et al. 1994).
Stress
Have you ever been “stressed out” in class? Say you’re in a large lecture hall with a hundred other people, or even in a small class where you don’t know anyone. You’re not sure about something the professor just said and you would really like to ask about it, so you start to raise your hand. Does your heart begin to pound and your mouth become dry? Do you get so nervous that you choose to ask a classmate after lecture instead? If so, you are not alone. Fear of speaking in public is one of the most common social phobias (APA 2013). It has been estimated that 75% of all people experience some degree of anxiety or nervousness when it comes to public speaking (Hamilton 2011), and surveys have shown that most people fear public speaking more than they fear death (Croston 2012).
We have evolution to thank for this.
Humans, like other primates, are social animals. Being part of a group helped us to survive predation, get enough to eat, and successfully raise our young. When faced with standing up in front of a group, or even speaking up in class, we break into a sweat because we are afraid of rejection. Psychologist Glenn Croston (2012) writes, “The fear is so great because we are not merely afraid of being embarrassed or judged. We are afraid of being rejected from the social group, ostracized and left to defend ourselves all on our own. We fear ostracism still so much today it seems, fearing it more than death, because not so long ago getting kicked out of the group probably really was a death sentence.” Hence, it is no surprise that public speaking triggers a stress response among much of humankind.
The human nervous system evolved in a context where quick responses to perceived threats presented an evolutionary advantage. The “fight or flight” response was honed during millions of years when threats more often took the form of an approaching lion than an approaching deadline. Our body’s stress response, however, is triggered by a wide variety of stressors that produce the same general pattern of hormonal and physiological adjustments (Martini et al. 2013). In today’s world, the system is often stuck in the “on” position due to the constant pressures of modern life, and this is a significant influence on health and disease.
It is important to recognize that there are different types of stress and the time in life when adult coping mechanisms are formed is in childhood. In children, some stressors can be positive—for example, stressors that are mild to moderate in magnitude, and accompanied by the support of a caring adult, which help children develop pathways by which stress is dealt with by the body throughout life. In a young child, a positive stress response might be going to the pediatrician to receive a vaccination and receiving encouragement and comfort from both parent and practitioner. A tolerable stress response is more serious, precipitated by something like a divorce or death of a relative. Again, buffered by positive support from surrounding adults, these types of stressors can be successfully managed by children. Toxic stress, however, “results from strong, frequent or prolonged activation of the stress response in the absence of the buffering protection of a supportive adult relationship” (Shonkoff and Garner 2012). Examples include child abuse or neglect, parental substance abuse, homelessness, and violence. In the absence of adequate psychological and physical support, the biological pathways of a child’s physiological stress response are altered and lead to reduced abilities to cope with life’s challenges as an adult.
The negative effects of sustained, elevated cortisol levels on health are well documented. These include higher levels of infectious disease and slowed growth in childhood (Flinn and England 2003) and increased incidence of heart disease, obesity, and diabetes in adults (Worthman and Kuzara 2005). Contrary to our evolutionary past, many causes of sustained stress in contemporary societies are psychosocial rather than physical threats. These can include an unhappy marriage or frustrations at work (Dimsdale 2008). Stressors can also be more subtle. For example, a review of research into the effects of stress on health indicated that experiencing racism was a significant stressor that was associated with alcohol consumption, psychological distress, overweight, abdominal obesity, and higher fasting-glucose levels among minority groups (Williams and Mohammed 2013). Chronic, everyday racial discrimination is also associated with the hardening of coronary arteries, elevated blood pressure, giving birth to lower-birth-weight infants, cognitive impairment, poor sleep, and visceral fat, which is fat stored deep inside the belly, wrapped around the organs, including the liver and intestines. Visceral fat is a sign of metabolic syndrome, increasing the risk of stroke, heart disease, and type 2 diabetes. These effects have been shown to increase morbidity and mortality among members of affected groups.
Epigenetics can also be a factor in how a person is able to deal with stressful situations. Maternal experiences of stress during pregnancy have the potential to permanently alter the physiology of mothers’ offspring, especially the hypothalamic-pituitary-adrenal (HPA) axis. The HPA axis regulates metabolism, blood pressure, and the immune response, and these alterations can predispose prenatally stressed individuals to suffer metabolic, cardiovascular, and mental disorders in adulthood (Palma-Gudiel et al. 2015). These experiences carry across generations, with children of Holocaust survivors who experienced PTSD demonstrating similar changes in neurochemistry in the absence of a sustained, traumatic event, as did infant offspring of mothers who developed PTSD during pregnancy after witnessing the traumatic events of 9/11 (Yehuda and LeDoux 2007).
Syndemics and the Ecological Model
It is important to recognize that disease risk is not spread evenly within or between populations. Diseases combine and interact to create a syndemic, where the coexistence of two or more conditions exacerbates the effects of one or all conditions. A syndemic (versus a pandemic, for example) takes into account social, political, economic, and environmental factors that increase risk for the clustering of two or more diseases (Singer et al. 2017). One of the first syndemics identified involved substance abuse, violence, and AIDS. In inner cities in the U.S., the health crisis around HIV/AIDS was related to tuberculosis, sexually transmitted infections, hepatitis, cirrhosis, infant mortality, drug abuse, suicide, and homicide. These were connected to poverty, homelessness, unemployment, poor nutrition, lack of social support, and social and ethnic inequality (Singer et al. 2017). Together, these factors and others, like health policy and unequal access to health care, form an ecological model of health and disease, one that moves beyond biology and medical intervention (Sallis et al. 2008).
The COVID-19 pandemic represents a syndemic in which systemic racism in the healthcare system, differential access to diagnosis and treatment, income, employment, housing, family structure, pre existing conditions, and public health policies combined to result in higher rates of infection and death for African Americans, Native Americans, Asians, and Hispanic populations in the United States (Figure 16.11).
Rate ratios compared to White, Non-Hispanic persons | American Indian or Alaska Native, Non-Hispanic persons | Asian, Non-Hispanic persons | Black or African American, Non-Hispanic persons | Hispanic or Latino persons |
Cases | 1.6x | .8x | 1.1x | 1.5x |
Hospitalization | 2.7x | .8x | 2.3x | 2.0x |
Death | 2.1x | .8x | 1.7x | 1.8x |
COVID-19 was the third leading cause of death in the U.S. in 2020 and 2021 (NIH 2022; Figure 16.12), but morbidity and mortality was not equally spread across the population. Working-class people and people of color in the U.S. are more likely to live in poverty, in areas with high rates of crime and violence, and in close proximity to freeways and environmental threats like petrochemical plants and waste incinerators (Singer and Baer 2012). Many such neighborhoods are also food “deserts” without ready access to a healthy, affordable diet, made more challenging by residents not owning a car (Food Empowerment Project n.d.). Low-income people also often lack access to high-quality health care and delay or avoid preventive care and health screenings (Ross et al. 2007). These factors contributed to higher rates of preexisting conditions, including obesity, diabetes, hypertension, asthma, heart disease, chronic obstructive pulmonary disease (COPD), and smoking behavior, which then led to more complications and higher death rates from COVID (Ghosh et al. 2021).
Family structure also affected COVID exposure and severity. Many Americans live in multigenerational households, including 27% of Hispanics, 29% of Asians, 26% of African Americans, and 20% of Whites (Cohn and Passel 2018). Not all multigenerational households are equal, however. Over twice as many African Americans as Whites are in multigenerational families in which at least one family member is unemployed, and over three times as many African Americans are in multigenerational families in which everyone is simultaneously unemployed (Park, Wiemers, and Seltzer 2019). Family members in multigenerational households were at a much higher risk of developing more severe forms of COVID due to decreased personal space and multiple exposures to the virus, as well as higher rates of diabetes, smoking, and residents living below the poverty line (Ghosh et al. 2021). While aimed at reducing overall infection rates from COVID, public health measures such as mandatory lockdowns only exacerbated the situation in overcrowded and multigenerational housing, resulting in higher rates of infection and death in these communities.
United States |
Worldwide |
1. Heart disease |
1. Heart disease |
2. Cancer |
2. Stroke |
3. COVID-19 |
3. COVID-19 |
4. Accidents |
4. Chronic Obstructive Pulmonary Disease |
5. Stroke |
5. Lower respiratory infections |
There is a long history of systemic racism and discrimination in the medical system in the United States (Washington 2006). African Americans have been subjected to medical testing and experimentation without their consent or knowledge since the time of slavery. They continue to routinely receive care of poorer quality than whites (Williams and Wyatt 2015), less pain medication during treatment and hospitalization (Green et al. 2003), and differential treatment during pregnancy and childbirth (Washington 2006). Many Americans, including 50% of White medical students and residents in one recent study (Hoffman et al. 2016), hold at least one false belief about African Americans, including “Black people’s skin is thicker than white people’s skin,” “Blacks have stronger immune systems than whites,” and “Blacks’ nerve endings are less sensitive than whites’.” Such beliefs affect health care for African Americans in medical emergencies and for chronic conditions.
During the COVID-19 pandemic, patients with darker skin in the United States were negatively affected by the very medical device most commonly used to assess oxygen levels in their blood. The pulse oximeter, a small device that clips onto the tip of your index finger and measures blood oxygen levels, experienced increased use in home, clinical, and hospital settings during the COVID-19 pandemic. Decisions regarding treatment and hospital admission for patients infected with COVID were often based on pulse oximeter readings (Valbuena, Merchant, and Hough 2022). The problem is the device overestimates oxygen saturation in patients with darker skin, an issue which has been recognized for over thirty years (Valbuena, Merchant, and Hough 2022). It would be as if a standard thermometer reported lower body temperatures for patients of color, making it seem as if they did not have a fever when they actually did. In the case of COVID-19, Asians, Hispanics, and African Americans experienced inaccurately high readings of their oxygen levels (with African Americans and darker-skinned Hispanics having the highest), resulting in delays in treatment, hospital admission, and access to medications to treat COVID and contributing to higher severity of illness and higher death rates among these populations in comparison to whites (Fawzi et al. 2022).
Employment was also a factor in unequal exposure to and death from COVID-19 (Raifman, Skinner, and Sojourner 2022), with many low-income workers making the choice (which, realistically, may not be a choice at all) to expose themselves to COVID in order to earn the funds necessary to purchase food, housing, and other necessities. Many such workers were then forced to miss work due to COVID infection. With only 35% of low-wage workers (as opposed to 95% of high-wage workers) having paid sick leave, this left many families struggling financially. Three years into the pandemic, low-wage workers continue to have the least access to COVID vaccines and boosters. The U.S. also lacks federal workplace-safety regulations with regard to vaccine and masking mandates that other nations enforce in times of high transmission, and it does not provide high-quality masks to its essential workers. Many occupations deemed essential by the CDC during the height of the pandemic—such as health care, emergency services, meat packing, agricultural work, teaching, and jobs in the hospitality sector—experienced higher rates of morbidity and mortality from COVID. Many of these fields disproportionately employ people of color (McKinsey and Company 2021). Given this, future policies that address the pandemic at a structural level—for example, providing monetary assistance to people who work in environments with a high risk of infection, such as cleaning, nursing, transportation, retail, restaurant work, and factory work, so that they can remain at home—may function more effectively to prevent transmission and curb future outbreaks (Arnot et al. 2020).
Food for Thought
This chapter focused primarily on health conditions prevalent in contemporary, industrialized societies that are due, in part, to the mismatch between our evolved biology and modern environments. These are the built environment and the social environment, which together form the obesogenic environment in which unhealthy behaviors are encouraged. This chapter will close by examining each of these in a college context.
Consider your campus from an evolutionary perspective. To what degree does the built environment lend itself to physical activity as part of daily life? Is your campus constructed in ways that promote driving at the expense of walking or biking? If driving is necessary, is parking available close to the buildings or do you need to walk a fair distance from the parking lot to your destination? Do the buildings have stairs or ramps or is it necessary to take the elevator? Is it possible to negotiate safely around campus on foot or by bike in all weather? After dark? How about the classrooms and computer labs? Do they have standing or treadmill desks? Does your class schedule encourage walking from building to building between classes, or are most courses in your major scheduled in the same location? Most college majors also lack a physical education requirement, leaving it up to students to incorporate exercise into already-challenging schedules (Figure 16.13).
Sociocultural factors that contribute to obesity include food advertising, ubiquitous fast-food and junk food options, and social pressure to consume, all of which are present on college campuses. Although nutrition on campuses has improved in recent years, many students find eating healthy in the dining halls and dorms challenging (Plotnikoff et al. 2015), and students who live off campus fare even worse (Small et al. 2013). There are also parties and other social events, a normal part of college life, that involve unhealthy food and encourage behaviors like alcohol consumption and smoking. Give some thought to the social atmosphere on your campus and the ways it may contribute to obesity. My own freshman orientation involved a succession of pizza parties, ice cream socials, and barbecues, followed by late-night runs to the nearest fast-food outlet. The purpose of these events was to encourage people to make friends and feel comfortable living away from home, but the foods served were unhealthy, and there was social pressure to join in and be part of the group. Such activities set students up for the “freshman fifteen” and then some. They also reinforce the idea that being social involves eating (and sometimes drinking and/or smoking).
Sedentarism and inactivity are also built into the academics of college life. Digital technology is a significant contributor to obesity. Students use laptops and cell phones to take notes, complete their work outside of class, and access social media. There are also video games, virtual reality headsets, and streaming television and movies for entertainment. The built environment of college already necessitates that students sit in class for hours each day, then sit at computers to complete work outside of class. The social environment enabled by digital technology encourages sitting around for entertainment. It is telling that we call it “binge watching” when we spend hours watching our favorite shows. Doing so often involves eating, as well as multiple exposures to food advertising embedded in the shows themselves. In these ways, college contributes to the development of obesity-causing behaviors that can have negative health ramifications long after college is over (Small et al. 2013).
In the U.S., the greatest increase in obesity is among young adults aged 18–29 years, a significant percentage of whom are college students (Plotnikoff et al. 2015). Analyses of college students’ behavior across semesters shows consumption of fruits and vegetables drops over time, as does the amount of physical activity, while consumption of sugar-sweetened beverages and fast-food goes up, leading to weight gain at nearly six times the rate of the general public (Small et al. 2013). In response, many colleges and universities have instituted programs to encourage healthier eating and more physical activity among students (Plotnikoff et al. 2015). It is important to emphasize that neither changes in diet or exercise are effective on their own.. A 2022 study of over 340,000 British participants demonstrated that physical activity and diet quality did not individually have an impact on cardiovascular disease or cancers (Ding et al. 2022). That is, hitting the gym won’t counteract the consequences of consuming high-calorie, fatty foods, and eating kale all day can’t cancel out sedentary habits.
Just as no one fad diet is going to prove healthier than another, no one type of exercise is better than another. Anything that raises your heart rate and that you enjoy doing for at least an hour each day will work. Take advantage of opportunities to build exercise into everyday life. Take the stairs, park as far away from buildings as possible, ride a bike or walk instead of driving, and take walks between classes instead of sitting down and checking your phone. As far as diets go, eating a few less unhealthy calories each day, one less soda, no sugar in your coffee, or letting that last slice of pizza go to someone else, make a difference in the long run. Little changes add up to bigger ones. We cannot change our biology, but we can certainly change our habits.
Review Questions
- Why do humans like foods that are “bad” for them? Describe the evolutionary underpinnings of our tastes for sugar, salt, and fat.
- How might understanding contemporary disease from an evolutionary perspective benefit medical practitioners in treating their patients?
- Several risk factors for conditions like heart disease, diabetes, and cancer are referred to as “lifestyle factors,” implying these are behavioral choices people make that put them at risk. These include unhealthy eating, lack of physical activity, smoking, and alcohol consumption. To what degree is unhealthy behavior structured by the physical and social environment? For example, how does being a college student influence your eating habits, physical activity patterns, smoking, and consumption of alcohol?
- Who benefits from the global obesity epidemic? Think about how the following industries and institutions might profit from it: The medical establishment? The fitness industry? The diet industry? Fashion? Pharmaceutical companies? Food manufacturers? Advertisers?
Key Terms
Body mass index (BMI): A person’s weight in kilograms divided by the square of their height in meters. This is the most widely used measure for identifying obesity. The formula using kilograms and meters (or centimeters) is: weight (kg) / [height (m)]2 . The formula using pounds and inches is: 703 x weight (lbs) / [height (in)]2 . Use of the BMI is controversial for several reasons, including that it does not take into account age, bone structure, muscle mass, fat distribution, or ethnic and racial differences in body type.
Cancer: A collection of related diseases in which some of the body’s cells begin to divide without stopping and spread into surrounding tissues.
Cardiovascular disease (CVD): A disease of the heart and blood vessels, often related to atherosclerosis. CVD is a condition in which a substance called plaque builds up in the walls of the arteries, the blood vessels that carry blood away from the heart, which compromises the flow of blood to the heart or brain.
Central nervous system: The complex of nerve tissues stemming from the brain and spinal cord that controls the activities of the body.
Circulatory (system): The biological system that circulates blood around the body via the heart, arteries, and veins, delivering oxygen and nutrients to organs and cells and carrying waste products away.
Diabetes mellitus: An endocrine disorder in which high glucose (blood sugar) levels occur over a prolonged period of time. Blood glucose is your body’s main source of energy and comes from the food you eat. Insulin, a hormone made by the pancreas, helps glucose from food get into your cells to be used for energy. Sometimes your body does not make enough—or any—insulin (type 1 diabetes) or does not take up insulin well (type 2 diabetes). Glucose then stays in your blood and does not reach your cells.
“Double burden”: Refers to parts of the world in which there is a prevalence of chronic disease (e.g., cancer, heart disease) while, at the same time, there are also high rates of infectious disease due to poverty, malnutrition, poor sanitation, and lack of access to health care, often accompanied by high rates of maternal and child mortality.
Ecological model: Ecological models of health and disease emphasize environmental and policy contexts of behavior, while incorporating social and psychological influences, rather than focusing on individual behaviors. These models encompass multiple levels of influence and can lend themselves to more comprehensive health interventions.
Emerging infectious diseases (EIDs): Infections that have recently appeared within a population or those whose incidence or geographic range is rapidly increasing or threatens to increase in the near future. Examples include Covid-19, Ebola, Zika, SARS, and avian (bird) flu.
Endocrine system: Those organs in the body whose primary function is the production of hormones.
Epidemiological transition: A transformation in patterns of disease (morbidity) and death (mortality) among a population.
Glycemic index (GI): A system that ranks foods on a scale from 1 to 100 based on their effect on blood-sugar levels. Carbohydrates with a low GI value (55 or less) are more slowly digested and metabolized causing a lower, slower rise in blood glucose and insulin levels.
Hypertension: High blood pressure. Blood pressure is the force exerted by the blood against the walls of the blood vessels. In a blood pressure reading, the top number (usually higher) refers to the systolic pressure, the amount of pressure in your arteries during the contraction of your heart muscle when your heart beats. The bottom number is the diastolic pressure when your heart muscle is resting between beats. Hypertension can lead to severe health complications and increases the risk of heart attack and stroke.
Incidence: The rate at which new cases of a disease occur in a population over a given period of time.
Insulin: A hormone produced in the pancreas that regulates the amount of glucose in the blood. Lack of insulin or the inability to absorb insulin causes diabetes.
Metabolic syndrome: A cluster of conditions, including increased blood pressure, high blood sugar, excess body fat around the waist, and abnormal cholesterol levels that occur together, increasing the risk of heart disease, stroke, and diabetes. Lifestyle changes like losing weight, exercising regularly, and making dietary changes can help prevent or reverse metabolic syndrome.
Menarche: The first occurrence of menstruation.
Morbidity: The number of cases of disease per unit of population occurring over a unit of time.
Mortality: The number of deaths attributable to a particular cause per unit of population over a unit of time.
Noncommunicable diseases (NCDs): Also known as chronic diseases, NCDs tend to be of long duration and are the result of a combination of genetic, physiological, environmental, and behavior factors. The main types of NCDs are cardiovascular diseases (like heart attacks and stroke), cancers, chronic respiratory diseases (such as chronic obstructive pulmonary disease and asthma), and diabetes.
Obesity: A medical condition in which excess body fat has accumulated to the point that it has adverse effects on health. Although controversial due to its lack of ethnic and racial specificity, the most widely used measure for identifying obesity is the body mass index (BMI), a person’s weight in kilograms divided by the square of their height in meters. A measure of 30 kg/m2 is considered obese and 25–29 kg/m2 is considered overweight. Distribution of body fat also matters. Fat in the abdominal region has a stronger association with type 2 diabetes and cardiovascular disease, meaning waist-to-hip ratio and waist circumference are also important indicators of obesity-related health risk.
Obesogenic: Promoting excessive weight gain.
Omnivorous: Able to eat and digest foods of both plant and animal origins.
Osteoarthritis: Refers to the degeneration of joint cartilage and underlying bone, causing pain and stiffness. In the absence of previous injury, it is most common in modern populations from middle age onward.
Prevalence: The proportion of individuals in a population who have a particular disease or condition at a given point in time.
Sedentarism: A way of life characterized by much sitting and little physical activity.
Sedentism: Living in groups settled permanently in one place.
Stress response: A predictable response to any significant threat to homeostasis. The human stress response involves the Central Nervous System and the endocrine system acting together. Sudden and severe stress incites the “flight or flight” response from the autonomic nervous system in conjunction with hormones secreted by the adrenal and pituitary glands, increasing our heart rate and breathing and releasing glucose from the liver for quick energy.
Stroke: A stroke occurs when a blood vessel leading to the brain is blocked or bursts, preventing that part of the brain from receiving blood and oxygen, leading to cell death.
Syndemic: The aggregation (grouping together) of two or more diseases or health conditions in a population in which there is some level of harmful biological or behavioral interface that exacerbates the negative health effects of any or all of the diseases involved. Syndemics involve the adverse interaction of diseases of all types, including infections, chronic noncommunicable diseases, mental health problems, behavioral conditions, toxic exposure, and malnutrition.
Tricep skinfold measurement: The triceps skinfold site is a common location used for the assessment of body fat using skinfold calipers. A section of skin on the posterior (back) surface of the arm that lays atop the tricep muscle is pinched between calipers. The resulting measurement is matched against a chart standardized for age and gender.
“Triple burden”: A fourth epidemiological transition currently underway in which some parts of the globe are suffering from the “double burden” of infectious and chronic diseases combined with injuries and diseases related to intensifying globalization, urbanization, deforestation, and climate change.
Vector-borne diseases: Human illnesses caused by parasites, viruses, and bacteria that are transmitted by mosquitoes, flies, ticks, mites, snails, and lice.
Zoonoses: Diseases that can be transmitted from animals to humans.
About the Author
Joylin Namie, Ph.D.
Truckee Meadows Community College, jnamie@tmcc.edu
Joylin Namie is Professor of Anthropology at Truckee Meadows Community College, where she teaches courses in biological and cultural anthropology. Her current research interest is in culturally and environmentally sustainable tourism in desert environments, particularly in the country of Jordan and the U.S. state of Nevada. She was awarded a fellowship to Jordan from the Council of American Overseas Research Centers (CAORC) in 2020 to explore this topic, including visiting Petra and other important tourism destinations in Jordan. Dr. Namie’s favorite things in life are teaching, competing in sports, and traveling.
For Further Exploration
Lents, Nathan H. 2018. Human Errors: A Panorama of Our Glitches, from Pointless Bones to Broken Genes. Boston: Houghton Mifflin Harcourt.
Stearns, Stephen C., and Jacob C. Koella, eds. 2008. Evolution in Health and Disease. 2nd edition. United Kingdom: Oxford University Press.
Zuk, Marlene. 2013. Paleofantasy: What Evolution Really Tells Us about Sex, Diet, and How We Live. New York: W. W. Norton & Company.
References
Abid, Zaynah, Amanda J. Cross, and Rashmi Sinha. 2014. “Meat, Dairy, and Cancer.” The American Journal of Clinical Nutrition 100 (S1): 386S–393S.
Akil, Luma, and H. Anwar Ahmad. 2011. “Relationships between Obesity and Cardiovascular Diseases in Four Southern States and Colorado.” Journal of Health Care for the Poor and Underserved 22 (S4): 61–72.
Ali, Aus Tariq. 2014. “Reproductive Factors and the Risk of Endometrial Cancers.” International Journal of Gynecological Cancer 24 (3): 384–393.
American Heart Association. 2018. “Heart Disease and Stroke Statistics-2018 Update: A Report.” Circulation 137 (12). Accessed April 7, 2023. https://www.ahajournals.org/doi/10.1161/cir.0000000000000558.
American Psychiatric Association (APA). 2013. Diagnostic and Statistical Manual of Mental Disorder. 5th Edition: DSM-5. Washington, DC: APA.
Armelagos, George J. 1990. “Health and Disease in Prehistoric Populations in Transition.” Disease in Populations in Transition: Anthropological and Epidemiological Perspectives, edited by George J. Armelagos and Alan C. Swedland, 127–144. New York: Bergin & Garvey.
Armelagos, George J., Peter J. Brown, and Bethany Turner. 2005. “Evolutionary, Historical and Political Economic Perspectives on Health and Disease.” Social Science and Medicine 61 (4): 755-765.
Arnot, Megan, Eva Brandl, O. L. K. Campbell, Yuan Chen, Mark Dyble, Emily H. Emmott, et al. 2020. Evolution, Medicine, and Public Health 2020 (1): 264–278. https://doi.org/10.1093/emph/eoaa038.
Asher, Claire. 2017. “Illegal Bushmeat Trade Threatens Human Health and Great Apes.” Mongabay, April 6. Accessed April 4, 2023. https://news.mongabay.com/2017/04/illegal-bushmeat-trade-threatens-human-health-and-great-apes/.
Baker, W. A., G. A. Hitman, K. Hawrami, M .I. McCarthy, A. Riikonen, E. Tuomilehto-Wolf, A. Nissinen, et al. 1994. “Apolipoprotein D Gene Polymorphism: A New Genetic Marker for Type 2 Diabetic Subjects in Nauru and South India.” Diabetic Medicine 11 (10): 947–952.
Baltic, Milan Z., and Marija Boskovic. 2015. “When Man Met Meat: Meat in Human Nutrition from Ancient Times Till Today.” Procedia Food Science 5: 6- 9.
Blue Cross Blue Shield Association (BCBSA). 2017. “Diabetes and the Commercially Insured U.S. Population.” The Health of America Report, August 1. Accessed April 4, 2023. https://www.bcbs.com/the-health-of-america/reports/diabetes-and-commercially-insured-us-population.
Bogin, Barry. 1991. “The Evolution of Human Nutrition.” In The Anthropology of Medicine: From Culture to Method, edited by Lola Romanucci-Ross, Daniel E. Moerman, and Laurence R. Tancredi, 158–195. New York: Bergin & Garvey.
Bouchard, Claude. 2007. “The Biological Predisposition to Obesity: Beyond the Thrifty Genotype Scenario.” International Journal of Obesity 31 (9): 1337–1339.
Buchman, Aron S., Lei Yu, Robert S. Wilson, Andrew Lim, Robert J. Dawe, Chris Gaiteri, Sue E. Leurgans, Julie A. Schneider, and David A. Bennett. 2019. “Physical Activity, Common Brain Pathologies, and Cognition in Community-Dwelling Older Adults.” Neurology 98 (2). https://doi.org/10.1212/WNL.0000000000006954.
Centers for Disease Control and Prevention (CDC). N.d. “Global Health - Pakistan.” Accessed April 4, 2023. https://www.cdc.gov/globalhealth/countries/pakistan/default.htm.
Centers for Disease Control and Prevention (CDC). 2014. “Breastfeeding Report Card: United States/2014.” Atlanta, GA: Centers for Disease Control and Prevention.
Centers for Disease Control and Prevention (CDC). 2020. National Diabetes Statistics Report: Estimates of Diabetes and Its Burden in the United States. Accessed April 7, 2023 from https://diabetesresearch.org/wp-content/uploads/2022/05/national-diabetes-statistics-report-2020.pdf.
Centers for Disease Control and Prevention (CDC). 2022. “CDC Releases Updated Maps of America’s High Levels of Inactivity.” January 20. https://www.cdc.gov/media/releases/2022/p0120-inactivity-map.html.
Cerón, Ella. 2017. “Here’s Every Oreo Flavor Ever Created.” TeenVogue.com, June 19. Accessed April 4, 2023. https://www.teenvogue.com/story/every-oreo-flavor-ranked.
Chin, Jessie, Brennan R. Payne, Wai-Tat Fu, Daniel G. Morrow, and Elizabeth A. L. Stine-Morrow. 2015. “Information Foraging across the Life Span: Search and Switch in Unknown Patches.” Topics in Cognitive Science 7 (3): 428–450.
Cohn, D’Vera, and Jeffrey S. Passel. 2018. “A Record 64 Million Americans Live In Multigenerational Households.” Pew Research Center, April 5. Accessed April 4, 2023. https://www.pewresearch.org/fact-tank/2018/04/05/a-record-64-million-americans-live-in-multigenerational-households/.
Cooling, Laura. 2015. “Blood Groups in Infection and Host Susceptibility.” Clinical Microbiology Reviews 28 (3): 801–870.
Croston, Glenn. 2012. “The Thing We Fear More Than Death: Why Predators Are Responsible for Our Fear of Public Speaking.” Psychology Today blog, November 29. Accessed April 4, 2023. https://www.psychologytoday.com/us/blog/the-real-story-risk/201211/the-thing-we-fear-more-death.
Currie, Candace, Naman Ahluwalia, Emmanuelle Godeau, Saoirse Nic Gabhainn, Pernille Due, and Dorothy B. Mille. 2012. “Is Obesity at Individual and National Level Associated with Lower Age at Menarche? Evidence from 34 Countries in the Health Behaviour in School-Aged Children Study.” Journal of Adolescent Health 50 (6): 621–626.
David, A. Rosalie, and Michael Zimmerman. 2010. “Cancer: An Old Disease, A New Disease or Something In Between?” Nature Reviews: Cancer 10 (10): 728–733.
Davis, Mike. 2005. The Monster at Our Door: The Global Threat of Avian Flu. New York: Owl Books.
de Silva, A. M., K. R. Walder, T. J. Aitman, T. Gotoda, A. P. Goldstone, A. M. Hodge, M. P. de Courten, P. Z. Zimmet, and G. R. Collier. 1999. “Combination of Polymorphisms in OB-R and the OB Gene Associated with Insulin Resistance in Nauruan Males.” International Journal of Obesity 23 (8): 816–822.
Dimsdale, Joel E. 2008. “Psychological Stress and Cardiovascular Disease.” Journal of the American College of Cardiology 51 (13): 1237–1246.
Ding, Ding, Joe Van Buskirk, Binh Nguyen, Emmanuel Stammotakis, Mona Elbarbary, Nicole Veronese, Philip J. Clare, et al. 2022. “Physical Activity, Diet Quality and All-Cause Cardiovascular Disease and Cancer Mortality: A Prospective Study of 346,627 UK Biobank Participants.” British Journal of Sports Medicine 56 (20): 1148–1156.
Dutton, Denis. 2009. The Art Instinct: Beauty, Pleasure, and Human Evolution. New York: Bloomsbury.
Eaton, S. Boyd, Melvin Konner, and Marjorie Shostak. 1988. “Stone Agers in the Fast Lane: Chronic Degenerative Diseases in Evolutionary Perspective.” American Journal of Medicine 84 (4): 739–749.
Eaton, S. Boyd, Malcolm C. Pike, Roger V. Short, Nancy C. Lee, James Trussell, Robert A. Hatcher, James W. Wood, et al. 1994. “Women’s Reproductive Cancers in Evolutionary Context.” The Quarterly Review of Biology 69 (3): 353–367.
Ewald, Paul W. 2018. “Ancient Cancers and Infection-Induced Oncogenesis.” International Journal of Paleopathology 21: 178–185. https://dx.doi.org/10.1016/J.ijpp.2017.08.007.
Farb, Peter, and George Armelagos. 1980. Consuming Passions: The Anthropology of Eating. New York: Washington Square Press.
Fawzi, Ashraf, Tianshi David Wu, Kunbo Wang, Matthew L. Robinson, Jad Farha, Amanda Bradke, Sherita H. Golden, et al. 2022. “Racial and Ethnic Discrepancy in Pulse Oximetry and Delayed Identification of Treatment Eligibility among Patients With COVID-19.” JAMA Internal Medicine 182 (7): 730–738.
Flinn, Mark V., and Barry G. England. 2003. “Childhood Stress: Endocrine and Immune Responses to Psychosocial Events.” In Social and Cultural Lives of Immune Systems: Theory and Practice in Medical Anthropology and International Health, edited by James M. Wilce Jr., 105–146. London: Routledge.
Food Empowerment Project. N.d. “Food Deserts.” Accessed April 4, 2023. https://foodispower.org/access-health/food-deserts/.
Gao, George. 2015. “Americans’ Ideal Family Size Is Smaller Than It Used to Be.” Pew Research Center, May 8. Accessed April 4, 2023. https://www.pewresearch.org/fact-tank/2015/05/08/ideal-size-of-the-american-family/.
Ghosh, A. K., S. Venkatraman, O. Soroka, E. Reshetnyak, M. Rajan, A. An, J. K. Chae, et al. 2021. “Association between Overcrowded Households, Multigenerational Households, and COVID-19: A Cohort Study.” Public Health 198: 273–279.
Government of Pakistan. “Pakistan: Updated Nationally Determined Contributions.” 2021. Accessed October 12, 2022. https://unfccc.int/sites/default/files/NDC/2022-06/Pakistan%20Updated%20NDC%202021.pdf.
Greaves, Mel. 2015. “Evolutionary Determinants of Cancer.” Cancer Discovery 5 (8): 806–820.
Green, Carmen R., Karen O. Anderson, Tamara A. Baker, Lisa C. Campbell, Sheila Decker, Robert B. Fillingim, Donna A. Kalukalani, et al. 2003. “The Unequal Burden of Pain: Confronting Racial and Ethnic Disparities in Pain.” Pain Medicine 4 (3): 277–294.
Greenspan, Louise, and Julianna Deardorff. 2014. The New Puberty: How to Navigate Early Development in Today’s Girls. New York: Rodale.
Hamilton, Cheryl. 2011. Communicating for Results, a Guide for Business and the Professions, 9th Edition. Belmont, CA: Thomson Wadsworth.
Harper, Kristin, and George Armelagos. 2010. “The Changing Disease-Scape in the Third Epidemiological Transition.” International Journal of Environmental Research and Public Health 7 (2): 675–697.
Hoffman, Kelly M., Sophie Trawalter, Jordan R. Axt, and M. Norman Oliver. 2016. “Racial Bias in Pain Assessment and Treatment Recommendations, and False Beliefs about Biological Differences among Blacks and Whites.” PNAS 113 (16): 4296–4301.
International Agency for Research on Cancer (IARC). 2018. Red Meat and Processed Meat IARC Monographs on the Evaluation of Carcinogenic Risks to Humans, Volume 114. ISBN-13: 978-92-832-0152-6.
International Diabetes Federation (IDF). 2022. IDF Diabetes Atlas 2022 Reports. 10th edition. https://diabetesatlas.org/.
Jain, H. K. 2012. “Transition to Twenty-First Century Agriculture: Change of Direction.” Agricultural Research 1 (1): 12–17.
Jones, Kate E., Nikkita G. Patel, Mark A. Levy, Adam Storeygard, Deborah Balk, John L. Gittleman, and Peter Daszak. 2008. “Global Trends in Emerging Infectious Disease.” Nature 451 (7181): 990–993.
Karn, Mitesh, and Muna Sharma. 2021. “Climate Change, Natural Calamities, and the Triple Burden of Disease.” Nature Climate Change 11: 796–797.
Larsen, Clark Spencer. 2014. “Foraging to Farming Transition: Global Health Impacts, Trends, and Variation.” In Encyclopedia of Global Archaeology, edited by Claire Smith, 2818–2824. New York: Springer.
Lee, Richard B. 2013. The Dobe Ju/’hoansi. 4th edition. Belmont, CA: Wadsworth/Cengage Learning.
Lieberman, Daniel E. 2015. “Human Locomotion and Heat Loss: An Evolutionary Perspective.” Comprehensive Physiology 5 (1): 99–117.
Lieberman, Leslie Sue. 2006. “Evolutionary and Anthropological Perspectives on Optimal Foraging in Obesogenic Environments.” Appetite 47 (1): 3–9.
Liu, Gang. 2018. “Abstract P184: Meat Cooking Methods and Risk of Hypertension: Results From Three Prospective Cohort Studies.” Circulation 137 (S1): AP184.
Liu, Gang, Geng Zong, Kana Wu, Yang Hu, Yanping Li, Walter C. Willett, David M. Eisenberg, Frank B. Hu, and Qi Sun. 2018. “Meat Cooking Methods and Risk of Type 2 Diabetes: Results from Three Prospective Cohort Studies.” Diabetes Care 41 (5): 1049–1060.
Livingston, Gretchen. 2015. “Childlessness.” Pew Research Center, May 7. Accessed April 4, 2023. https://www.pewsocialtrends.org/2015/05/07/childlessness/.
Lucock, Mark D., Charlotte E. Martin, Zoe R. Yates, and Martin Veysey. 2014. “Diet and Our Genetic Legacy in the Recent Anthropocene: A Darwinian Perspective to Nutritional Health.” Journal of Evidence-Based Complementary and Alternative Medicine 19 (1): 68–83.
Mandsager, Kyle, Serge Harb, and Paul Cremer. 2018. “Association of Cardiorespiratory Fitness with Long-term Mortality among Adults Undergoing Exercise Treadmill Testing.” JAMA Network Open 1 (6): e183605. https://doi.org/10.1001/jamanetworkopen.2018.3605.
Marciniak, Stephanie, and George H. Perry. 2017. “Harnessing Ancient Genomes to Study the History of Human Adaptation.” Nature Reviews Genetics 18: 659–674.
Marlowe, Frank W. 2005. “Hunter-Gatherers and Human Evolution.” Evolutionary Anthropology 14 (2): 54–67.
Marmot, Michael. 2005. “Social Determinants of Health Inequality.” The Lancet 365 (9464): 1099–1104.
Martini, Frederic H., William C. Ober, Edwin F. Bartholomew, and Judi L. Nath. 2013. Visual Essentials of Anatomy & Physiology. Boston, MA: Pearson.
Mathews, T. J., and Brady E. Hamilton. 2016. “Mean Age of Mothers Is on the Rise: United States, 2000–2014.” National Center for Health Statistics (CHS) Data Brief. No. 232. Accessed April 4, 2023. https://www.cdc.gov/nchs/data/databriefs/db232.pdf.
McClellan, James E., and Harold Dorn. 2006. Science and Technology in World History: An Introduction. 2nd edition. Baltimore, MD: The Johns Hopkins University Press.
McElroy, Ann, and Patricia Townsend. 2009. Medical Anthropology in Ecological Perspective. 5th edition. Boulder, CO: Westview Press.
McKee, Trudy, and James R. McKee. 2015. Biochemistry: The Molecular Basis of Life. 6th edition. Oxford, UK: Oxford University Press.
McKinsey and Company. 2021. “Race in the Workplace: Black Workers in the U.S. Private Sector.” February 21. Accessed April 4, 2023. https://www.mckinsey.com/featured-insights/diversity-and-inclusion/race-in-the-workplace-the-black-experience-in-the-us-private-sector.
Moss, Michael. 2013. Salt, Sugar, Fat: How the Food Giants Hooked Us. New York: Random House.
National Institutes of Health (NIH). 2022. “COVID-19 Was Third Leading Cause of Death in the United States in Both 2020 and 2021.” Media Advisory. July 5. Accessed April 4, 2023. https://www.nih.gov/news-events/news-releases/covid-19-was-third-leading-cause-death-united-states-both-2020-2021.
National Institutes of Health (NIH). 2017. “Obesity and Cancer Fact Sheet.” Last modified January, 2017. Accessed April 4, 2023. https://www.cancer.gov/about-cancer/causes-prevention/risk/obesity/obesity-fact-sheet.
Odes, Edward J., Patrick S. Randolph-Quinney, Maryna Steyn, Zach Throckmorton, Jacqueline S. Smilg, Bernhard Zipfel, Tanya N. Augustine, et al. 2016. “Earliest Hominin Cancer: 1.7-Million-Year-Old Osteosarcoma from Swartkrans Cave, South Africa.” South African Journal of Science 112 (7–8): 1–5.
Omran, Abdel R. 2005. “The Epidemiologic Transition: A Theory of the Epidemiology of Population Change.” The Milbank Quarterly, December 83 (4): 731-757. Accessed April 7, 2023. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2690264/.
Palma-Gudiel, H., A. Córdova-Palomera, E. Eixarch, M. Deuschle, and L. Fañanás. 2015. “Maternal Psychosocial Stress during Pregnancy Alters the Epigenetic Signature of the Glucocorticoid Receptor Gene Promoter in Their Offspring: A Meta-Analysis.” Epigenetics 10 (10): 893–902.
Park, Sung S., Emily E. Wiemers, and Judith A. Seltzer. 2019. “The Family Safety Net of Black and White Multigenerational Families.” Population and Development Review 45 (2): 351–378.
Pierce, Mary, and Rebecca Hardy. 2012. “Commentary: The Decreasing Age of Puberty—As Much a Psychosocial as Biological Problem?” International Journal of Epidemiology 41 (1): 300–302.
Plotnikoff, Ronald C., Sarah A. Costigan, Rebecca L. Williams, Melinda J. Hutchesson, Sarah G. Kennedy, Sara L. Robards, Jennifer Allen, Clare E. Collins, Robin Callister, and John Germov. 2015. “Effectiveness of Interventions Targeting Physical Activity, Nutrition and Healthy Weight for University and College Students: A Systematic Review and Meta-analysis.” International Journal of Behavioral Nutrition and Physical Activity 12 (1): 1–10.
Poinar, George. 2018. “Vertebrate Pathogens Vectored by Ancient Hematophagous Arthropods.” Historical Biology, November 7. https://doi.org/10.1080/08912963.2018.1545018.
Pontzer, Herman, David A. Raichlen, Brian M. Wood, Audax Z. P. Mabulla, Susan B. Racette, and Frank W. Marlowe. 2012. “Hunter-Gatherer Energetics and Obesity.” PLoS ONE 7 (7): e40503. https://doi.org/10.1371/journal.pone.0040503.
Pritchard, Jonathan K. 2010. “How We Are Evolving.” Scientific American 303 (4): 4047.
Raichlen, David A., and Gene E. Alexander. 2014. “Exercise, APOE Genotype, and the Evolution of the Human Lifespan.” Trends in Neurosciences 37 (5): 247–255.
Raichlen, David A., Herman Pontzer, Jacob A. Harris, Audax Z. P. Mabulla, Frank W. Marlowe, J. Josh Snodgrass, Geeta Eick, J. Colette Berbesque, Amelia Sancilio, and Brian M. Wood. 2016. “Physical Activity Patterns and Biomarkers of Cardiovascular Disease Risk in Hunter-Gatherers.” American Journal of Human Biology 29 (2): e22919. https://doi.org/10.1002/ajhb.22919.
Raifman, Julia, Alexander Skinner, and Aaron Sojourner. 2022. “The Unequal Toll of COVID-19 On Workers.” Working Economics Blog (of the Economic Policy Institute), February 7. Accessed April 4, 2023. https://www.epi.org/blog/the-unequal-toll-of-covid-19-on-workers/.
Randolph-Quinney, Patrick S., Scott A. Williams, Maryna Stein, Mark R. Meyer, Jacqueline S. Smilg, Steven E. Churchill, Edward J. Odes, Tanya Augustine, Paul Tafforeau and Lee Berger. 2016. “Osteogenic Tumour in Australopithecus sediba: Earliest Hominin Evidence for Neoplastic Disease.” South African Journal of Science 112 (7-8). Accessed April 7, 2023. https://www.scielo.org.za/scielo.php?script=sci_arttext&pid=S0038-23532016000400013. https://dx.doi.org/10.17159/sajs.2016/20150470
Rankinen, Tuomo, Aamir Zuberi, Yvon C. Chagnon, S. John Weisnagel, George Argyropoulos, Brandon Walts, Louis Pérusse, and Claude Bouchard. 2006. “The Human Obesity Gene Map: The 2005 Update.” Obesity 14 (4): 529–644.
Richards, M. P. 2002. “A Brief Review of the Archaeological Evidence for Palaeolithic and Neolithic Subsistence.” European Journal of Clinical Nutrition 56 (12): 1270–1278.
Ross, Joseph S., Susannah M. Bernheim, Elizabeth H. Bradley, Hsun-Mei Teng, and William T. Gallo. 2007. “Use of Preventive Care by the Working Poor in the United States.” Preventive Medicine 44 (3): 254–259.
Sallis, James F., Neville Owen, and Edwin B. Fisher. 2008. “Chapter 20: Ecological Models of Health Behavior.” In Health Behavior and Health Education: Theory, Research, and Practice, edited by Karen Glanz, Barbara K. Rimer, and K. Viswanath, 465–485. 4th edition. San Francisco, CA: Jossey-Bass.
Sharp, Paul M., and Beatrice H. Hahn. 2011. “Origins of HIV and the AIDS Pandemic.” Cold Springs Harbor Perspectives in Medicine 1 (1): a006841.
Sheerazi, Hadia A. 2022. “The Flood Seen from Space: Pakistan’s Apocalyptic Crisis.” State of the Planet (news from the Columbia Climate School), September 12. Accessed April 4, 2023. https://news.climate.columbia.edu/2022/09/12/the-flood-seen-from-space-pakistans-apocalyptic-crisis/.
Shiels, Meredith S., Anika T. Haque, Amy Berrington de González, and Neal D. Freedman. 2022. “Leading Causes of Death in the US During the COVID-19 Pandemic, March 2020 to October 2021.” JAMA 182 (8): 883-886. doi:10.1001/jamainternmed.2022.2476.
Shonkoff, Jack P., and Andrew S. Garner. 2012. “The Lifelong Effects of Early Childhood Adversity and Toxic Stress.” Pediatrics 129 (1): e232–e246. https://doi.org/10.1542/peds.2011-2663.
Singer, Merrill, and Hans Baer. 2012. “Health Disparity, Health Inequality.” In Introducing Medical Anthropology: A Discipline in Action, 2nd edition, edited by Merrill Singer and Hans Baer, 175–205. Lanham, MD: AltaMira.
Singer, Merrill, Nicola Bulled, Bayla Ostrach, and Emily Mendenhall. 2017. “Syndemics and the Biosocial Conception of Health.” Lancet 389 (10072): 941–950.
Small, Meg, Lisa Bailey-Davis, Nicole Morgan, and Jennifer Maggs. 2013. “Changes in Eating and Physical Activity Behaviors across Seven Semesters of College: Living On or Off Campus Matters.” Health Education and Behavior 40 (4): 435–441.
Smith-Morris, Carolyn M. 2004. “Reducing Diabetes in Indian Country: Lessons from the Three Domains Influencing Pima Diabetes.” Human Organization 63 (1): 34–46.
Sonfield, Adam, Kinsey Hasstedt, Megan L. Cavanaugh, and Ragnar Anderson. 2013. “The Social and Economic Benefits of Women’s Ability to Determine Whether and When to Have Children.” Report, March 13. New York: Guttmacher Institute. Accessed April 4, 2023. https://www.guttmacher.org/report/social-and-economic-benefits-womens-ability-determine-whether-and-when-have-children.
St. Pierre, Danielle. 2018. “The 15 Best Potato Chips for Every Flavor Craving.” Best, April 6. Accessed June 13, 2018. https://www.bestproducts.com/eats/food/g972/best-potato-chips/.
Stearns, Stephen C., Randolph M. Nesse, and David Haig. 2008. “Introducing Evolutionary Thinking into Medicine.” In Evolution in Health and Disease, edited by Stephen C. Stearns and Jacob C. Koella, 3–15. United Kingdom: Oxford University Press.
Strassmann, Beverly I. 1997. “The Biology of Menstruation in Homo Sapiens: Total Lifetime Menses, Fecundity, and Nonsynchrony in a Natural-Fertility Population.” Current Anthropology 38 (1): 123–129.
Strassmann, Beverly I. 1999. “Menstrual Cycling and Breast Cancer: An Evolutionary Perspective.” Journal of Women’s Health 8 (2): 193–202.
Trafialek, Joanna, and Wojciech Kolanowski. 2014. “Dietary Exposure to Meat-Related Carcinogenic Substances: Is There a Way to Estimate the Risk?” International Journal of Food Sciences and Nutrition 65 (6): 774–780.
Troeger, Christopher. 2023. “Just How Do Deaths Due to COVID-19 Stack Up?” Think Global Health. February 15. Accessed April 8, 2023. https://www.thinkglobalhealth.org/article/just-how-do-deaths-due-covid-19-stack#:~:text=Looking%20at%20official%20statistics%20alone,since%20the%20beginning%20of%202020.
Valbuena, Valeria S.M., Raina M. Merchant, and Catherine L. Hough. 2022. “Racial and Ethnic Bias in Pulse Oximetry and Clinical Outcomes.” Editorial. JAMA Internal Medicine 182 (7): 699–700.
Ventola, C. Lee. 2015. “The Antibiotic Resistance Crisis: Part I: Causes and Threats.” Pharmacy & Therapeutics 40 (4): 277–283.
Versini, Mathilde, Pierre-Yves Jeandel, Tomer Bashi, Giorgia Bizzaro, Miri Blank, and Yahuda Shoenfeld. 2015. “Unraveling the Hygiene Hypothesis of Helminthes and Autoimmunity: Origins, Pathophysiology, and Clinical Applications.” BMC Medicine, 13: 81. https://doi.org/10.1186/s12916-015-0306-7.
Vucenik, Ivana, and Joseph P. Stains. 2012. “Obesity and Cancer Risk: Evidence, Mechanisms, and Recommendations.” Special issue, “Nutrition and Physical Activity in Aging, Obesity, and Cancer,” Annals of the New York Academy of Sciences 1271 (1): 37–43.
Walker, Polly, Pamela Rhubart-Berg, Shawn McKenzie, Kristin Kelling, and Robert S. Lawrence. 2005. “Public Health Implications of Meat Production and Consumption.” Public Health Nutrition 8 (4): 348–356.
Washington, Harriet A. 2006. Medical Apartheid: The Dark History of Medical Experimentation on Black Americans from Colonial Times to the Present. New York: Anchor Books.
Williams, David R., and Selina A. Mohammed. 2013. “Racism and Health I: Pathways and Scientific Evidence.” American Behavioral Scientist 57 (8). https://doi.org/10.1177/0002764213487340.
Williams, David R., and Ronald Wyatt. 2015. “Racial Bias in Health Care and Health: Challenges and Opportunities.” JAMA 314 (6): 555–556. https://doi.org10.1001/jama.2015.9260.
Wolfe, Nathan, Claire P. Dunavan, and Jared Diamond. 2012. “Origins Of Major Human Infectious Diseases.” In Institute of Medicine: Improving Food Safety through a One Health Approach: Workshop Summary, A16. Washington, DC: National Academies Press. Accessed April 4, 2023. https://www.ncbi.nlm.nih.gov/books/NBK114494/.
World Health Organization (WHO). 2016. Global Report on Diabetes. Accessed April 4, 2023. https://apps.who.int/iris/bitstream/handle/10665/204871/9789241565257_eng.pdf.
World Health Organization (WHO). 2017. “Obesity and Overweight.” Fact Sheet. Last modified June 9, 2021; accessed April 4, 2023. https://www.who.int/mediacentre/factsheets/fs311/en/.
World Health Organization (WHO). 2018. “Cancer.” Fact Sheet. Last modified February 3, 2022; accessed April 5, 2023. https://www.who.int/news-room/fact-sheets/detail/cancer.
Worobey, Michael, Joshua I. Levy, Lorena Malpica Serrano, Alexander Crits-Christoph, Jonathan E. Pekar, Stephan A. Goldstein, Angela L. Rassmussen, et al. July 26, 2022. “The Huanan Seafood Wholesale Market in Wuhan was the Early Epicenter of the COVID-19 Pandemic.” SCIENCE 26 (377): 951–959. https://doi.org/10.1126/science.abp8715.
Worthman, Carol M., and Jennifer Kuzara. 2005. “Life History and the Early Origins of Health Differentials.” American Journal of Human Biology 17 (1): 95–112.
Wrangham, Richard. 2009. Catching Fire: How Cooking Made Us Human. New York: Basic Books.
Yehuda, Rachel, and Joseph LeDoux. 2007. “Response Variation Following Trauma: A Translational Neuroscience Approach to Understanding PTSD.” Neuron 56 (1): 19–32.
Zuckerman, Molly K., and George J. Armelagos. 2014. “The Hygiene Hypothesis and the Second Epidemiologic Transition.” In Modern Environments and Human Health: Revisiting the Second Epidemiologic Transition, edited by Molly K. Zuckerman, 301–320. Hoboken, NJ: Wiley-Blackwell.
Zuckerman, Molly Kathleen, Kristin Nicole Harper, Ronald Barrett, and George John Armelagos. 2014. “The Evolution of Disease: Anthropological Perspectives on Epidemiologic Transitions.” Special issue, “Epidemiological Transitions: Beyond Omran’s Theory,” Global Health Action 7 (1): 23303. https://doi.org/10.3402/gha.v7.23303.
Andrea J. Alveshere, Ph.D., Western Illinois University
This chapter is a revision from "Chapter 4: Forces of Evolution” by Andrea J. Alveshere. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Learning Objectives
- Outline a 21st-century perspective of the Modern Synthesis.
- Define populations and population genetics as well as the methods used to study them.
- Identify the forces of evolution and become familiar with examples of each.
- Discuss the evolutionary significance of mutation, genetic drift, gene flow, and natural selection.
- Explain how allele frequencies can be used to study evolution as it happens.
- Contrast micro- and macroevolution.
It’s hard for us, with our typical human life spans of less than 100 years, to imagine all the way back, 3.8 billion years ago, to the origins of life. Scientists still study and debate how life came into being and whether it originated on Earth or in some other region of the universe (including some scientists who believe that studying evolution can reveal the complex processes that were set in motion by God or a higher power). What we do know is that a living single-celled organism was present on Earth during the early stages of our planet’s existence. This organism had the potential to reproduce by making copies of itself, just like bacteria, many amoebae, and our own living cells today. In fact, with modern technologies, we can now trace genetic lineages, or phylogenies, and determine the relationships between all of today’s living organisms—eukaryotes (animals, plants, fungi, etc.), archaea, and bacteria—on the branches of the phylogenetic tree of life (Figure 4.1).
Looking at the common sequences in modern genomes, we can even make educated guesses about the likely genetic sequence of the Last Universal Common Ancestor (LUCA) of all living things. Through a wondrous series of mechanisms and events over nearly four billion years, that ancient single-celled organism gave rise to the rich diversity of species that fill the lands, seas, and skies of our planet. This chapter explores the mechanisms by which that amazing transformation occurred and considers some of the crucial scientific experiments that shaped our current understanding of the evolutionary process.
Updating the Modern Synthesis: Tying it All Together
Chapter 2 examined the roles played by many different scientists, and their many careful scientific experiments, in providing the full picture of evolution. The term Modern Synthesis describes the integration of these various lines of evidence into a unified theory of evolution.
While the biggest leap forward in understanding how evolution works came with the joining (synthesis) of Darwin’s concept of natural selection with Mendel’s insights about particulate inheritance (described in detail in Chapter 3), there were some other big contributions that were crucial to making sense of the variation that was being observed. Mathematical models for evolutionary change provided the tools to study variation and became the basis for the study of population genetics (Fisher 1919; Haldane 1924). Other experiments revealed the existence of chromosomes as carriers of collections of genes (Dobzhansky 1937; Wright 1932).
Studies on wild butterflies confirmed the mathematical predictions and also led to the definition of the concept of polymorphisms to describe multiple forms of a trait (Ford 1949). These studies led to many useful advances such as the discovery that human blood type polymorphisms are maintained in the human population because they are involved in disease resistance (Ford 1942; see also the Special Topic box on Sickle Cell Anemia below).
Population Genetics
Defining Populations and the Variations within Them
One of the major breakthroughs in understanding the mechanisms of evolutionary change came with the realization that evolution takes place at the level of populations, not within individuals. In the biological sciences, a population is defined as a group of individuals of the same species who are geographically near enough to one another that they can breed and produce new generations of individuals.
For the purpose of studying evolution, we recognize populations by their even smaller units: genes. Remember, a gene is the basic unit of information that encodes the proteins needed to grow and function as a living organism. Each gene can have multiple alleles, or variants—each of which may produce a slightly different protein. Each individual, for genetic inheritance purposes, carries a collection of genes that can be passed down to future generations. For this reason, in population genetics, we think of populations as gene pools, which refers to the entire collection of genetic material in a breeding community that can be passed on from one generation to the next.
For genes carried on our human chromosomes (our nuclear DNA), we inherit two copies of each, one from each parent. This means we may carry two of the same alleles (a homozygous genotype) or two different alleles (a heterozygous genotype) for each nuclear gene.
Defining Evolution
In order to understand evolution, it’s crucial to remember that evolution is always studied at the population level. Also, if a population were to stay exactly the same from one generation to the next, it would not be evolving. So evolution requires both a population of breeding individuals and some kind of a genetic change occurring within it. Thus, the simple definition of evolution is a change in the allele frequencies in a population over time. What do we mean by allele frequencies? Allele frequencies refer to the ratio, or percentage, of one allele (one variant of a gene) compared to the other alleles for that gene within the study population (Figure 4.2). By contrast, genotype frequencies are the ratios or percentages of the different homozygous and heterozygous genotypes in the population. Because we carry two alleles per genotype, the total count of alleles in a population will usually be exactly double the total count of genotypes in the same population (with the exception being rare cases in which an individual carries a different number of chromosomes than the typical two; e.g., Down syndrome results when a child carries three copies of Chromosome 21).
The Forces of Evolution
Today, we recognize that evolution takes place through a combination of mechanisms: mutation, genetic drift, gene flow, and natural selection. These mechanisms are called the “forces of evolution”; together they account for all the genotypic variation observed in the world today. Keep in mind that each of these forces was first defined and then tested—and retested—through the experimental work of the many scientists who contributed to the Modern Synthesis.
Mutation
The first force of evolution we will discuss is mutation, and for good reason: mutation is the original source of all the genetic variation found in every living thing. Imagine all the way back in time to the very first single-celled organism, floating in Earth’s primordial sea. Based on what we observe in simple, single-celled organisms today, that organism probably spent its lifetime absorbing nutrients and dividing to produce cloned copies of itself. While the numbers of individuals in that population would have grown (as long as the environment was favorable), nothing would have changed in that perfectly cloned population. There would not have been variety among the individuals. It was only through a copying error—the introduction of a mutation, or change, into the genetic code—that new alleles were introduced into the population.
After many generations have passed in our primordial population, mutations have created distinct chromosomes. The cells are now amoeba-like, larger than many of their tiny bacterial neighbors, who have long since become their favorite source of nutrients. Without mutation to create this diversity, all living things would still be identical to LUCA, our universal ancestor (Figure 4.3).
When we think of genetic mutation, we often first think of deleterious mutations—the ones associated with negative effects such as the beginnings of cancers or heritable disorders. The fact is, though, that every genetic adaptation that has helped our ancestors survive since the dawn of life is directly due to beneficial mutations—changes in the DNA that provided some sort of advantage to a given population at a particular moment in time. For example, a beneficial mutation allowed chihuahuas and other tropical-adapted dog breeds to have much thinner fur coats than their cold-adapted cousins the northern wolves, malamutes, and huskies.
Every one of us has genetic mutations. Yes, even you. The DNA in some of your cells today differs from the original DNA that you inherited when you were a tiny, fertilized egg. Mutations occur all the time in the cells of our skin and other organs, due to chemical changes in the nucleotides. Exposure to the UV radiation in sunlight is one common cause of skin mutations. Interaction with UV light causes UV crosslinking, in which adjacent thymine bases bind with one another (Figure 4.4). Many of these mutations are detected and corrected by DNA repair mechanisms, enzymes that patrol and repair DNA in living cells, while other mutations may cause a new freckle or mole or, perhaps, an unusual hair to grow. For people with the autosomal recessive disease xeroderma pigmentosum, these repair mechanisms do not function correctly, resulting in a host of problems especially related to sun exposure, including severe sunburns, dry skin, heavy freckling, and other pigment changes.
Most of our mutations exist in somatic cells, which are the cells of our organs and other body tissues. Those will not be passed onto future generations and so will not affect the population over time. Only mutations that occur in the gametes, the reproductive cells (i.e., the sperm or egg cells), will be passed onto future generations. When a new mutation pops up at random in a family lineage, it is known as a spontaneous mutation. If the individual born with this spontaneous mutation passes it on to his offspring, those offspring receive an inherited mutation. Geneticists have identified many classes of mutations and the causes and effects of many of these.
Point Mutations
A point mutation is a single-letter (single-nucleotide) change in the genetic code resulting in the substitution of one nucleic acid base for a different one. As you learned in Chapter 3, the DNA code in each gene is translated through three-letter “words” known as codons. So depending on how the point mutation changes the “word,” the effect it will have on the protein may be major or minor or may make no difference at all.
If a mutation does not change the resulting protein, then it is called a synonymous mutation. Synonymous mutations do involve a letter (nucleic acid) change, but that change results in a codon that codes for the same “instruction” (the same amino acid or stop code) as the original codon. Mutations that do cause a change in the protein are known as nonsynonymous mutations. Nonsynonymous mutations may change the resulting protein’s amino acid sequence by altering the DNA sequence that encodes the mRNA or by changing how the mRNA is spliced prior to translation (refer to Chapter 3 for more details).
Insertions and Deletions
In addition to point mutations, another class of mutations are insertions and deletions, or indels, for short. As the name suggests, these involve the addition (insertion) or removal (deletion) of one or more coding sequence letters (nucleic acids). These typically first occur as an error in DNA replication, wherein one or more nucleotides are either duplicated or skipped in error. Entire codons or sets of codons may also be removed or added if the indel is a multiple of three nucleotides.
Frameshift mutations are types of indels that involve the insertion or deletion of any number of nucleotides that is not a multiple of three (e.g., adding one or two extra letters to the code). Because these indels are not consistent with the codon numbering, they “shift the reading frame,” causing all the codons beyond the mutation to be misread. Like point mutations, small indels can also disrupt splice sites.
Transposable elements, or transposons, are fragments of DNA that can “jump” around in the genome. There are two types of transposons: retrotransposons are transcribed from DNA into RNA and then “reverse transcribed,” to insert the copied sequence into a new location in the DNA, and DNA transposons, which do not involve RNA. DNA transposons are clipped out of the DNA sequence itself and inserted elsewhere in the genome. Because transposable elements insert themselves into existing DNA sequences, they are frequent gene disruptors. At certain times, and in certain species, it appears that transposons became very active, likely accelerating the mutation rate (and thus, the genetic variation) in those populations during the active periods.
Chromosomal Alterations
The final major category of genetic mutations are changes at the chromosome level: crossover events, nondisjunction events, and translocations. Crossover events occur when DNA is swapped between homologous chromosomes while they are paired up during meiosis I. Crossovers are thought to be so common that some DNA swapping may happen every time chromosomes go through meiosis I. Crossovers don’t necessarily introduce new alleles into a population, but they do make it possible for new combinations of alleles to exist on a single chromosome that can be passed to future generations. This also enables new combinations of alleles to be found within siblings who share the same parents. Also, if the fragments that cross over don’t break at exactly the same point, they can cause genes to be deleted from one of the homologous chromosomes and duplicated on the other.
Nondisjunction events occur when the homologous chromosomes (in meiosis I) or sister chromatids (in meiosis II and mitosis) fail to separate after pairing. The result is that both chromosomes or chromatids end up in the same daughter cell, leaving the other daughter cell without any copy of that chromosome (Figure 4.5). Most nondisjunctions at the gamete level are fatal to the embryo. The most widely known exception is Trisomy 21, or Down syndrome, which results when an embryo inherits three copies of Chromosome 21: two from one parent (due to a nondisjunction event) and one from the other (Figure 4.6). Trisomies (triple chromosome conditions) of Chromosomes 18 (Edwards syndrome) and 13 (Patau syndrome) are also known to result in live births, but the children usually have severe complications and rarely survive beyond the first year of life.
Sex chromosome trisomies (XXX, XXY, XYY) and X chromosome monosomies (inheritance of an X chromosome from one parent and no sex chromosome from the other) are also survivable and fairly common. The symptoms vary but often include atypical sexual characteristics, either at birth or at puberty, and often result in sterility. The X chromosome carries unique genes that are required for survival; therefore, Y chromosome monosomies are incompatible with life.
Chromosomal translocations involve transfers of DNA between nonhomologous chromosomes. This may involve swapping large portions of two or more chromosomes. The exchanges of DNA may be balanced or unbalanced. In balanced translocations, the genes are swapped, but no genetic information is lost. In unbalanced translocations, there is an unequal exchange of genetic material, resulting in duplication or loss of genes. Translocations result in new chromosomal structures called derivative chromosomes, because they are derived or created from two different chromosomes. Translocations are often found to be linked to cancers and can also cause infertility. Even if the translocations are balanced in the parent, the embryo often won’t survive unless the baby inherits both of that parent’s derivative chromosomes (to maintain the balance).
Genetic Drift
The second force of evolution is commonly known as genetic drift. This is an unfortunate misnomer, as this force actually involves the drifting of alleles, not genes. Genetic drift refers to random changes (“drift”) in allele frequencies from one generation to the next. The genes are remaining constant within the population; it is only the alleles of the genes that are changing in frequency. The random nature of genetic drift is a crucial point to understand: it specifically occurs when none of the variant alleles confer an advantage.
Let’s imagine far back in time, again, to that ancient population of amoeba-like cells, subsisting and occasionally dividing, in the primordial sea. A mutation occurs in one of the cells that changes the texture of the cell membrane from a relatively smooth surface to a highly ruffled one (Figure 4.7). This has absolutely no effect on the cell’s quality of life or ability to reproduce. In fact, eyes haven’t evolved yet, so no one in the world at the time would even notice the difference. The cells in the population continue to divide, and the offspring of the ruffled cell inherit the ruffled membrane. The frequency (percentage) of the ruffled allele in the population, from one generation to the next, will depend entirely on how many offspring that first ruffled cell ends up having, and the random events that might make the ruffled alleles more common or more rare (such as population bottlenecks and founder effects, which are discussed below).
Sexual Reproduction and Random Inheritance
Tracking alleles gets a bit more complicated in our primordial cells when, after a number of generations, a series of mutations have created populations that reproduce sexually. These cells now must go through an extra round of cell division (meiosis) to create haploid gametes. The combination of two gametes is now required to produce each new diploid offspring.
In the earlier population, which reproduced via asexual reproduction, a cell either carried the smooth allele or the ruffled allele. With sexual reproduction, a cell inherits one allele from each parent, so there are homozygous cells that contain two smooth alleles, homozygous cells that contain two ruffled alleles, and heterozygous cells that contain one of each allele (Figure 4.8). If the new, ruffled allele happens to be dominant (and we’ll imagine that it is), the heterozygotes will have ruffled cell phenotypes but also will have a 50/50 chance of passing on a smooth allele to each offspring. As long as neither phenotype (ruffled nor smooth) provides any advantage over the other, the variation in the population from one generation to the next will remain completely random.
In sexually reproducing populations (including humans and many other animals and plants in the world today), that 50/50 chance of inheriting one or the other allele from each parent plays a major role in the random nature of genetic drift.
Population Bottlenecks
A population bottleneck occurs when the number of individuals in a population drops dramatically due to some random event. The most obvious, familiar examples are natural disasters. Tsunamis and hurricanes devastating island and coastal populations and forest fires and river floods wiping out populations in other areas are all too familiar. When a large portion of a population is randomly wiped out, the allele frequencies (i.e., the percentages of each allele) in the small population of survivors are often much different from the frequencies in the predisaster, or “parent,” population.
If such an event happened to our primordial ocean cell population—perhaps a volcanic fissure erupted in the ocean floor and only the cells that happened to be farthest from the spewing lava and boiling water survived—we might end up, by random chance, with a surviving population that had mostly ruffled alleles, in contrast to the parent population, which had only a small percentage of ruffles (Figure 4.9).
One of the most famous examples of a population bottleneck is the prehistoric disaster that led to the extinction of dinosaurs, the Cretaceous–Paleogene extinction event (often abbreviated K–Pg; previously K-T). This occurred approximately 66 million years ago. Dinosaurs and all their neighbors were going about their ordinary routines when a massive asteroid zoomed in from space and crashed into what is now the Gulf of Mexico, creating an impact so enormous that populations within hundreds of miles of the crash site were likely immediately wiped out. The skies filled with dust and debris, causing temperatures to plummet worldwide. It’s estimated that 75% of the world’s species went extinct as a result of the impact and the deep freeze that followed (Jablonski and Chaloner 1994).
The populations that emerged from the K-Pg extinction were markedly different from their predisaster communities. Surviving mammal populations expanded and diversified, and other new creatures appeared. The ecosystems of Earth were filled with new organisms and have never been the same (Figure 4.10).
Much more recently in geological time, during the colonial period, many human populations experienced bottlenecks as a result of the fact that imperial powers were inclined to slaughter communities who were reluctant to give up their lands and resources. This effect was especially profound in the Americas, where Indigenous populations faced the compounded effects of brutal warfare, exposure to new bacteria and viruses (against which they had no immunity), and ultimately segregation on resource-starved reservations. The populations in Europe, Asia, and Africa had experienced regular gene flow during the 10,000-year period in which most kinds of livestock were being domesticated, giving them many generations of experience building up immunity against zoonotic diseases (those that can pass from animals to humans). In contrast, the residents of the Americas had been almost completely isolated during those millennia, so all these diseases swept through the Americas in rapid succession, creating a major loss of genetic diversity in the Indigenous American population. It is estimated that between 50% and 95% of the Indigenous American populations died during the first decades after European contact, around 500 years ago (Livi-Bacci 2006).
An urgent health challenge facing humans today involves human-induced population bottlenecks that produce antibiotic-resistant bacteria. Antibiotics are medicines prescribed to treat bacterial infections. The typical prescription includes enough medicine for ten days. People often feel better much sooner than ten days and sometimes decide to quit taking the medicine ahead of schedule. This is often a big mistake. The antibiotics have quickly killed off a large percentage of the bacteria—enough to reduce the symptoms and make you feel much better. However, this has created a bacterial population bottleneck. There are usually a small number of bacteria that survive those early days. If you take the medicine as prescribed for the full ten days, it’s quite likely that there will be no bacterial survivors. If you quit early, though, the survivors—who were the members of the original population who were most resistant to the antibiotic—will begin to reproduce again. Soon the infection will be back, possibly worse than before, and now all of the bacteria are resistant to the antibiotic that you had been prescribed.
Other activities that have contributed to the rise of antibiotic-resistant bacteria include the use of antibacterial cleaning products and the inappropriate use of antibiotics as a preventative measure in livestock or to treat infections that are viral instead of bacterial (viruses do not respond to antibiotics). In 2017, the World Health Organization published a list of twelve antibiotic-resistant pathogens that are considered top priority targets for the development of new antibiotics (World Health Organization 2017).
Founder Effects
Founder effects occur when members of a population leave the main or “parent” group and form a new population that no longer interbreeds with the other members of the original group. Similar to survivors of a population bottleneck, the newly founded population often has allele frequencies that are different from the original group. Alleles that may have been relatively rare in the parent population can end up being very common due to the founder effect. Likewise, recessive traits that were seldom seen in the parent population may be seen frequently in the descendants of the offshoot population.
One striking example of the founder effect was first noted in the Dominican Republic in the 1970s. During a several-year period, eighteen children who had been born with female genitalia and raised as girls suddenly grew penises at puberty. This culture tended to value sons over daughters, so these transitions were generally celebrated. They labeled the condition guevedoces, which translates to “penis at twelve,” due to the average age at which this occurred. Scientists were fascinated by the phenomenon.
Genetic and hormonal studies revealed that the condition, scientifically termed 5-alpha reductase deficiency, is an autosomal recessive syndrome that manifests when a child having both X and Y sex chromosomes inherits two nonfunctional (mutated) copies of the SRD5A2 gene (Imperato-McGinley and Zhu 2002). These children develop testes internally, but the 5-alpha reductase 2 steroid, which is necessary for development of male genitals in babies, is not produced. In absence of this male hormone, the baby develops female-looking genitalia (in humans, “female” is the default infant body form, if the full set of the necessary male hormones are not produced). At puberty, however, a different set of male hormones are produced by other fully functional genes. These hormones complete the male genital development that did not happen in infancy. This condition became quite common in the Dominican Republic during the 1970s due to founder effect—that is, the mutated SRD5A2 gene happened to be much more common among the Dominican Republic’s founding population than in the parent populations. (The Dominican population derives from a mixture of Indigenous Americans [Taino] peoples, West Africans, and Western Europeans.) Five-alpha reductase syndrome has since been observed in other small, isolated populations around the world.
Founder effect is closely linked to the concept of inbreeding, which in population genetics does not necessarily mean breeding with immediate family relatives. Instead, inbreeding refers to the selection of mates exclusively from within a small, closed population—that is, from a group with limited allelic variability. This can be observed in small, physically isolated populations but also can happen when cultural practices limit mates to a small group. As with the founder effect, inbreeding increases the risk of inheriting two copies of any nonfunctional (mutant) alleles.
The Amish in the United States are a population that, due to their unique history and cultural practices, emerged from a small founding population and have tended to select mates from within their groups. The Old Order Amish population of Lancaster County, Pennsylvania, has approximately 50,000 current members, all of whom can trace their ancestry back to a group of approximately 80 individuals. This small founding population immigrated to the United States from Switzerland in the mid-1700s to escape religious persecution. Since the Amish keep to themselves and almost exclusively select mates from within their own communities, they have more recessive traits compared to their parent population.
One of the genetic conditions that has been observed much more frequently in the Lancaster County Amish population is Ellis-van Creveld syndrome, which is an autosomal recessive disorder characterized by short stature (dwarfism), polydactyly (the development of more than five digits [fingers or toes] on the hands or feet], abnormal tooth development, and heart defects (Figure 4.11). Among the general world population, Ellis-van Creveld syndrome is estimated to affect approximately 1 in 60,000 individuals; among the Old Order Amish of Lancaster County, the rate is estimated to be as high as 1 in every 200 births (D’Asdia et al. 2013).
One important insight that has come from the study of founder effects is that a limited gene pool carries a much higher risk for genetic diseases. Genetic diversity in a population greatly reduces these risks.
Gene Flow
The third force of evolution is traditionally called gene flow. As with genetic drift, this is a misnomer, because it refers to flowing alleles, not genes. (All members of the same species share the same genes; it is the alleles of those genes that may vary.) Gene flow refers to the movement of alleles from one population to another. In most cases, gene flow can be considered synonymous with migration.
Returning again to the example of our primordial cell population, let’s imagine that, after the volcanic fissure opened up in the ocean floor, wiping out the majority of the parent population, two surviving populations developed in the waters on opposite sides of the fissure. Ultimately, the lava from the fissure cooled into a large island that continued to provide a physical barrier between the populations (Figure 4.12).
In the initial generations after the eruption, due to founder effect, isolation, and random inheritance (genetic drift), the population to the west of the islands contained a vast majority of the ruffled membrane alleles while the eastern population carried only the smooth alleles. Ocean currents in the area typically flowed from east to west, sometimes carrying cells (facilitating gene flow) from the eastern (smooth) population to the western (ruffled) population. Due to the ocean currents, it was almost impossible for any cells from the western population to be carried eastward. Thus, for inheritance purposes, the eastern (smooth) population remained isolated. In this case, the gene flow is unidirectional (going only in one direction) and unbalanced (only one population is receiving the new alleles).
Among humans, gene flow is often described as admixture. In forensic cases, anthropologists and geneticists are often asked to estimate the ancestry of unidentified human remains to help determine whether they match any missing persons’ reports. This is one of the most complicated tasks in these professions because, while “race” or “ancestry” involves simple checkboxes on a missing person’s form, among humans today there are no truly distinct genetic populations. All modern humans are members of the same fully breeding compatible species, and all human communities have experienced multiple episodes of gene flow (admixture), leading all humans today to be so genetically similar that we are all members of the same (and only surviving) human subspecies: Homo sapiens sapiens.
Gene flow between otherwise isolated nonhuman populations is often termed hybridization.. One example of this involves the hybridization and spread of Scutellata honey bees (a.k.a. “killer bees”) in the Americas. All honey bees worldwide are classified as Apis mellifera. Due to distinct adaptations to various environments around the world, there are 28 different subspecies of Apis mellifera.
During the 1950s, a Brazilian biologist named Warwick E. Kerr experimented with hybridizing African and European subspecies of honey bees to try to develop a strain that was better suited to tropical environments than the European honey bees that had long been kept by North American beekeepers. Dr. Kerr was careful to contain the reproductive queens and drones from the African subspecies, but in 1957, a visiting beekeeper accidentally released 26 queen bees of the Scutellata subspecies (Apis mellifera scutellata) from southern Africa into the Brazilian countryside. The Scutellata bees quickly interbred with local European honey bee populations. The hybridized bees exhibited a much more aggressively defensive behavior, fatally or near-fatally attacking many humans and livestock that ventured too close to their hives. The hybridized bees spread throughout South America and reached Mexico and California by 1985. By 1990, permanent colonies had been established in Texas, and by 1997, 90% of trapped bee swarms around Tucson, Arizona, were found to be Scutellata hybrids (Sanford 2006).
Another example involves the introduction of the Harlequin ladybeetle, Harmonia axyridis, native to East Asia, to other parts of the world as a “natural” form of pest control. Harlequin ladybeetles are natural predators of some of the aphids and other crop-pest insects. First introduced to North America in 1916, the “biocontrol” strains of Harlequin ladybeetles were considered to be quite successful in reducing crop pests and saving farmers substantial amounts of money. After many decades of successful use in North America, biocontrol strains of Harlequin ladybeetles were also developed in Europe and South America in the 1980s.
Over the seven decades of biocontrol use, the Harlequin ladybeetle had never shown any potential for development of wild colonies outside of its native habitat in China and Japan. New generations of beetles always had to be reared in the lab. That all changed in 1988, when a wild colony took root near New Orleans, Louisiana. Either through admixture with a native ladybeetle strain, or due to a spontaneous mutation, a new allele was clearly introduced into this population that suddenly enabled them to survive and reproduce in a wide range of environments. This population spread rapidly across the Americas and had reached Africa by 2004.
In Europe, the invasive, North American strain of Harlequin ladybeetle admixed with the European strain (Figure 4.13), causing a population explosion (Lombaert et al. 2010). Even strains specifically developed to be flightless (to curtail the spreading) produced flighted offspring after admixture with members of the North American population (Facon et al. 2011). The fast-spreading, invasive strain has quickly become a disaster, out-competing native ladybeetle populations (some to the point of extinction), causing home infestations, decimating fruit crops, and contaminating many batches of wine with their bitter flavor after being inadvertently harvested with the grapes (Pickering et al. 2004).
Natural Selection
The final force of evolution is natural selection. This is the evolutionary process that Charles Darwin first brought to light, and it is what the general public typically evokes when considering the process of evolution. Natural selection occurs when certain phenotypes confer an advantage or disadvantage in survival and/or reproductive success. The alleles associated with those phenotypes will change in frequency over time due to this selective pressure. It’s also important to note that the advantageous allele may change over time (with environmental changes) and that an allele that had previously been benign may become advantageous or detrimental. Of course, dominant, recessive, and codominant traits will be selected upon a bit differently from one another. Because natural selection acts on phenotypes rather than the alleles themselves, deleterious (disadvantageous) alleles can be retained by heterozygotes without any negative effects.
In the case of our primordial ocean cells, up until now, the texture of their cell membranes has been benign. The frequencies of smooth to ruffled alleles, and smooth to ruffled phenotypes, has changed over time, due to genetic drift and gene flow. Let’s now imagine that the Earth’s climate has cooled to a point that the waters frequently become too cold for survival of the tiny bacteria that are the dietary staples of our smooth and ruffled cell populations. The way amoeba-like cells “eat” is to stretch out the cell membrane, almost like an arm, to encapsulate, then ingest, the tiny bacteria. When the temperatures plummet, the tiny bacteria populations plummet with them. Larger bacteria, however, are better able to withstand the temperature change.
The smooth cells were well-adapted to ingesting tiny bacteria but poorly suited to encapsulating the larger bacteria. The cells with the ruffled membranes, however, are easily able to extend their ruffles to encapsulate the larger bacteria. They also find themselves able to stretch their entire membrane to a much larger size than their smooth-surfaced neighbors, allowing them to ingest more bacteria at a given time and to go for longer periods between feedings (Figure 4.14).
The smooth and ruffled traits, which had previously offered no advantage or disadvantage while food was plentiful, now are subject to natural selection. During the cold snaps, at least, the ruffled cells have a definite advantage. We can imagine that the western population that has mostly ruffled alleles will continue to do well, while the eastern population is at risk of dying out if the smaller bacteria remain scarce and no ruffled alleles are introduced.
A classic example of natural selection involves the study of an insect called the peppered moth (Biston betularia) in England during the Industrial Revolution in the 1800s. Prior to the Industrial Revolution, the peppered moth population was predominantly light in color, with dark (pepper-like) speckles on the wings. The “peppered” coloration was very similar to the appearance of the bark and lichens that grew on the local trees (Figure 4.15). This helped to camouflage the moths as they rested on a tree, making it harder for moth-eating birds to find and snack on them. There was another phenotype that popped up occasionally in the population. These individuals were heterozygotes that carried an overactive, dominant pigment allele, producing a solid black coloration. As you can imagine, the black moths were much easier for birds to spot, making this phenotype a real disadvantage.
The situation changed, however, as the Industrial Revolution took off. Large factories began spewing vast amounts of coal smoke into the air, blanketing the countryside, including the lichens and trees, in black soot. Suddenly, it was the light-colored moths that were easy for birds to spot and the black moths that held the advantage. The frequency of the dark pigment allele rose dramatically. By 1895, the black moth phenotype accounted for 98% of observed moths (Grant 1999).
Thanks to new environmental regulations in the 1960s, the air pollution in England began to taper off. As the soot levels decreased, returning the trees to their former, lighter color, this provided the perfect opportunity to study how the peppered moth population would respond. Repeated follow-up studies documented the gradual rise in the frequency of the lighter-colored phenotype. By 2003, the maximum frequency of the dark phenotype was 50% and in most parts of England had decreased to less than 10% (Cook 2003).
Directional, Balancing/Stabilizing, and Disruptive/Diversifying Selection
Natural selection can be classified as directional, balancing/stabilizing, or disruptive/diversifying, depending on how the pressure is applied to the population (Figure 4.16).
Both of the above examples of natural selection involve directional selection: the environmental pressures favor one phenotype over the other and cause the frequencies of the associated advantageous alleles (ruffled membranes, dark pigment) to gradually increase. In the case of the peppered moths, the direction shifted three times: first, it was selecting for lighter pigment; then, with the increase in pollution, the pressure switched to selection for darker pigment; finally, with reduction of the pollution, the selection pressure shifted back again to favoring light-colored moths.
Balancing selection (a.k.a. stabilizing selection) occurs when selection works against the extremes of a trait and favors the intermediate phenotype. For example, humans maintain an average birth weight that balances the need for babies to be small enough not to cause complications during pregnancy and childbirth but big enough to maintain a safe body temperature after they are born. Another example of balancing selection is found in the genetic disorder called sickle cell anemia (see “Special Topic: Sickle Cell Anemia”).
Disruptive selection (a.k.a. diversifying selection), the opposite of balancing selection, occurs when both extremes of a trait are advantageous. Since individuals with traits in the mid-range are selected against, disruptive selection can eventually lead to the population evolving into two separate species. Darwin believed that the many species of finches (small birds) found in the remote Galapagos Islands provided a clear example of disruptive selection leading to speciation. He observed that seed-eating finches either had large beaks, capable of eating very large seeds, or small beaks, capable of retrieving tiny seeds. The islands did not have many plants that produced medium-size seeds. Thus, birds with medium-size beaks would have trouble eating the very large seeds and would also have been inefficient at picking up the tiny seeds. Over time, Darwin surmised, this pressure against mid-size beaks may have led the population to divide into two separate species.
Sexual Selection
Sexual selection is an aspect of natural selection in which the selective pressure specifically affects reproductive success (the ability to successfully breed and raise offspring) rather than survival. Sexual selection favors traits that will attract a mate. Sometimes these sexually appealing traits even carry greater risks in terms of survival.
A classic example of sexual selection involves the brightly colored feathers of the peacock. The peacock is the male sex of the peafowl genera Pavo and Afropavo. During mating season, peacocks will fan their colorful tails wide and strut in front of the peahens in a grand display. The peahens will carefully observe these displays and will elect to mate with the male that they find the most appealing. Many studies have found that peahens prefer the males with the fullest, most colorful tails. While these large, showy tails provide a reproductive advantage, they can be a real burden in terms of escaping predators. The bright colors and patterns as well as the large size of the peacock tail make it difficult to hide. Once predators spot them, peacocks also struggle to fly away, with the heavy tail trailing behind and weighing them down (Figure 4.17). Some researchers have argued that the increased risk is part of the appeal for the peahens: only an especially strong, alert, and healthy peacock would be able to avoid predators while sporting such a spectacular tail.
It’s important to keep in mind that sexual selection relies on the trait being present throughout mating years. Reflecting on the NF1 genetic disorder (see “Special Topic: Neurofibromatosis Type 1 [NF1]”), given how disfiguring the symptoms can become, some might find it surprising that half of the babies born with NF1 inherited it from a parent. Given that the disorder is autosomal dominant and fully penetrant (meaning it has no unaffected carriers), it may seem surprising that sexual selection doesn’t exert more pressure against the mutated alleles. One important factor is that, while the neurofibromas typically begin to appear during puberty, they usually emerge only a few at a time and may grow very slowly. Many NF1 patients don’t experience the more severe or disfiguring symptoms until later in life, long after they have started families of their own.
Some researchers prefer to classify sexual selection separately, as a fifth force of evolution. The traits that underpin mate selection are entirely natural, of course. Research has shown that subtle traits, such as the type of pheromones (hormonal odors related to immune system alleles) someone emits and how those are perceived by the immune system genotype of the “sniffer,” may play crucial and subconscious roles in whether we find someone attractive or not (Chaix, Cao, and Donnelly 2008).
Special Topic: Neurofibromatosis Type 1 (NF1)
Neurofibromatosis Type 1, also known as NF1, is a genetic disorder that illustrates how a mutation in a single gene can affect multiple systems in the body. Surprisingly common, more people have NF1 than cystic fibrosis and muscular dystrophy combined. Even more surprising, given how common it is, is how few people have heard of it. One in every 3,000 babies is born with NF1, and this holds true for all populations worldwide (Riccardi 1992). This means that, for every 3,000 people in your community, there is likely at least one person living with this disorder. NF1 is an autosomal dominant condition, which means that everyone born with a mutation in the gene, whether inherited or spontaneous, has a 50/50 chance of passing it on to each of their own children.
The NF1 disorder results from mutation of the NF1 gene on Chromosome 17. Almost any mutation that affects the sequence of the gene’s protein product, neurofibromin, will cause the disorder. Studies of individuals with NF1 have identified over 3,000 different mutations of all kinds (including point mutations, small and large indels, and translocations). The NF1 gene is one of the largest known genes, containing at least 60 exons (protein-encoding sequences) in a span of about 300,000 nucleotides.
We know that neurofibromin plays an important role in preventing tumor growth because one of the most common symptoms of the NF1 disorder is the growth of benign (noncancerous) tumors, called neurofibromas. Neurofibromas sprout from nerve sheaths—the tissues that encase our nerves—throughout the body, usually beginning around puberty. There is no way to predict where the tumors will occur, or when or how quickly they will grow, although only about 15% turn malignant (cancerous). The two types of neurofibromas that are typically most visible are cutaneous neurofibromas, which are spherical bumps on, or just under, the surface of the skin (Figure 4.18), and plexiform neurofibromas, growths involving whole branches of nerves, often giving the appearance that the surface of the skin is “melting” (Figure 4.19).
Unfortunately, there is currently no cure for NF1. Surgical removal of neurofibromas risks paralysis, due to the high potential for nerve damage, and often results in the tumors growing back even more vigorously. This means that patients are often forced to live with disfiguring and often painful neurofibromas. People who are not familiar with NF1 often mistake neurofibromas for something contagious. This makes it especially hard for people living with NF1 to get jobs working with the public or even to enjoy spending time away from home. Raising public awareness about NF1 and its symptoms can be a great help in improving the quality of life for people living with this condition.
One of the first symptoms of NF1 in a small child is usually the appearance of café-au-lait spots, or CALS, which are flat, brown birthmark-like spots on the skin (Figure 4.20). CALS are often light brown, similar to the color of coffee with cream, which is the reason for the name, although the shade of the pigment depends on a person’s overall complexion. Some babies are born with CALS, but for others the spots appear within the first few years of life. Having six or more CALS larger than five millimeters (mm) across is a strong indicator that a child may have NF1.
Other common symptoms include the following: gliomas (tumors) of the optic nerve, which can cause vision loss; thinning of bones and failure to heal if they break (often requiring amputation); low muscle tone (poor muscle development, often delaying milestones such as sitting up, crawling, and walking); hearing loss, due to neurofibromas on auditory nerves; and learning disabilities, especially those involving spatial reasoning. Approximately 50% of people with NF1 have some type of speech and/or learning disability and often benefit greatly from early intervention services. Generalized developmental disability, however, is not common with NF1, so most people with NF1 live independently as adults. Many people with NF1 live full and successful lives, as long as their symptoms can be managed.
Based on the wide variety of symptoms, it’s clear that the neurofibromin protein plays important roles in many biochemical pathways. While everyone who has NF1 will exhibit some symptoms during their lifetime, there is a great deal of variation in the types and severity of symptoms, even between individuals from the same family who share the exact same NF1 mutation. It seems crazy that a gene with so many important functions would be so susceptible to mutation. Part of this undoubtedly has to do with its massive size—a gene with 300,000 nucleotides has ten times more nucleotides available for mutation than does a gene of 30,000 bases. This also suggests that the mutability of this gene might provide some benefits, which is a possibility that we will revisit later in this chapter.
Special Topic: Sickle Cell Anemia
Sickle cell anemia is an autosomal recessive genetic disorder that affects millions of people worldwide. It is most common in Africa, countries around the Mediterranean Sea, and eastward as far as India. Populations in the Americas that have high percentages of ancestors from these regions also have high rates of sickle cell anemia. In the United States, it’s estimated that 72,000 people live with the disease, with one in approximately 1,200 Hispanic-American babies and one in every 500 African-American babies inheriting the condition (World Health Organization 1996).
Sickle cell anemia affects the hemoglobin protein in red blood cells. Normal red blood cells are somewhat doughnut-shaped—round with a depression on both sides of the middle. They carry oxygen around the bloodstream to cells throughout the body. Red blood cells produced by the mutated form of the gene take on a stiff, sickle-like crescent shape when stressed by low oxygen or dehydration (Figure 4.21). Because of their elongated shape and the fact that they are stiff rather than flexible, they tend to form clumps in the blood vessels, inhibiting blood flow to adjacent areas of the body. This causes episodes of extreme pain and can cause serious problems in the oxygen-deprived tissues. The sickle cells also break down much more quickly than normal cells, often lasting only 20 days rather than the 120 days of normal cells. This causes an overall shortage of blood cells in the sickle cell patient, resulting in low iron (anemia) and problems associated with it such as extreme fatigue, shortness of breath, and hindrances to children’s growth and development.
The devastating effects of sickle cell anemia made its high frequency a pressing mystery. Why would an allele that is so deleterious in its homozygous form be maintained in a population at levels as high as the one in twelve African Americans estimated to carry at least one copy of the allele? The answer turned out to be one of the most interesting cases of balancing selection in the history of genetic study.
While looking for an explanation, scientists noticed that the countries with high rates of sickle cell disease also shared a high risk for another disease called malaria, which is caused by infection of the blood by a Plasmodium parasite. These parasites are carried by mosquitoes and enter the human bloodstream via a mosquito bite. Once infected, the person will experience flu-like symptoms that, if untreated, can often lead to death. Researchers discovered that many people living in these regions seemed to have a natural resistance to malaria. Further study revealed that people who carry the sickle cell allele are far less likely to experience a severe case of malaria. This would not be enough of a benefit to make the allele advantageous for the sickle cell homozygotes, who face shortened life spans due to sickle cell anemia. The real benefit of the sickle cell allele goes to the heterozygotes.
People who are heterozygous for sickle cell carry one normal allele, which produces the normal, round, red blood cells, and one sickle cell allele, which produces the sickle-shaped red blood cells. Thus, they have both the sickle and round blood cell types in their bloodstream. They produce enough of the round red blood cells to avoid the symptoms of sickle cell anemia, but they have enough sickle cells to provide protection from malaria.
When the Plasmodium parasites infect an individual, they begin to multiply in the liver, but then must infect the red blood cells to complete their reproductive cycle. When the parasites enter sickle-type cells, the cells respond by taking on the sickle shape. This prevents the parasite from circulating through the bloodstream and completing its life cycle, greatly inhibiting the severity of the infection in the sickle cell heterozygotes compared to non–-sickle cell homozygotes. See Chapter 14 for more discussion of sickle cell anemia.
Special Topic: The Real Primordial Cells—Dictyostelium Discoideum
The amoeba-like primordial cells that were used as recurring examples throughout this chapter are inspired by actual research that is truly fascinating. In 2015, Gareth Bloomfield and colleagues reported on their genomic study of the social amoeba Dictyostelium discoideum (a.k.a. “slime molds,” although technically they are amoebae, not molds). Strains of these amoebae have been grown in research laboratories for many decades and are useful in studying the mechanisms that amoeboid single-celled organisms use to ingest food and liquid. For simplification of our examples in this chapter, our amoeba-like cells remained ocean dwellers. Wild Dictyostelium discoideum, however, live in soil and feed on soil bacteria by growing ruffles in their membranes that reach out to encapsulate the bacterial cell. Laboratory strains, however, are typically raised on liquid media (agar) in Petri dishes, which is not suitable for the wild-type amoebae. It was widely known that the laboratory strains must have developed mutations in one or more genes to allow them to ingest the larger nutrient particles in the agar and larger volumes of liquid, but the genes involved were not known.
Bloomfield and colleagues performed genomic testing on both the wild and the laboratory strains of Dictyostelium discoideum. Their discovery was astounding: every one of the laboratory strains carried a mutation in the NF1 gene, the very same gene associated with Neurofibromatosis Type 1 (NF1) in humans. The antiquity of this massive, easily mutated gene is incredible. It originated in an ancestor common to both humans and these amoebae, and it has been retained in both lineages ever since. As seen in Dictyostelium discoideum, breaking the gene can be advantageous. Without a functioning copy of the neurofibromin protein, the cell membrane is able to form much-larger feeding structures, allowing the NF1 mutants to ingest larger particles and larger volumes of liquid. For these amoebae, this may provide dietary flexibility that functions somewhat like an insurance policy for times when the food supply is limited.
Dictyostelium discoideum are also interesting in that they typically reproduce asexually, but under certain conditions, one cell will convert into a “giant” cell, which encapsulates surrounding cells, transforming into one of three sexes. This cell will undergo meiosis, producing gametes that must combine with one of the other two sexes to produce viable offspring. This ability for sexual reproduction may be what allows Dictyostelium discoideum to benefit from the advantages of NF1 mutation, while also being able to restore the wild type NF1 gene in future generations.
What does this mean for humans living with NF1? Well, understanding the role of the neurofibromin protein in the membranes of simple organisms like Dictyostelium discoideum may help us to better understand how it functions and malfunctions in the sheaths of human neurons. It’s also possible that the mutability of the NF1 gene confers certain advantages to humans as well. Alleles of the NF1 gene have been found to reduce one’s risk for alcoholism (Repunte-Canonigo Vez et al. 2015), opiate addiction (Sanna et al. 2002), Type 2 diabetes (Martins et al. 2016), and hypomusicality (a lower-than-average musical aptitude; Cota et al. 2018). This research is ongoing and will be exciting to follow in the coming years.
Studying Evolution in Action
The Hardy-Weinberg Equilibrium
This chapter has introduced you to the forces of evolution, the mechanisms by which evolution occurs. How do we detect and study evolution, though, in real time, as it happens? One tool we use is the Hardy-Weinberg Equilibrium: a mathematical formula that allows estimation of the number and distribution of dominant and recessive alleles in a population. This aids in determining whether allele frequencies are changing and, if so, how quickly over time, and in favor of which allele? It’s important to note that the Hardy-Weinberg formula only gives us an estimate based on the data for a snapshot in time. We will have to calculate it again later, after various intervals, to determine if our population is evolving and in what way the allele frequencies are changing. To learn how to calculate the Hardy-Weinberg formula, see “Special Topic: Calculating the Hardy-Weinberg Equilibrium” at the end of the chapter.
Special Topic: Calculating the Hardy-Weinberg Equilibrium (include here but not as special topic just normal paragraph)
Interpreting Evolutionary Change: Nonrandom Mating
Once we have detected change occurring in a population, we need to consider which evolutionary processes might be the cause of the change. It is important to watch for nonrandom mating patterns, to see if they can be included or excluded as possible sources of variation in allele frequencies.
Nonrandom mating (also known as assortative mating) occurs when mate choice within a population follows a nonrandom pattern.
Positive assortative mating patterns result from a tendency for individuals to mate with others who share similar phenotypes. This often happens based on body size. Taking as an example dog breeds, it is easier for two Chihuahuas to mate and have healthy offspring than it is for a Chihuahua and a St. Bernard to do so. This is especially true if the Chihuahua is the female and would have to give birth to giant St. Bernard pups.
Negative assortative mating patterns occur when individuals tend to select mates with qualities different from their own. This is what is at work when humans choose partners whose pheromones indicate that they have different and complementary immune alleles, providing potential offspring with a better chance at a stronger immune system.
Among domestic animals, such as pets and livestock, assortative mating is often directed by humans who decide which pairs will mate to increase the chances of offspring having certain desirable traits. This is known as artificial selection.
Among humans, in addition to phenotypic traits, cultural traits such as religion and ethnicity may also influence assortative mating patterns.
Defining a Species
Species are organisms whose individuals are capable of breeding because they are biologically and behaviorally compatible to produce viable, fertile offspring. Viable offspring are those offspring that are healthy enough to survive to adulthood. Fertile offspring are able to reproduce successfully, resulting in offspring of their own. Both conditions must be met for individuals to be considered part of the same species. As you can imagine, these criteria complicate the identification of distinct species in fossilized remains of extinct populations. In those cases, we must examine how much phenotypic variation is typically found within a comparable modern-day species; we can then determine whether the fossilized remains fall within the expected range of variation for a single species.
Some species have subpopulations that are regionally distinct. These are classified as separate subspecies because they have their own unique phenotypes and are geographically isolated from one another. However, if they do happen to encounter one another, they are still capable of successful interbreeding.
There are many examples of sterile hybrids that are offspring of parents from two different species. For example, horses and donkeys can breed and have offspring together. Depending on which species is the mother and which is the father, the offspring are either called mules, or hennies. Mules and hennies can live full life spans but are not able to have offspring of their own. Likewise, tigers and lions have been known to mate and have viable offspring. Again, depending on which species is the mother and which is the father, these offspring are called either ligers or tigons. Like mules and hennies, ligers and tigons are unable to reproduce. In each of these cases, the mismatched set of chromosomes that the offspring inherit produce an adequate set of functioning genes for the hybrid offspring; however, once mixed and divided in meiosis, the gametes don’t contain the full complement of genes needed for survival in the third generation.
Micro- to Macroevolution
Microevolution refers to changes in allele frequencies within breeding populations—that is, within single species. Macroevolution describes how the similarities and differences between species, as well as the phylogenetic relationships with other taxa, lead to changes that result in the emergence of new species. Consider our example of the peppered moth that illustrated microevolution over time, via directional selection favoring the peppered allele when the trees were clean and the dark pigment allele when the trees were sooty. Imagine that environmental regulations had cleaned up the air pollution in one part of the nation, while the coal-fired factories continued to spew soot in another area. If this went on long enough, it’s possible that two distinct moth populations would eventually emerge—one containing only the peppered allele and the other only harboring the dark pigment allele.
When a single population divides into two or more separate species, it is called speciation. The changes that prevent successful breeding between individuals who descended from the same ancestral population may involve chromosomal rearrangements, changes in the ability of the sperm from one species to permeate the egg membrane of the other species, or dramatic changes in hormonal schedules or mating behaviors that prevent members from the new species from being able to effectively pair up.
There are two types of speciation: allopatric and sympatric. Allopatric speciation is caused by long-term isolation (physical separation) of subgroups of the population (Figure 4.22). Something occurs in the environment—perhaps a river changes its course and splits the group, preventing them from breeding with members on the opposite riverbank. Over many generations, new mutations and adaptations to the different environments on each side of the river may drive the two subpopulations to change so much that they can no longer produce fertile, viable offspring, even if the barrier is someday removed.
Sympatric speciation occurs when the population splits into two or more separate species while remaining located together without a physical barrier. This typically results from a new mutation that pops up among some members of the population that prevents them from successfully reproducing with anyone who does not carry the same mutation. This is seen particularly often in plants, as they have a higher frequency of chromosomal duplications.
One of the quickest rates of speciation is observed in the case of adaptive radiation. Adaptive radiation refers to the situation in which subgroups of a single species rapidly diversify and adapt to fill a variety of ecological niches. An ecological niche is a set of constraints and resources that is available in an environmental setting. Evidence for adaptive radiations is often seen after population bottlenecks. A mass disaster kills off many species, and the survivors have access to a new set of territories and resources that were either unavailable or much coveted and fought over before the disaster. The offspring of the surviving population will often split into multiple species, each of which stems from members in that first group of survivors who happened to carry alleles that were advantageous for a particular niche.
The classic example of adaptive radiation brings us back to Charles Darwin and his observations of the many species of finches on the Galapagos Islands. We are still not sure how the ancestral population of finches first arrived on that remote Pacific Island chain, but they found themselves in an environment filled with various insects, large and tiny seeds, fruit, and delicious varieties of cactus. Some members of that initial population carried alleles that gave them advantages for each of these dietary niches. In subsequent generations, others developed new mutations, some of which were beneficial. These traits were selected for, making the advantageous alleles more common among their offspring. As the finches spread from one island to the next, they would be far more likely to find mates among the birds on their new island. Birds feeding in the same area were then more likely to mate together than birds who have different diets, contributing to additional assortative mating. Together, these evolutionary mechanisms caused rapid speciation that allowed the new species to make the most of the various dietary niches (Figure 4.23).
In today’s modern world, understanding these evolutionary processes is crucial for developing immunizations and antibiotics that can keep up with the rapid mutation rate of viruses and bacteria. This is also relevant to our food supply, which relies, in large part, on the development of herbicides and pesticides that keep up with the mutation rates of pests and weeds. Viruses, bacteria, agricultural pests, and weeds have all shown great flexibility in developing alleles that make them resistant to the latest medical treatment, pesticide, or herbicide. Billion-dollar industries have specialized in trying to keep our species one step ahead of the next mutation in the pests and infectious diseases that put our survival at risk.
Special Topic: Calculating the Hardy-Weinberg Equilibrium
In the Hardy-Weinberg formula, p represents the frequency of the dominant allele, and q represents the frequency of the recessive allele. Remember, an allele’s frequency is the proportion, or percentage, of that allele in the population. For the purposes of Hardy-Weinberg, we give the allele percentages as decimal numbers (e.g., 42% = 0.42), with the entire population (100% of alleles) equaling 1. If we can figure out the frequency of one of the alleles in the population, then it is simple to calculate the other. Simply subtract the known frequency from 1 (the entire population): 1 – p = q and 1 – q = p.
The Hardy-Weinberg formula is p2 + 2pq + q2, where:
p2 represents the frequency of the homozygous dominant genotype;
2pq represents the frequency of the heterozygous genotype; and
q2 represents the frequency of the homozygous recessive genotype.
It is often easiest to determine q2 first, simply by counting the number of individuals with the unique, homozygous recessive phenotype (then dividing by the total individuals in the population to arrive at the “frequency”). Once we have this number, we simply need to calculate the square root of the homozygous recessive phenotype frequency. That gives us q. Remember, 1 – q equals p, so now we have the frequencies for both alleles in the population. If we needed to figure out the frequencies of heterozygotes and homozygous dominant genotypes, we’d just need to plug the p and q frequencies back into the p2 and 2pq formulas.
Let’s imagine we have a population of ladybeetles that carries two alleles: a dominant allele that produces red ladybeetles and a recessive allele that produces orange ladybeetles. Since red is dominant, we’ll use R to represent the red allele, and r to represent the orange allele. Our population has ten beetles, and seven are red and three are orange (Figure 4.24). Let’s calculate the number of genotypes and alleles in this population.
Of ten total beetles, we have three orange beetles3/10 = .30 (30%) frequency—and we know they are homozygous recessive (rr). So:
rr = .3; therefore, r = √.3 = .5477
R = 1 – .5477 = .4523
Using the Hardy-Weinberg formula:
1=.45232 + 2 x .4523 x .5477 +.54772 = .20 + .50 + .30 = 1
Thus, the genotype breakdown is 20% RR, 50% Rr, and 30% rr
(2 red homozygotes, 5 red heterozygotes, and 3 orange homozygotes).
Since we have 10 individuals, we know we have 20 total alleles: 4 red from the RR group, 5 red and 5 orange from the Rr group, and 6 orange from the rr group, for a grand total of 9 red and 11 orange (45% red and 55% orange, just like we estimated in the 1 – q step).
Reminder: The Hardy-Weinberg formula only gives us an estimate for a snapshot in time. We will have to calculate it again later, after various intervals, to determine if our population is evolving and in what way the allele frequencies are changing.
Review Questions
- Summarize the Modern Synthesis and provide several examples of how it is relevant to questions and problems in our world today.
- You inherit a house from a long-lost relative that contains a fancy aquarium, filled with a variety of snails. The phenotypes include large snails and small snails; red, black, and yellow snails; and solid, striped, and spotted snails. Devise a series of experiments that would help you determine how many snail species are present in your aquarium.
- Match the correct force of evolution with the correct real-world example:
a. Mutationi. 5-alpha reductase deficiency
b. Genetic Driftii. Peppered Moths
c. Gene Flowiii. Neurofibromatosis Type 1
d. Natural Selectioniv. Scutellata Honey Bees - Imagine a population of common house mice (Mus musculus). Draw a comic strip illustrating how mutation, genetic drift, gene flow, and natural selection might transform this population over several (or more) generations.
-
The many breeds of the single species of domestic dog (Canis familiaris) provide an extreme example of microevolution. Discuss why this is the case. What future scenarios can you imagine that could potentially transform the domestic dog into an example of macroevolution?
-
The ability to roll one’s tongue (lift the outer edges of the tongue to touch each other, forming a tube) is a dominant trait. In a small town of 1,500 people, 500 can roll their tongues. Use the Hardy-Weinberg formula to determine how many individuals in the town are homozygous dominant, heterozygous, and homozygous recessive.
Key Terms
5-alpha reductase deficiency: An autosomal recessive syndrome that manifests when a child having both X and Y sex chromosomes inherits two nonfunctional (mutated) copies of the SRD5A2 gene, producing a deficiency in a hormone necessary for development in infancy of typical male genitalia. These children often appear at birth to have female genitalia, but they develop a penis and other sexual characteristics when other hormones kick in during puberty.
Adaptive radiation: The situation in which subgroups of a single species rapidly diversify and adapt to fill a variety of ecological niches.
Admixture: A term often used to describe gene flow between human populations. Sometimes also used to describe gene flow between nonhuman populations.
Allele frequency: The ratio, or percentage, of one allele compared to the other alleles for that gene within the study population.
Alleles: Variant forms of genes.
Allopatric speciation: Speciation caused by long-term isolation (physical separation) of subgroups of the population.
Antibiotics: Medicines prescribed to treat bacterial infections.
Artificial selection: Human-directed assortative mating among domestic animals, such as pets and livestock, designed to increase the chances of offspring having certain desirable traits.
Asexual reproduction: Reproduction via mitosis, whereby offspring are clones of the parents.
Autosomal dominant: A phenotype produced by a gene on an autosomal chromosome that is expressed, to the exclusion of the recessive phenotype, in heterozygotes.
Autosomal recessive: A phenotype produced by a gene on an autosomal chromosome that is expressed only in individuals homozygous for the recessive allele.
Balanced translocations: Chromosomal translocations in which the genes are swapped but no genetic information is lost.
Balancing selection: A pattern of natural selection that occurs when the extremes of a trait are selected against, favoring the intermediate phenotype (a.k.a. stabilizing selection).
Beneficial mutations: Mutations that produce some sort of an advantage to the individual.
Benign: Noncancerous. Benign tumors may cause problems due to the area in which they are located (e.g., they might put pressure on a nerve or brain area), but they will not release cells that aggressively spread to other areas of the body.
Café-au-lait spots (CALS): Flat, brown birthmark-like spots on the skin, commonly associated with Neurofibromatosis Type 1.
Chromosomal translocations: The transfer of DNA between nonhomologous chromosomes.
Chromosomes: Molecules that carry collections of genes.
Codons: Three-nucleotide units of DNA that function as three-letter “words,” encoding instructions for the addition of one amino acid to a protein or indicating that the protein is complete.
Cretaceous–Paleogene extinction: A mass disaster caused by an asteroid that struck the earth approximately 66 million years ago and killed 75% of life on Earth, including all terrestrial dinosaurs. (a.k.a. K-Pg Extinction, Cretatious-Tertiary Extinction, and K-T Extinction).
Crossover events: Chromosomal alterations that occur when DNA is swapped between homologous chromosomes while they are paired up during meiosis I.
Cutaneous neurofibromas: Neurofibromas that manifest as spherical bumps on or just under the surface of the skin.
Deleterious mutation: A mutation producing negative effects to the individual such as the beginnings of cancers or heritable disorders.
Deletions: Mutations that involve the removal of one or more nucleotides from a DNA sequence.
Derivative chromosomes: New chromosomal structures resulting from translocations.
Dictyostelium discoideum: A species of social amoebae that has been widely used for laboratory research. Laboratory strains of Dictyostelium discoideum all carry mutations in the NF1 gene, which is what allows them to survive on liquid media (agar) in Petri dishes.
Directional selection: A pattern of natural selection in which one phenotype is favored over the other, causing the frequencies of the associated advantageous alleles to gradually increase.
Disruptive selection: A pattern of natural selection that occurs when both extremes of a trait are advantageous and intermediate phenotypes are selected against (a.k.a. diversifying selection).
DNA repair mechanisms: Enzymes that patrol and repair DNA in living cells.
DNA transposons: Transposons that are clipped out of the DNA sequence itself and inserted elsewhere in the genome.
Ecological niche: A set of constraints and resources that are available in an environmental setting.
Ellis-van Creveld syndrome: An autosomal recessive disorder characterized by short stature (dwarfism), polydactyly (the development of more than five digits [fingers or toes] on the hands or feet), abnormal tooth development, and heart defects. Estimated to affect approximately one in 60,000 individuals worldwide, among the Old Order Amish of Lancaster County, the rate is estimated to be as high as one in every 200 births.
Evolution: A change in the allele frequencies in a population over time.
Exons: The DNA sequences within a gene that directly encode protein sequences. After being transcribed into messenger RNA, the introns (DNA sequences within a gene that do not directly encode protein sequences) are clipped out, and the exons are pasted together prior to translation.
Fertile offspring: Offspring that can successfully reproduce, resulting in offspring of their own.
Founder effect: A type of genetic drift that occurs when members of a population leave the main or “parent” group and form a new population that no longer interbreeds with the other members of the original group.
Frameshift mutations: Types of indels that involve the insertion or deletion of any number of nucleotides that is not a multiple of three. These “shift the reading frame” and cause all codons beyond the mutation to be misread.
Gametes: The reproductive cells, produced through meiosis (a.k.a. germ cells or sperm or egg cells).
Gene: A sequence of DNA that provides coding information for the construction of proteins.
Gene flow: The movement of alleles from one population to another. This is one of the forces of evolution.
Gene pool: The entire collection of genetic material in a breeding community that can be passed on from one generation to the next.
Genetic drift: Random changes in allele frequencies within a population from one generation to the next. This is one of the forces of evolution.
Genotype: The set of alleles that an individual has for a given gene.
Genotype frequencies: The ratios or percentages of the different homozygous and heterozygous genotypes in the population.
Guevedoces: The term coined locally in the Dominican Republic for the condition scientifically known as 5-alpha reductase deficiency. The literal translation is “penis at twelve.”
Hardy-Weinberg Equilibrium: A mathematical formula (1=p2 + 2pq + q2 ) that allows estimation of the number and distribution of dominant and recessive alleles in a population.
Harlequin ladybeetle: A species of ladybeetle, native to East Asia, that was introduced to Europe and the Americas as a form of pest control. After many decades of use, one of the North American strains developed the ability to reproduce in diverse environments, causing it to spread rapidly throughout the Americas, Europe, and Africa. It has hybridized with European strains and is now a major pest in its own right.
Heterozygous genotype: A genotype comprising two different alleles.
Homozygous genotype: A genotype comprising an identical set of alleles.
Hybridization: A term often used to describe gene flow between nonhuman populations.
Inbreeding: The selection of mates exclusively from within a small, closed population.
Indels: A class of mutations that includes both insertions and deletions.
Inherited mutation: A mutation that has been passed from parent to offspring.
Insertions: Mutations that involve the addition of one or more nucleotides into a DNA sequence.
Isolation: Prevention of a population subgroup from breeding with other members of the same species due to a physical barrier or, in humans, a cultural rule.
Last Universal Common Ancestor (LUCA): The ancient organism from which all living things on Earth are descended.
Macroevolution: Changes that result in the emergence of new species, how the similarities and differences between species, as well as the phylogenetic relationships with other taxa, lead to changes that result in the emergence of new species.
Malaria: A frequently deadly mosquito-borne disease caused by infection of the blood by a Plasmodium parasite.
Malignant: Cancerous. Malignant tumors grow aggressively and their cells may metastasize (travel through the blood or lymph systems) to form new, aggressive tumors in other areas of the body.
Microevolution: Changes in allele frequencies within breeding populations—that is, within a single species.
Modern Synthesis: The integration of Darwin’s, Mendel’s, and subsequent research into a unified theory of evolution.
Monosomies: Conditions resulting from a nondisjunction event, in which a cell ends up with only one copy of a chromosome. In humans, a single X chromosome is the only survivable monosomy.
Mutation: A change in the nucleotide sequence of the genetic code. This is one of the forces of evolution.
Natural selection: An evolutionary process that occurs when certain phenotypes confer an advantage or disadvantage in survival and/or reproductive success. This is one of the forces of evolution, and it was first identified by Charles Darwin.
Negative assortative mating: A pattern that occurs when individuals tend to select mates with qualities different from their own.
Neurofibromas: Nerve sheath tumors that are common symptoms of Neurofibromatosis Type 1.
Neurofibromatosis Type 1: An autosomal dominant genetic disorder affecting one in every 3,000 people. It is caused by mutation of the NF1 gene on Chromosome 17, resulting in a defective neurofibromin protein. The disorder is characterized by neurofibromas, café-au-lait spots, and a host of other potential symptoms.
NF1: An abbreviation for Neurofibromatosis Type 1. When italicized, NF1 refers to the gene on Chromosome 17 that encodes the neurofibromin protein.
Nondisjunction events: Chromosomal abnormalities that occur when the homologous chromosomes (in meiosis I) or sister chromatids (in meiosis II and mitosis) fail to separate after pairing. The result is that both chromosomes or chromatids end up in the same daughter cell, leaving the other daughter cell without any copy of that chromosome.
Nonrandom mating: A scenario in which mate choice within a population follows a nonrandom pattern (a.k.a. assortative mating).
Nonsynonymous mutation: A point mutation that causes a change in the resulting protein.
Old Order Amish: A culturally isolated population in Lancaster County, Pennsylvania, that has approximately 50,000 current members, all of whom can trace their ancestry back to a group of approximately eighty individuals. This group has high rates of certain genetics disorders, including Ellis-van Creveld syndrome.
Origins of life: How the first living organism came into being.
Peacock: The male sex of the peafowl, famous for its large, colorful tail, which it dramatically displays to attract mates. (The female of the species is known as a peahen.)
Peppered moth: A species of moth (Biston betularia) found in England that has light and dark phenotypes. During the Industrial Revolution, when soot blackened the trees, the frequency of the previously rare dark phenotype dramatically increased, as lighter-colored moths were easier for birds to spot against the sooty trees. After environmental regulations eliminated the soot, the lighter-colored phenotype gradually became most common again.
Phenotype: The observable traits that are produced by a genotype.
Phylogenetic tree of life: A family tree of all living organisms, based on genetic relationships.
Phylogenies: Genetically determined family lineages.
Plasmodium: A genus of mosquito-borne parasite. Several Plasmodium species cause malaria when introduced to the human bloodstream via a mosquito bite.
Plexiform neurofibromas: Neurofibromas that involve whole branches of nerves, often giving the appearance that the surface of the skin is “melting.”
Point mutation: A single-letter (single-nucleotide) change in the genetic code, resulting in the substitution of one nucleic acid base for a different one.
Polymorphisms: Multiple forms of a trait; alternative phenotypes within a given species.
Population: A group of individuals who are genetically similar enough and geographically near enough to one another that they can breed and produce new generations of individuals.
Population bottleneck: A type of genetic drift that occurs when the number of individuals in a population drops dramatically due to some random event.
Positive assortative mating: A pattern that results from a tendency for individuals to mate with others who share similar phenotypes.
Retrotransposons: Transposons that are transcribed from DNA into RNA, and then are “reverse transcribed,” to insert the copied sequence into a new location in the DNA.
Scutellata honey bees: A strain of honey bees that resulted from the hybridization of African and European honey bee subspecies. These bees were accidentally released into the wild in 1957 in Brazil and have since spread throughout South and Central America and into the United States. Also known as “killer bees,” they tend to be very aggressive in defense of their hives and have caused many fatal injuries to humans and livestock.
Sexual reproduction: Reproduction via meiosis and combination of gametes. Offspring inherit genetic material from both parents.
Sexual selection: An aspect of natural selection in which the selective pressure specifically affects reproductive success (the ability to successfully breed and raise offspring).
Sickle cell anemia: An autosomal recessive genetic disorder that affects millions of people worldwide. It is most common in Africa, countries around the Mediterranean Sea, and eastward as far as India. Homozygotes for the recessive allele develop the disorder, which produce misshapen red blood cells that cause iron deficiency, painful episodes of oxygen-deprivation in localized tissues, and a host of other symptoms. In heterozygotes, though, the sickle cell allele confers a greater resistance to malaria.
Somatic cells: The cells of our organs and other body tissues (all cells except gametes) that replicate by mitosis.
Speciation: The process by which a single population divides into two or more separate species.
Species: Organisms whose individuals are capable of breeding because they are biologically and behaviorally compatible to produce viable, fertile offspring.
Spontaneous mutation: A mutation that occurs due to random chance or unintentional exposure to mutagens. In families, a spontaneous mutation is the first case, as opposed to mutations that are inherited from parents.
Subspecies: A distinct subtype of a species. Most often, this is a geographically isolated population with unique phenotypes; however, it remains biologically and behaviorally capable of interbreeding with other populations of the same species.
Sympatric speciation: When a population splits into two or more separate species while remaining located together without a physical (or cultural) barrier.
Synonymous mutation: A point mutation that does not change the resulting protein.
Transposable elements: Fragments of DNA that can “jump” around in the genome.
Transposon: Another term for “transposable element.”
Trisomies: Conditions in which three copies of the same chromosome end up in a cell, resulting from a nondisjunction event. Down syndrome, Edwards syndrome, and Patau syndrome are trisomies.
Unbalanced translocations: Chromosomal translocations in which there is an unequal exchange of genetic material, resulting in duplication or loss of genes.
UV crosslinking: A type of mutation in which adjacent thymine bases bind to one another in the presence of UV light.
Viable offspring: Offspring that are healthy enough to survive to adulthood.
Xeroderma pigmentosum: An autosomal recessive disease in which DNA repair mechanisms do not function correctly, resulting in a host of problems especially related to sun exposure, including severe sunburns, dry skin, heavy freckling, and other pigment changes.
About the Author
Andrea J. Alveshere, Ph.D.
Western Illinois University, a-alveshere@wiu.edu, Andrea Alveshere: Archaeologist & Biological Anthropologist
Dr. Andrea Alveshere is an associate professor of anthropology and chemistry at Western Illinois University. Her research focuses on relationships between humans and their environments, including cultural and biological adaptations surrounding ancient diet, health, and knowledge systems; genetic disorders such as Neurofibromatosis Type 1 (NF1); effects of environmental factors on the preservation of bones, plant remains, and the molecules within them; and the comparative utility of field and laboratory techniques to produce informative archaeological, nutritional, and forensic data.
Dr. Alveshere earned her B.A. in anthropology at the University of Washington with an emphasis in archaeology and an undergraduate research focus on the analysis of skeletal remains and geoarchaeological deposits. At the University of Minnesota, she completed her Ph.D. in anthropology, with a minor in human genetics. Her graduate thesis investigated factors that influence the preservation and detection of DNA in ancient and forensic specimens.
Dr. Alveshere also worked for several years as a forensic scientist in the DNA/Biology section of the Minnesota Bureau of Criminal Apprehension Forensic Science Laboratory. She led the WIU Archaeological Field School, on alternate summers since 2017, and conducted archaeological excavations in Israel, South Africa, and throughout the midwestern United States.
For Further Exploration
Explore Evolution on HHMI’s Biointeractive website.
Teaching Evolution through Human Examples, Smithsonian Museum of Natural History websites.
References
Bloomfield, Gareth, David Traynor, Sophia P. Sander, Douwe M. Veltman, Justin A. Pachebat, and Robert R. Kay. 2015. “Neurofibromin Controls Macropinocytosis and Phagocytosis in Dictyostelium.” eLife 4:e04940.
Chaix, Raphaëlle, Chen Cao, and Peter Donnelly. 2008. “Is Mate Choice in Humans MHC-Dependent?” PLoS Genetics 4 (9): e1000184.
Cook, Laurence M. 2003. "The Rise and Fall of the Carbonaria Form of the Peppered Moth." The Quarterly Review of Biology 78 (4): 399–417.
Cota, Bruno Cézar Lage, João Gabriel Marques Fonseca, Luiz Oswaldo Carneiro Rodrigues, Nilton Alves de Rezende, Pollyanna Barros Batista, Vincent Michael Riccardi, and Luciana Macedo de Resende. 2018. “Amusia and Its Electrophysiological Correlates in Neurofibromatosis Type 1.” Arquivos de Neuro-Psiquiatria 76 (5): 287–295.
D’Asdia, Maria Cecilia, Isabella Torrente, Federica Consoli, Rosangela Ferese, Monia Magliozzi, Laura Bernardini, Valentina Guida, et al. 2013. “Novel and Recurrent EVC and EVC2 Mutations in Ellis-van Creveld Syndrome and Weyers Acrofacial Dyostosis.” European Journal of Medical Genetics 56 (2): 80–87.
Dobzhansky, Theodosius. 1937. Genetics and the Origin of Species. Columbia University Biological Series. New York: Columbia University Press.
Facon, Benoît, Laurent Crespin, Anne Loiseau, Eric Lombaert, Alexandra Magro, and Arnaud Estoup. 2011. “Can Things Get Worse When an Invasive Species Hybridizes? The Harlequin Ladybird Harmonia axyridis in France as a Case Study.” Evolutionary Applications 4 (1): 71–88.
Fisher, Ronald A. 1919. "The Correlation between Relatives on the Supposition of Mendelian Inheritance." Transactions of the Royal Society of Edinburgh 52 (2): 399–433.
Ford, E. B. 1942. Genetics for Medical Students. London: Methuen.
Ford, E. B. 1949. Mendelism and Evolution. London: Methuen.
Grant, Bruce S. 1999. “Fine-tuning the Peppered Moth Paradigm.” Evolution 53 (3): 980–984.
Haldane, J. B. S. 1924. “A Mathematical Theory of Natural and Artificial Selection (Part 1).” Transactions of the Cambridge Philosophical Society 23 (2):19–41.
Imperato-McGinley, J., and Y.-S. Zhu. 2002. “Androgens and Male Physiology: The Syndrome of 5 Alpha-Reductase-2 Deficiency.” Molecular and Cellular Endocrinology 198 (1-2): 51–59.
Jablonski, David, and W. G. Chaloner. 1994. "Extinctions in the Fossil Record.” Philosophical Transactions of the Royal Society of London B: Biological Sciences 344 (1307): 11–17.
Livi-Bacci, Massimo. 2006. “The Depopulation of Hispanic America after the Conquest.” Population Development and Review 32 (2): 199–232.
Lombaert, Eric, Thomas Guillemaud, Jean-Marie Cornuet, Thibaut Malausa, Benoît Facon, and Arnaud Estoup. 2010. "Bridgehead Effect in the Worldwide Invasion of the Biocontrol Harlequin Ladybird.” PLoS ONE 5 (3): e9743.
Martins, Aline Stangherlin, Ann Kristine Jansen, Luiz Oswaldo Carneiro Rodrigues, Camila Maria Matos, Marcio Leandro Ribeiro Souza, Juliana Ferreira de Souza, Maria de Fátima Haueisen Sander Diniz, et al. 2016. “Lower Fasting Blood Glucose in Neurofibromatosis Type 1.” Endocrine Connections 5 (1): 28–33.
Pickering, Gary, James Lin, Roland Riesen, Andrew Reynolds, Ian Brindle, and George Soleas. 2004. "Influence of Harmonia axyridis on the Sensory Properties of White and Red Wine." American Journal of Enology and Viticulture 55 (2): 153–159.
Repunte-Canonigo Vez, Melissa A. Herman, Tomoya Kawamura, Henry R. Kranzler, Richard Sherva, Joel Gelernter, Lindsay A. Farrer, Marisa Roberto, and Pietro Paolo Sanna. 2015. “NF1 Regulates Alcohol Dependence-Associated Excessive Drinking and Gamma-Aminobutyric Acid Release in the Central Amygdala in Mice and Is Associated with Alcohol Dependence in Humans.” Biological Psychiatry 77 (10): 870–879.
Riccardi, Vincent M. 1992. Neurofibromatosis: Phenotype, Natural History, and Pathogenesis. Baltimore: Johns Hopkins University Press.
Sanford, Malcolm T. 2006. "The Africanized Honey Bee in the Americas: A Biological Revolution with Human Cultural Implications, Part V—Conclusion." American Bee Journal 146 (7): 597–599.
Sanna, Pietro Paolo, Cindy Simpson, Robert Lutjens, and George Koob. 2002. “ERK Regulation in Chronic Ethanol Exposure and Withdrawal.” Brain Research 948 (1–2): 186–191.
World Health Organization. 1996. “Control of Hereditary Disorders: Report of WHO Scientific meeting (1996).” WHO Technical Reports 865. Geneva: World Health Organization.
World Health Organization. 2017. “Global Priority List of Antibiotic-Resistant Bacteria to Guide Research, Discovery, and Development of New Antibiotics.” Global Priority Pathogens List, February 27. Geneva: World Health Organization. https://www.who.int/medicines/publications/WHO-PPL-Short_Summary_25Feb-ET_NM_WHO.pdf.
Wright, Sewall. 1932. "The Roles of Mutation, Inbreeding, Crossbreeding, and Selection in Evolution." Proceedings of the Sixth International Congress on Genetics 1 (6): 356–366.
Acknowledgment
Many thanks to Dr. Vincent M. Riccardi for sharing his vast knowledge of neurofibromatosis and for encouraging me to explore it from an anthropological perspective.
Stephanie Etting, Ph.D., Sacramento City College
This chapter is a revision from "Chapter 5: Meet the Living Primates” by Stephanie Etting. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Learning Objectives
- Describe how studying nonhuman primates is important in anthropology.
- Compare two ways of categorizing taxa: grades and clades.
- Define different types of traits used to evaluate primate taxa.
- Identify key ways that primates differ from other mammals.
- Distinguish between the major primate taxa using their key characteristics.
- Describe your place in nature by learning your taxonomic classification.
You may be wondering why a field dedicated to the study of humans includes discussions of nonhuman animals. Our primary goal in biological anthropology is to understand how humans are similar to and different from the rest of the natural world, why we have the traits we have, and how we got to be the way we are. But to fully grasp our place in nature, we must look to our closest living relatives, the nonhuman primates. In this chapter, we focus on the organization and diversity within the Order Primates.
Studying Primates in Biological Anthropology
Primates are one of at least twenty Orders belonging to the Class Mammalia, and probably one of the oldest. One genetic estimate puts the origin of primates at approximately 91 million years ago (mya), predating the extinction of the dinosaurs (Bininda-Emonds et al. 2007). Today, the Order Primates is a diverse group of animals that includes lemurs and lorises, tarsiers, monkeys, apes, and humans, all of which are united in sharing a suite of anatomical, behavioral, and life history characteristics. While nonhuman primates are fascinating animals in their own right, their close relationship to humans makes them ideal for studying humans via homology, looking at traits that are shared between taxa because they inherited the trait from a common ancestor. For example, humans (genus Homo) and chimpanzees (genus Pan) both share the trait of male cooperation in hunting. This trait—along with many others that chimpanzees and humans share—is likely homologous, meaning it was probably passed down from the last common ancestor of Homo and Pan, which lived about 6–8 million years ago.
Nonhuman primates also make excellent comparators for learning about humans via analogy. Many nonhuman primates live in environments similar to those in which our ancestors lived and therefore exhibit traits similar to what we see in humans. For example, baboons and humans both have long legs. In humans, this is because about 1.7 million years ago, our ancestors moved into savanna habitats where longer legs helped them move more efficiently over long distances. Baboons, who also live in savanna habitats, independently evolved longer arms and legs for the same reason—to be able to cover more ground, more efficiently. This means that having long legs is an analogous trait in baboons and humans: —that is, this adaptation evolved independently in the two species but for the same purpose. Using homology and analogy, our closest living relatives provide the critical context in which to understand human biology, morphology, and behavior. It is only by studying how humans compare with our primate relatives that we can fully comprehend our place in nature.
Ways of Organizing Taxa
You learned in Chapter 2 about Linnaeus and the hierarchical nature of taxonomic classification. Our goal in classifying taxa is to create categories that reflect clade relationships. A clade is a grouping of organisms based on relatedness that reflects a branch of the evolutionary tree. Clade relationships are determined using traits shared by groups of taxa as well as genetic similarities. An example of a clade would be a grouping that includes humans, chimpanzees, bonobos, and gorillas (Figure 5.1). These taxa are in what is referred to as the African clade of hominoids (a taxonomic group you will learn about later in this chapter). The African clade grouping reflects how humans, chimpanzees, bonobos, and gorillas all share a more recent ancestor with each other than any of them do with other species—that is, we are on the same branch of the evolutionary tree. We know members of the African clade are most closely related based on shared morphological traits as well as genetic similarities. Excluded from this grouping is the orangutan, which is considered a member of the Asian clade of hominoids.
In contrast, grades are groupings that reflect levels of adaptation or overall similarity and not necessarily evolutionary relationships. An example of a grade would be placing orangutans, gorillas, bonobos, and chimpanzees into a group, and excluding humans. Grouping in this way is based on the superficial similarities of the apes in being large-bodied, having lots of body hair, living in tropical forests, climbing and sleeping in trees, and so on. According to these criteria, humans seem to be unusual in that we differ in our morphology, behavior, and ecology. Separating humans from the large-bodied apes is the system that was used historically. We now know that grouping orangutans, gorillas, bonobos, and chimpanzees and excluding humans does not accurately reflect our true evolutionary relationships. Since our goal in taxonomic classification is to organize animals to reflect their evolutionary relationships, we prefer to use clade classifications.
Types of Traits
When evaluating relationships between taxa, we use key traits that allow us to determine which species are most closely related to one another. Traits can be either ancestral or derived. Ancestral traits are those that a taxon has because it has inherited the trait from a distant ancestor. For example, all primates have body hair because we are mammals and all mammals share an ancestor hundreds of millions of years ago that had body hair. This trait has been passed down to all mammals from a shared ancestor, so all mammals alive today have body hair. Derived traits are those that have been more recently altered. This type of trait is most useful when we are trying to distinguish one group from another because derived traits tell us which taxa are more closely related to each other. For example, humans walk on two legs.The many adaptations that humans possess that allow us to move in this way evolved after humans split from the Genus Pan. This means that when we find fossil taxa that share derived traits for walking on two legs, we can conclude that they are likely more closely related to humans than to chimpanzees and bonobos.
There are a couple of other important points about ancestral and derived traits that will become apparent as we discuss primate diversity. First, the terms ancestral and derived are relative terms, meaning that a trait can be either one depending on the taxa being compared. For example, in the previous paragraph, body hair was used as an example for an ancestral trait among primates. All mammals have body hair because we share a distant ancestor who had this trait. The presence of body hair therefore doesn’t allow you to distinguish whether monkeys are more closely related to apes or lemurs because they all share this trait. However, if we are comparing mammals to birds and fish, then body hair becomes a derived trait of mammals. It evolved after mammals diverged from birds and fish, and it tells us that all mammals are more closely related to each other than they are to birds or fish.The second important point is that very often when one lineage splits into two, one taxon will stay more similar to the last common ancestor in retaining more ancestral traits, whereas the other lineage will usually become more different from the last common ancestor by developing more derived traits. This will become very apparent when we discuss the two suborders of primates, Strepsirrhini and Haplorrhini. When these two lineages diverged, strepsirrhines retained more ancestral traits (those present in the earliest primates) and haplorrhines developed more derived traits (became more different from ancestral primates).
There are two other types of traits that will be relevant to our discussions here: generalized and specialized traits. Generalized traits are those characteristics that are useful for a wide range of things. Having opposable thumbs that go in a different direction than the rest of your fingers is a very useful, generalized trait. You can hold a pen, grab a branch, peel a banana, or text your friends all thanks to your opposable thumbs! Specialized traits are those that have been modified for a specific purpose. These traits may not have a wide range of uses, but they will be very efficient at their job. Hooves in horses are a good example of a specialized trait: they allow horses to run quickly on the ground on all fours. You can think of generalized traits as a Swiss Army knife, useful for a wide range of tasks but not particularly good at any one of them. That is, if you’re in a bind, then a Swiss Army knife can be very useful to cut a rope or fix a loose screw, but if you were going to build furniture or fix a kitchen sink, then you’d want specialized tools for the job. As we will see, most primate traits tend to be generalized.
What Makes Something a Primate?
The Order Primates is distinguished from other groups of mammals in having a suite of characteristics. This means that there is no individual trait that you can use to instantly identify an animal as a primate; instead, you have to look for animals that possess a collection of traits. What this also means is that each individual trait we discuss may be found in nonprimates, but if you see an animal that has most or all of these traits, there is a good chance it is a primate.
Primates are most distinguishable from other organisms in traits related to our vision. Our Order relies on vision as a primary sense, which is reflected in many areas of our anatomy and behavior. All primates have eyes that face forward with convergent (overlapping) visual fields. So if you cover one eye with your hand, you can still see most of the room with your other one. This also means that we cannot see on the sides or behind us as well as some other animals can. In order to protect the sides of the eyes from the muscles we use for chewing, all primates have at least a postorbital bar, a bony ring around the outside of the eye (Figure 5.2). Primate taxa with more convergent eyes need extra protection, so animals with greater orbital convergence will have a postorbital plate or postorbital closure in addition to the bar (Figure 5.2).The postorbital bar is a derived trait of primates, appearing in our earliest ancestors.
Another distinctive trait of our Order is that many primates have trichromatic color vision, the ability to distinguish reds and yellows in addition to blues and greens. Birds, fish, and reptiles are tetrachromatic (they can see reds, yellows, blues, greens, and even ultraviolet), but most mammals, including some primates, are only dichromatic (they see only in blues and greens). It is thought that the nocturnal ancestors of mammals benefited from seeing better at night rather than in color, and so dichromacy is the ancestral condition for mammals. Trichromatic primates are known to use their color vision for all sorts of purposes: finding young leaves and ripe fruits, identifying other species, and evaluating signals of health and fertility.
The primate visual system uses a lot of energy, so primates have compensated by cutting back on other sensory systems, particularly our sense of smell. Compared to other mammals, primates have reduced snouts, another derived trait that appears even in the earliest primate ancestors. There is variation across primate taxa in how much snouts are reduced. Those with a better sense of smell usually have poorer vision than those with a relatively dull sense of smell. The reason for this is that all organisms have a limited amount of energy to spend on running our bodies, so we make evolutionary trade-offs, as energy spent on one trait cuts back on energy spent on another. So primates with better vision are spending more energy on vision and thus have a poorer smell (and shorter snout), and those who spend less energy on vision will have a better sense of smell (and a longer snout).
Primates also differ from other mammals in the size and complexity of our brains. On average, primates have brains that are twice as big for their body size when compared to other mammals. Not unexpectedly, the visual centers of the brain are larger in primates and the wiring is different from that in other animals, reflecting our reliance on this sense. The neocortex, which is used for higher functions like consciousness and language in humans, as well as sensory perception and spatial awareness, is also larger in primates relative to other animals. In nonprimates this part of the brain is often smooth, but in primates it is made up of many folds, which increase the surface area. It has been proposed that the more complex neocortex of primates is related to diet, with fruit-eating primates having larger relative brain sizes than leaf-eating primates, due to the more challenging cognitive demands required to find and process fruits (Clutton-Brock and Harvey 1980). An alternative hypothesis argues that larger brain size is necessary for navigating the complexities of primate social life, with larger brains occurring in species who live in bigger, more complex groups relative to those living in pairs or solitarily (Dunbar 1998). There seems to be support for both hypotheses, as large brains are a benefit under both sets of selective pressures.
Animals with large brains usually have extended life history patterns, and primates are no exception. Life history refers to the pace at which an organism grows, reproduces, and ages. Some animals grow very quickly and reproduce many offspring in a short time frame but do not live very long. Other animals grow slowly, reproduce few offspring, reproduce infrequently, and live a long time. Primates are all in the “slow lane” of life history patterns. Compared to animals of similar body size, primates grow and develop more slowly, have fewer offspring per pregnancy, reproduce less often, and live longer. Primates also invest heavily in each offspring. With a few exceptions, most primates only have one offspring at a time. A group of small-bodied monkeys in South America regularly give birth to twins, and some lemurs can give birth to multiple offspring at a time, but these primates are the exception rather than the rule. Primates also reproduce relatively infrequently. The fastest-reproducing primates will produce offspring about every six months, while the slowest, the orangutan, reproduces only once every seven to nine years. This very slow reproductive rate makes the orangutan the slowest-reproducing animal on the planet! Primates are also characterized by having long lifespans. The group that includes humans and large-bodied apes has the most extended life history patterns among all primates, with some large-bodied apes estimated to live up to 58 years in the wild (Robson et al. 2006).
Primates also differ from other animals in our hands and feet. The Order Primates is a largely arboreal taxonomic group, meaning that most primates spend a significant amount of their time in trees. As a result, the hands and feet of primates have evolved to move in a three-dimensional environment. Primates have the generalized trait of pentadactyly— possessing five digits (fingers and toes) on each limb. Many nonprimates, like dogs and horses, have fewer digits because they are specialized for high-speed, terrestrial (on the ground) running. Pentadactyly is also an ancestral trait, one that dates back to the earliest four-footed animals. Primates today have opposable thumbs and, with the exception of humans, opposable big toes (Figure 5.3). Opposable thumbs and toes are a derived trait that appeared in the earliest primate fossils about 55 million years ago. Having thumbs and big toes that go in a different direction from the rest of the fingers and toes allow primates to be excellent climbers in trees as well as to manipulate objects. Our ability to manipulate objects is further enhanced by the flattened nails on the backs of our fingers and toes that we possess in the place of the claws and hooves that many other mammals have. On the other side of our digits, we have sensitive tactile pads that allow us to have a fine sense of touch. Primates use this fine sense of touch for handling food and, in many species, grooming themselves and others. In primates, grooming is an important social currency, through which individuals forge and maintain social bonds.
Primate suite of traits |
Convergent eyes Postorbital bar Many have trichromatic color vision Short snouts Opposable thumbs and big toes Pentadactyly Flattened nails Tactile pads Highly arboreal Large brains Extended life histories Live in the tropics |
Lastly, primates are very social animals. All primates, even those that search for food alone, establish strong social networks within species. Unlike many animals, primates do not migrate: they stay in a relatively stable area for their whole life, often interacting with the same individuals for their long lives. The long-term relationships that primates form with others of their species lead to complex and fascinating social behaviors (see Chapter 6). Finally, nonhuman primates show a clear preference for tropical regions of the world. Most primates are found between the Tropic of Cancer and the Tropic of Capricorn, with only a few taxa living outside these regions. Figure 5.4 shows a summary of primate traits.
Key Traits Used to Distinguish Between Primate Taxa
When placing primate species into specific taxonomic groups, we focus on dental characteristics, behavioral adaptations, and locomotor adaptations. Differences in these characteristics across groups reflect constraints of evolutionary history as well as variation in adaptations.
Dental Characteristics
Teeth may not seem like the most exciting topic with which to start, but we can learn a tremendous amount about an organism from its teeth. First, teeth are vital to survival. Wild animals do not have the benefit of knives and forks; they rely on their teeth to process their food. Because of this, teeth of any species have evolved to reflect what that organism eats and therefore have a lot to tell us about their diet. Second, variation in tooth size, shape, and number reveals an organism’s evolutionary history. Some taxa have more teeth than others or different forms of teeth. Furthermore, differences in teeth between males and females can tell us about competition over mates (see Chapter 6). Lastly, teeth are overly represented in the fossil record. Enamel is hard, and there is little meat on jaws so carnivores and scavengers often leave them behind. Sometimes, the only remains we have from an extinct taxon is its teeth!
Like other mammals, primates are heterodont: they have multiple types of teeth that are used for different purposes. We have incisors for slicing; premolars and molars for grinding up our food; and canines, which most primates (not humans) use as weapons against predators and each other. The sizes of canines vary across species and can often be sexually dimorphic, with males tending to have larger canines than females. Some nonhuman primates hone, or sharpen, their canines by gnashing the teeth together to sharpen the sides. The upper canine sharpens on the first lower premolar and the lower canine sharpens on the front of the upper canine. As canines get larger, they require a space to fit in order for the jaws to close. This space between the teeth is called a diastema (Figure 5.5).
We use a dental formula to specify how many incisors, canines, premolars, and molars are in each quadrant of the mouth (half of the top or bottom). For example, Figure 5.6 shows half of the lower teeth of a human. You can see that in half of the mandible, there are two incisors, one canine, two premolars, and three molars. This dental formula is written as 2:1:2:3. (The first number represents the number of incisors, followed by the number of canines, premolars, and molars).
To determine the dental formula, you need to be able to identify the different types of teeth. You can recognize incisors because they often look like spatulas with a flat, blade-like surface. Premolars and molars can be differentiated by the number of cusps that they have. Cusps are the bumps that you can feel with your tongue on the surface of your back teeth. Premolars are smaller than molars and, in primates, often have one or two cusps on them. Molars are bigger, providing a larger chewing surface, and have more cusps. Depending on the species and whether you’re looking at upper or lower teeth, primate molars can have between three and five cusps. Molar cusps can also vary between taxa in how they are arranged; you will learn more about this later in this chapter. Canines are often easy to distinguish because, in most taxa, they are much longer and more conical than the other teeth.
Teeth also directly reflect an organism’s diet. Primates are known to eat a wide range of plant parts, insects, gums, and, rarely, meat. While all primates eat a variety of foods, what differs among primates are the proportions of each of these food items in the diet. That is, two primates living in the same forest may be eating the same foods but in vastly different proportions, and so we would categorize them as different dietary types. The most common dietary types among primates are those whose diets consist primarily of fruit (frugivores), those who eat mostly insects (insectivores), and those who eat primarily leaves (folivores). A few primate taxa are gummivores, specializing in eating gums and saps, but we will only focus on the adaptations found in the three primary dietary groups.
Frugivores
Plants want animals to eat their fruits because, in doing so, animals eat the seeds of the fruit and then disperse them far away from the parent plant. Therefore, plants often “advertise” fruits by making them colorful and easy to spot, full of easy-to-digest sugars that make them taste good and, often, easy to chew and digest (not being too fibrous or tough). For these reasons, frugivores often do not need a lot of specialized traits to consume a diet rich in fruits (Figure 5.7). Their molars usually have a broad chewing surface with low, rounded cusps (referred to as bunodont molars). Frugivores have large incisors for slicing through the outer coatings on fruit, and they tend to have stomachs, colons, and small intestines that are intermediate in terms of size and complexity between insectivores and folivores (Chivers and Hladik 1980). They are also usually of intermediate body size between the other two dietary types. Because fruit does not contain protein, frugivores must supplement their diet with protein from insects, leaves, and/or seeds. Frugivores who get protein by eating seeds evolved to have thicker enamel on their teeth to protect them from excessive wear.
Insectivores
While insects can be difficult to find and catch, they are easy to chew and digest. As a result, insectivorous primates usually have small molars with pointed cusps to puncture the exoskeleton of the insects (Figure 5.8), and they have simple stomachs and colons with a long small intestine to process the insects. Nutritionally, insects provide a lot of protein and fat but are not plentiful enough in the environment to support large-bodied animals, so insectivores are usually the smallest of the primates.
Folivores
Plants rely on leaves to get energy from the sun, so plants do not want animals to eat their leaves (unlike their fruit). As a result, plants evolved to try to discourage animals from eating their leaves. Leaves often carry toxins, taste bitter, are very fibrous and difficult to chew, and are made of large cellulose molecules that are difficult to break down into usable sugars. Thus, animals who eat leaves need a lot of specialized traits (Figure 5.9). Folivorous primates have broad molars with high, sharp cusps connected by shearing crests. These molar traits allow folivores to physically break down fibrous leaves when chewing. Folivores then chemically break down cellulose molecules into usable energy. To do this, some folivores have complex stomachs with multiple compartments, while others have large, long intestines and special gut bacteria that can break up cellulose. Folivores are usually the largest bodied of all primates, and they tend to spend a large portion of their day digesting their food, so they are less active than frugivores or insectivores.
Behavioral Adaptations
Since Chapter 6 is dedicated to primate behavior, we will only briefly discuss variations in activity patterns, social grouping, and habitat use. Primate groups differ in activity patterns: whether they are active during the day (diurnal), at night (nocturnal), or through the 24-hour period (cathemeral). Primate taxa vary in social groupings: some are primarily solitary, others live in pairs, and still others live in groups of varying sizes and compositions. Lastly, some taxa are primarily arboreal while others are more terrestrial.
Locomotor Adaptations
Finally, primate groups vary in their adaptations for different forms of locomotion, or how they move around. Living primates are known to move by vertical clinging and leaping, quadrupedalism, brachiation, and bipedalism.
Vertical clinging and leaping is when an animal grasps a vertical branch with its body upright, pushes off with long hind legs, and then lands on another vertical support branch (Figure 5.10a). Animals who move in this way usually have longer legs than arms, long fingers and toes, and smaller bodies. Vertical clinger leapers also tend to have elongated ankle bones, which serve as a lever to help them push off with their legs and leap to another branch (Figure 5.10b).
Quadrupedalism, walking on all fours, is the most common form of locomotion among primates. Quadrupedal animals usually have legs and arms that are about the same length and a tail for balance. Arboreal quadrupeds (Figure 5.11a) usually have shorter arms and legs and longer tails, while terrestrial quadrupeds (Figure 5.11b) have longer arms and legs and, often, shorter tails. These differences relate to the lower center of gravity needed by arboreal quadrupeds for balance in trees and the longer tail required for better balance when moving along the tops of branches. Terrestrial quadrupeds have longer limbs to help them cover more distance more efficiently.
The third form of locomotion seen in primates is brachiation, the way of moving you used if you played on “monkey bars” as a child. Brachiation involves swinging below branches by the hands (Figure 5.12a). To be an efficient brachiator, a primate needs to have longer arms than legs, flexible shoulders and wrists, a short lower back, and no tail (Figure 5.12b). Some primates move via semi-brachiation, in which they swing below branches but do not have all of the same specializations as brachiators. Semi-brachiators have flexible shoulders, but their arms and legs are about the same length, which is useful because they are quadrupedal when on the ground. They also use long prehensile tails as a third limb when swinging (Figure 5.13). The underside of the tail has a tactile pad, resembling your fingerprints, for better grip.
Lastly, humans move around on two feet, called bipedalism. Some nonhuman primates will occasionally travel on two feet but do so awkwardly and never for long distances. Among mammals, only humans have evolved to walk with a striding gait on two legs as a primary form of locomotion.
Primate Diversity
As we begin exploring the different taxa of primates, it is important to keep in mind the hierarchical nature of taxonomic classification and how this relates to the key characteristics that will be covered. Figure 5.14 summarizes the major taxonomic groups of primates that you will learn about here. If you locate humans on the chart, you can trace our classification and see all of the categories getting more inclusive as you work your way up to the Order Primates. This means that humans will have the key traits of each of those groups. It is a good idea to refer to the figure to orient yourself as we discuss each taxon.
Suborder Strepsirrhini
The Order Primates is subdivided into Suborder Strepsirrhini and Suborder Haplorrhini, which, according to molecular estimates, split about 70–80 million years ago (Pozzi et al. 2014). The strepsirrhines include the groups commonly called lemurs, lorises, and galagos (Figure 5.15). Strepsirrhines differ from haplorrhines in many ways, most of which involve retaining ancestral traits from the earliest primates. Strepsirrhines do have two key derived traits that evolved after they diverged from the haplorrhines: the grooming claw (Figure 5.16) on the second digit of each foot, and the tooth comb (or dental comb) located on the lower, front teeth (Figure 5.17). In most strepsirrhines, there are six teeth in the toothcomb—four incisors and two canines. Other than the tooth comb, the teeth of strepsirrhines are fairly simple and are neither large or distinctive relative to haplorrhines.
Compared to haplorrhines, strepsirrhines rely more on nonvisual senses. Strepsirrhines get their name because they have wet noses (rhinariums) like cats and dogs, a trait that, along with a longer snout, reflect strepsirrhines’ greater reliance on olfaction relative to haplorrhines. Many strepsirrhines use scent marking, including rubbing scent glands or urine on objects in the environment to communicate with others. Additionally, many strepsirrhines have mobile ears that they use to locate insect prey and predators. While strepsirrhines have a better sense of smell than haplorrhines, their visual adaptations are more ancestral. Strepsirrhines have less convergent eyes than haplorrhines and therefore all have postorbital bars, whereas haplorrhines have full postorbital closure (see Figure 5.2). All strepsirrhines have a tapetum lucidum, a reflective layer at the back of the eye that reflects light and thereby enhances the ability to see in low-light conditions. It is the same layer that causes your dog or cat to have “yellow eye” when you take photos of them with the flash on. This is a trait thought to be ancestral among mammals as a whole.
Strepsirrhines also differ from haplorrhines in some aspects of their ecology and behavior. The majority of strepsirrhines are solitary, traveling alone to search for food; a few taxa are more social. Most strepsirrhines are also nocturnal and arboreal. Strepsirrhines are, on average, smaller than haplorrhines, and so many of them have a diet consisting of insects and fruit, with few taxa eating primarily leaves. Lastly, most strepsirrhines are good at leaping, with several taxa specialized for vertical clinging and leaping. In fact, among primates, all but one of the vertical clinger leapers belong to the Suborder Strepsirrhini.
Strepsirrhines can be found all across Asia, Africa, and on the island of Madagascar (Figure 5.18). The Suborder Strepsirrhini is divided into two groups: (1) the lemurs of Madagascar and (2) the lorises, pottos, and galagos of Africa and Asia. By molecular estimates, these two groups split about 65 million years ago (Pozzi et al. 2014).
Lemurs of Madagascar
Madagascar is an island off the east coast of Africa, and it is roughly the size of California, Oregon, and Washington combined. It has been separated from Africa for about 130 million years and from India for about 85 million years, which means it was already an island when strepsirrhines got there approximately 60–70 million years ago. Only a few mammal species ever reached Madagascar, and so when lemurs arrived they were able to flourish into a variety of forms.
The lemurs of Madagascar are much more diverse compared to their mainland counterparts, the lorises and galagos. While many Malagasy strepsirrhines are nocturnal, plenty of others are diurnal or cathemeral. They range in body size from the smallest of all primates, the mouse lemur, some species of which weigh a little over an ounce (see Figure 5.15), up to the largest of all strepsirrhines, the indri, which weighs up to about 20 pounds (Figure 5.19). Lemurs include species that are insectivorous, frugivorous, and folivorous. A couple of members of this group have unusual diets for primates, including the gummivorous fork-marked and bamboo lemurs, who are able to metabolize the cyanide in bamboo. The most unique lemur is the aye-aye (depicted in Figure 5.15). This nocturnal lemur has rodent-like front teeth that grow continuously and a long-bony middle finger that it uses to fish grubs out of wood. It has a very large brain compared to other strepsirrhines, which it fuels with a diet that includes bird’s eggs and other animal matter. Based on genetic estimates and morphological studies, it is believed that aye-ayes were the first lemurs to separate from all other strepsirrhines and to evolve on their own since strepsirrhines arrived in Madagascar (Matsui et al. 2009).
Lemurs are also diverse in terms of social behavior: Many lemurs are solitary foragers, some live in pairs, others in small groups, still others in larger groups, and some, like the red-ruffed lemur, live in unique and complex social groups (Vasey 2006). Lemurs include some of the best vertical clingers and leapers, and while many lemurs are quadrupedal, even the quadrupedal lemurs are quite adept at leaping. Malagasy strepsirrhines also exhibit a few unusual traits. They are highly seasonal breeders, often mating only during a short window once a year (Wright 1999). Female ring-tailed lemurs, for example, come into estrus one day a year for a mere six hours. Unlike most primates, where males are typically large and dominant, Malagasy strepsirrhines feature socially dominant females that are similar in size to males and have priority access to resources.
Lorises, Pottos, and Galagos of Asia and Africa
Unlike the lemurs of Madagascar, lorises, pottos, and galagos live in areas where they share their environments with monkeys and apes, who often eat similar foods. Lorises live across South and Southeast Asia, while pottos and galagos live across Central Africa. Because of competition with larger-bodied monkeys and apes, mainland strepsirrhines are more restricted in the niches they can fill in their environments and so are less diverse than the lemurs.
The strepsirrhines of Africa and Asia are all nocturnal and solitary, with little variation in body size and diet. For the most part, the diet of lorises, pottos, and galagos consists of fruits and insects. A couple of species eat more gum, but overall the diet of this group is narrow when compared to the Malagasy lemurs. Lorises (Figure 5.20) and pottos are known for being slow, quadrupedal climbers, moving quietly through the forests to avoid being detected by predators. These strepsirrhines have developed additional defenses against predators. Lorises, for example, eat a lot of caterpillars, which makes their saliva slightly toxic. Loris mothers bathe their young in this toxic saliva, making the babies unappealing to predators. In comparison to the slow-moving lorises and pottos, galagos are active quadrupedal runners and leapers that scurry about the forests at night. Galagos make distinctive calls that sound like a baby crying, which has led to their nickname “bushbabies.” Figure 5.21 summarizes the key differences between these two groups of strepsirrhines.
|
Lemurs |
Lorises, Pottos, and Galagos |
Geographic range |
Madagascar |
South and Southeast Asia Central Africa |
Activity patterns |
Diurnal, nocturnal, or cathemeral |
Nocturnal |
Dietary types |
Insectivore, frugivore, or folivore |
Insectivore, frugivore |
Social groupings |
Solitary, pairs, or small to large groups |
Solitary |
Forms of locomotion |
Vertical clinger leapers, quadrupedal |
Slow quadrupedal climbers and active quadrupedal runners |
Suborder Haplorrhini
When the two primate suborders split from one another, strepsirrhines retained more ancestral traits while haplorrhines developed more derived traits, which are discussed below.
As mentioned earlier, haplorrhines have better vision than strepsirrhines. This is demonstrated by the full postorbital closure protecting the more convergent eyes that haplorrhines possess (with one exception seen in Figure 5.2). Most haplorrhines are trichromatic, and all have a fovea, a depression in the retina at the back of the eye containing concentrations of cells that allows them to see things very close up in great detail. The heavier reliance on vision over olfaction is also reflected in the shorter snouts ending with the dry nose (no rhinarium) of haplorrhines. All but two genera of living haplorrhines are active during the day, so this group lacks the tapetum lucidum that is so useful to nocturnal species. On average, haplorrhines also have larger brains relative to their body size when compared with strepsirrhines.
The Haplorrhini differ from the Strepsirrhini in their ecology and behavior as well. Haplorrhines are generally larger than strepsirrhines, and they tend to be folivorous and frugivorous. This dietary difference is reflected in the teeth of haplorrhines, which are broader with more surface area for chewing. The larger body size of this taxon also influences locomotion. Only one haplorrhine is a vertical clinger and leaper. Most members of this suborder are quadrupedal, with one subgroup specialized for brachiation. A few haplorrhine taxa are monomorphic, meaning males and females are the same size, but many members of this group show moderate to high sexual dimorphism in body size and canine size. Haplorrhines also differ in social behavior. All but two haplorrhines live in groups, which is very different from the primarily solitary strepsirrhines. Differences between the two suborders are summarized in Figure 5.22.
|
Suborder Strepsirrhini |
Suborder Haplorrhini |
Sensory adaptations |
Rhinarium Longer snout Eyes less convergent Postorbital bar Tapetum lucidum Mobile ears |
No rhinarium Short snout Eyes more convergent Postorbital plate No tapetum lucidum Many are trichromatic Fovea |
Dietary differences |
Mostly insectivores and frugivores, few folivores |
Few insectivores, mostly frugivores and folivores |
Activity patterns and Ecology |
Mostly nocturnal, few diurnal or cathemeral Almost entirely arboreal |
Only two are nocturnal, rest are diurnal Many arboreal taxa, also many terrestrial taxa |
Social groupings |
Mostly solitary, some pairs, small to large groups |
Only two are solitary, all others live in pairs, small to very large groups |
Sexual dimorphism |
Minimal to none |
Few taxa have little/none, many taxa show moderate to high dimorphism |
Suborder Haplorrhini is divided into three infraorders: Tarsiiformes, which includes the tarsiers of Asia; Platyrrhini, which includes the monkeys of Central and South America; and Catarrhini, a group that includes the monkeys of Asia and Africa, apes, and humans. According to molecular estimates, tarsiers split from the other haplorrhines close to 70 million years ago, and platyrrhines split from catarrhines close to 46 million years ago (Pozzi et al. 2014).
Infraorder Tarsiiformes of Asia
Today, the Infraorder Tarsiiformes includes only one genus, Tarsius (Figure 5.23). Tarsiers are small-bodied primates that live in Southeast Asian forests (Figure 5.24) and possess an unusual collection of traits that have led to some debate about their position in the primate taxonomy. They are widely considered members of the haplorrhine group because they share several derived traits with monkeys, apes, and humans, including dry noses, a fovea, not having a tapetum lucidum, and eyes that are more convergent. Tarsiers also have some traits that are more like strepsirrhines and some that are unique. Tarsiers are the only haplorrhine that are specialized vertical clinger leapers, a form of locomotion only otherwise seen in some strepsirrhines. Tarsiers actually get their name because their ankle (tarsal) bones are elongated to provide a lever for vertical clinging and leaping. Tarsiiformes are also small, with most species weighing between 100 and 150 grams. Like strepsirrhines, tarsiers are nocturnal, but because they lack a tapetum lucidum, tarsiers compensate by having enormous eyes. In fact, each eye of a tarsier is larger than its brain. These large eyes allow enough light in for tarsiers to still be able to see well at night without the reflecting layer in their eyes. To protect their large eyes, tarsiers have a partially closed postorbital plate that appears somewhat intermediate between the postorbital bar of strepsirrhines and the full postorbital closure of other haplorrhines (Figure 5.25). Tarsiers have different dental formulas on their upper and lower teeth. On the top, the dental formula is 2:1:3:3, but on the bottom it is 1:1:3:3. Other unusual traits of tarsiers include having two grooming claws on each foot and the ability to rotate their heads around 180 degrees, a trait useful in locating insect prey. The tarsier diet is considered faunivorous because it consists entirely of animal matter, making them the only primate not to eat any vegetation. They are only one of two living haplorrhines to be solitary, the other being the orangutan. Most tarsiers are not sexually dimorphic, like strepsirrhines, although males of a few species are slightly larger than females.
Two alternative classifications have emerged due to the unusual mix of traits that tarsiers have. Historically, tarsiers were grouped with lemurs, lorises, and galagos into a suborder called Prosimii. This classification was based on tarsiers, lemurs, lorises, and galagos all having grooming claws and similar lifestyles. Monkeys, apes, and humans were then separated into a suborder called the Anthropoidea. These suborder groupings were based on grade rather than clade. Today, most people use Suborders Strepsirrhini and Haplorrhini, which are clade groupings based on the derived traits that tarsiers share with monkeys, apes, and humans. The Strepsirrhini/Haplorrhini dichotomy is also supported by the genetic evidence that indicates tarsiers are more closely related to monkeys, apes, and humans (Jameson et al. 2011). Figure 5.26 summarizes the unusual mix of traits seen in tarsiers.
Like Strepsirrhini |
Unique |
Like Haplorrhini |
Very small Nocturnal Highly insectivorous Solitary Vertical clinger-leapers Little/no sexual dimorphism |
Two grooming claws 2:1:3:3/1:1:3:3 dental formula Do not eat vegetation Can rotate their heads nearly 180 degrees |
Almost full PO closure More convergent eyes No tapetum lucidum No rhinarium Genetic evidence Fovea |
Infraorder Platyrrhini of Central and South America
The platyrrhines are the only nonhuman primates in Central and South America (Figure 5.27) and so, like the lemurs of Madagascar, have diversified into a variety of forms in the absence of competition. Infraorder Platyrrhini get their name from their distinctive nose shape. “Platy” means flat and “rhini” refers to noses, and, indeed, platyrrhines have noses that are flat and wide, with nostrils that are far apart, facing outward, and usually round in shape (Figure 5.28). This nose shape is very different from what we see in catarrhines.
On average, platyrrhines are smaller and less sexually dimorphic than catarrhines, and they have retained the more ancestral primate dental formula of 2:1:3:3. Platyrrhines are all highly arboreal, whereas many catarrhines spend significant time on the ground. The monkeys in Central and South America also differ in having less well-developed vision. This is reflected in the wiring in the visual system of the brain as well as in their polymorphic color vision. The genes that enable individuals to distinguish reds and yellows from blues and greens are on the X chromosome. Different genes code for being able to see different wavelengths of light so to distinguish between them you need to be heterozygous for seeing color. The X chromosomes of platyrrhines each carry the genes for seeing one wavelength, so male platyrrhines (with only one X chromosome) are always dichromatic. Female platyrrhines can be dichromatic (if they are homozygous for one version of the color vision gene) or trichromatic (if they are heterozygous) (Kawamura et al. 2012). We currently know of two exceptions to this pattern among platyrrhines. Nocturnal owl monkeys are monochromatic, meaning that they cannot distinguish any colors. The other exception are howler monkeys, which have evolved to have two color vision genes on each X chromosome. This means that both male and female howler monkeys are able to see reds and yellows. By contrast, catarrhine males and females are all trichromatic.
Platyrrhines include the smallest of the monkeys, the marmosets and tamarins (Figure 5.29), all of which weigh less than one kilogram and live in cooperative family groups, wherein usually only one female reproduces and everyone else helps carry and raise the offspring. They are unusual primates in that they regularly produce twins. Marmosets and tamarins largely eat gums and saps, so these monkeys have evolved claw-like nails that enable them to cling to the sides of tree trunks like squirrels as well as special teeth that allow them to gnaw through bark. Except for the Goeldi’s monkey, these small monkeys have one fewer molar than other platyrrhines, giving them a dental formula of 2:1:3:2.
The largest platyrrhines are a family that include spider monkeys, woolly spider monkeys, woolly monkeys, and howler monkeys (Figure 5.30). These monkeys can weigh up to 9–15 kg and have evolved prehensile tails that can hold their entire body weight. It is among this group that we see semi-brachiators, like the spider monkey (see Figure 5.13). To make them more efficient in this form of locomotion, spider monkeys evolved to not have thumbs so that their hands work more like hooks that can easily let go of branches while swinging. Howler monkeys are another well-known member of this group, earning their name due to their loud calls, which can be heard miles away. To make these loud vocalizations, howler monkeys have a specialized vocal system that includes a large larynx and hyoid bone. Howler monkeys are the most folivorous of the platyrrhines and are known for spending a large portion of their day digesting their food.
There are many other monkeys in Central and South America, including the gregarious capuchins (see Figure 5.28) and squirrel monkeys, the pair-living titi monkeys, and the nocturnal owl monkeys. There are also the seed-eating saki monkeys and uakaris. In many areas across Central and South America, multiple species of platyrrhines share the forests, with some even traveling together in association. According to molecular evidence, the diversity of platyrrhines that we see today seems to have originated about 25 million years ago (Schneider and Sampaio 2015). Figure 5.31 summarizes the key traits of platyrrhines relative to the other infraorders of Haplorrhini.
Platyrrhini traits |
Flat nose with rounded nostrils pointing to the side Highly arboreal Less sexually dimorphic on average 2:1:3:3 dental formula* Polymorphic color vision* |
Infraorder Catarrhini of Asia and Africa
Infraorder Catarrhini includes Superfamily Cercopithecoidea (the monkeys of Africa and Asia) and Superfamily Hominoidea (apes and humans). Nonhuman catarrhines are found all over Africa and South and Southeast Asia, with some being found as far north as Japan. The most northerly and southerly catarrhines are cercopithecoid monkeys. In contrast, apes are less tolerant of drier, more seasonal environments and so have a relatively restricted geographic range.
Relative to other haplorrhine infraorders, catarrhines are distinguished by several characteristics. Catarrhines have a distinctive nose shape, with teardrop-shaped nostrils that are close together and point downward (Figure 5.32) and one fewer premolar than most other primates, giving us a dental formula of 2:1:2:3 (Figure 5.33). On average, catarrhines are the largest and most sexually dimorphic of all primates. Gorillas are the largest living primates, with males weighing up to 220 kg. The most sexually dimorphic of all primates are mandrills. Mandrill males not only have much more vibrant coloration than mandrill females but also have larger canines and can weigh up to three times more (Setchell et al. 2001). The larger body size of catarrhines is related to the more terrestrial lifestyle of many members of this infraorder. In fact, the most terrestrial of living primates can be found in this group. Among all primates, vision is the most developed in catarrhines. Catarrhines independently evolved the same adaptation as howler monkeys in having each X chromosome with genes to distinguish both reds and yellows, so all male and female catarrhines are trichromatic, which is useful for these diurnal primates.
The two superfamilies of catarrhines—Superfamily Cercopithecoidea, the monkeys of Africa and Asia, and Superfamily Hominoidea, which includes apes and humans—are believed to have split about 32 million years ago based on molecular evidence (Pozzi et al. 2014). This fits with the fossil record, which shows evidence of these lineages by about 25 million years ago (see Chapter 8).
Superfamily Cercopithecoidea of Africa and Asia
Compared to hominoids, cercopithecoids have an ancestral quadrupedal body plan with two key derived traits. The first derived trait of cercopithecoids is their bilophodont molars (“bi” meaning two, “loph” referring to ridge, and “dont” meaning tooth). If you refer back to Figure 5.33, you will see how the molars of cercopithecoids have four cusps arranged in a square pattern and have two ridges connecting them. It is thought that this molar enabled these monkeys to eat a wide range of foods, thus allowing them to live in habitats that apes cannot. The other key derived trait that all cercopithecoids share is having ischial callosities (Figure 5.34). The ischium is the part of your pelvis that you are sitting on right now (see Appendix A: Osteology). In cercopithecoids, this part of the pelvis has a flattened surface that, in living animals, has callused skin over it. These function as seat pads for cercopithecoids, who often sit above branches when feeding and resting.
Cercopithecoid monkeys are the most geographically widespread group of nonhuman primates (Figure 5.35). Since their divergence from hominoids, this monkey group has increased in numbers and diversity due, in part, to their fast reproductive rates. On average, cercopithecoids will reproduce every one to two years, whereas hominoids will reproduce once every four to nine years, depending on the taxon.
Cercopithecoidea is split into two groups, the leaf monkeys and the cheek-pouch monkeys. Both groups coexist in Asia and Africa; however, the majority of leaf monkey species live in Asia with only a few taxa in Africa. In contrast, only one genus of cheek-pouch monkey lives in Asia, and all the rest of them in Africa. As you can probably guess based on their names, the two groups differ in terms of diet. Leaf monkeys are primarily folivores, with some species eating a significant amount of seeds. Cheek-pouch monkeys tend to be more frugivorous or omnivorous, with one taxon, geladas, eating primarily grasses. The two groups also differ in some other interesting ways. Leaf monkeys tend to produce infants with natal coats—infants whose fur is a completely different color from their parents (Figure 5.36). Leaf monkeys are also known for having odd noses (Figure 5.37), and so they are sometimes called “odd-nosed monkeys.” Cheek-pouch monkeys are able to pack food into their cheek pouches (Figure 5.38), thus allowing them to move to a location safe from predators or aggressive individuals of their own species where they can eat in peace.
Superfamily Hominoidea of Africa and Asia
Superfamily Hominoidea of Africa and Asia (Figure 5.39) includes the largest of the living primates: apes and humans. Whereas cercopithecoid monkeys have bilophodont molars, hominoids have the more ancestral Y-5 molars, which feature five cusps separated by a “Y”-shaped groove pattern (see Figure 5.33). The Y-5 molar was present in the common ancestors of hominoids and cercopithecoids, thus it is the more ancestral molar pattern of the two. Hominoids differ the most from other primates in our body plans, due to the unique form of locomotion that hominoids are adapted for: brachiation (Figure 5.40).
To successfully swing below branches, many changes to the body needed to occur. Hominoid arms are much longer than the legs to increase reach, and the lower back is shorter and less flexible to increase control when swinging. The torso, shoulders, and arms of hominoids have evolved to increase range of motion and flexibility (see again Figure 5.12). The clavicle, or collar bone, is longer to stabilize the shoulder joint out to the side, thus enabling us to rotate our arms 360 degrees. Hominoid rib cages are wider side to side and shallower front to back than those of cercopithecoids and we do not have tails, as tails are useful for balance when running on all fours but generally not useful while swinging. Hominoids also have modified ulnae, one of the two bones in the forearm (see Appendix A: Osteology). At the elbow end of the ulna, hominoids have a short olecranon process, which allows for improved extension in our arms. At the wrist end of the ulna, hominoids have a short styloid process, which enables us to have very flexible wrists, a trait critical for swinging. Both the olecranon process and styloid process are long in quadrupedal animals who carry much of their weight on their forelimbs when traveling and who therefore need greater stability rather than flexibility in those joints.
|
Quadrupedalism |
Brachiation |
Arm length vs. leg length |
About equal |
Arms are longer |
Shoulder position |
More on the front |
Out to the side |
Ribcage shape |
Deep front-to-back Narrow side-to-side |
Shallow front-to-back Wide side-to-side |
Length of lower back |
Long |
Short |
Collar bone length |
Short |
Long |
Ulnar olecranon process |
Long |
Short |
Ulnar styloid process |
Long |
Short |
Tail |
Short to long |
None |
Apes and humans also differ from other primates in behavior and life history characteristics. Hominoids all seem to show some degree of female dispersal at sexual maturity but, as you will learn in Chapter 6, it is more common that males leave. Some apes show males dispersing in addition to females, but the hominoid tendency for female dispersal is a bit unusual among primates. Our superfamily is also characterized by the most extended life histories of all primates. All members of this group take a long time to grow and reproduce much less frequently compared to cercopithecoids. The slow pace of this life history is likely related to why hominoids have decreased in diversity since they first evolved. Figure 5.41 summarizes the key traits of Infraorder Catarrhini and its two superfamilies. Today, there are only five types of hominoids left: gibbons and siamangs, orangutans, gorillas, chimpanzees and bonobos, and humans.
Infraorder Catarrhini |
|
Downward facing, tear-drop shaped nostrils, close together Arboreal and more terrestrial taxa On average, largest primates On average, most sexually dimorphic taxonomic group 2:1:2:3 dental formula All trichromatic |
Superfamily Cercopithecoidea |
Superfamily Hominoidea |
Wide geographic distribution Bilophodont molars Ischial callosities Reproduce every 1–2 years |
Tropical forests of Africa and Asia Y-5 molars Adaptations for brachiation Reproduce every 4–9 years |
Family Hylobatidae of Southeast Asia
The number of genera in this group has been changing in recent years, but the taxa broadly encompasses gibbons and siamangs. Both are found across Southeast Asian tropical forests. Gibbons weigh, on average, about 13 pounds and tend to be more frugivorous, whereas siamangs are larger than gibbons and also more folivorous. Unlike the larger-bodied apes (orangutans, chimps, bonobos, and gorillas) who make nests to sleep in every night, gibbons and siamangs will develop callused patches on their ischium resembling ischial callosities. Gibbon species are quite variable in their coloration and markings, while siamangs are all black with big throat sacs that are used in their exuberant vocalizations (Figure 5.42). Both gibbons and siamangs live in pairs with very little sexual dimorphism, although males and females do differ in coloration in some gibbon species.
Pongo of Southeast Asia
The Genus Pongo refers to orangutans. These large red apes are found in Southeast Asia, with the two well-known species each living on the islands of Borneo and Sumatra. A third, very rare species, was recently discovered in Southern Sumatra (Nater et al. 2017). Orangutans are highly frugivorous but will supplement their diet with leaves and bark when fruit is less available. As mentioned earlier, orangutans are the only diurnal, solitary taxon among primates and are extremely slow to reproduce, producing only one offspring about every seven to nine years. They are highly sexually dimorphic (Figure 5.43 a and b), with fully developed, “flanged” males being approximately twice the size of females. These males have large throat sacs; long, shaggy coats; and cheek flanges. The skulls of male orangutans often feature a sagittal crest, which is believed to function as additional attachment area for chewing muscles as well as a trait used in sexual competition (Balolia, Soligo, and Wood 2017). An unusual feature of orangutan biology is male bimaturism. Male orangutans are known to delay maturation until one of the more dominant, flanged males disappears. The males that delay maturation are called “unflanged” males, and they can remain in this state for their entire life. Unflanged males resemble females in their size and appearance and will sneak copulations with females while avoiding the bigger, flanged males. Flanged and unflanged male orangutans represent alternative reproductive strategies, both of which successfully produce offspring (Utami et al. 2002).
Gorilla of Africa
There are several species of gorillas that can be found across Central Africa. Gorilla males, like orangutan males, are about twice the size of female gorillas (Figure 5.44a and b). When on the ground, gorillas use a form of quadrupedalism called knuckle-walking, wherein the fingers are curled under and the weight is carried on the knuckles. Male gorillas have a large sagittal crest and large canines compared with females. Adult male gorillas are often called “silverbacks” because when they reach about twelve to thirteen years old, the hair on their backs turns silvery gray. Gorillas typically live in groups of one male and several females. Gorillas are considered folivorous, although some species can be more frugivorous depending on fruit seasonality (Remis 1997).
Pan of Africa
The Genus Pan includes two species: Pan troglodytes (the common chimpanzee) and Pan paniscus (the bonobo). These species are separated by the Congo River, with chimpanzees ranging across West and Central Africa and bonobos located in a restricted area south of the Congo River. Chimpanzees and bonobos both have broad, largely frugivorous diets.The two species differ morphologically in that bonobos are slightly smaller, have their hair parted down the middle of their foreheads, and are born with dark faces (Figure 5.45). In contrast, chimpanzees do not have the distinctive parted hair and are born with light faces that darken as they mature (Figure 5.46). Chimpanzees and bonobos live in a grouping called a fission-fusion community, which you will learn more about in Chapter 6. Both species are moderately sexually dimorphic, with males about 20% larger than females. When on the ground, chimpanzees and bonobos knuckle-walk like gorillas do.
Homo
The last member of the Hominoidea to discuss is our own taxon, Genus Homo. Later chapters will discuss the many extinct species of Homo, but today there is only one living species of Homo, our own species, sapiens. While it is interesting to focus on how humans differ from apes in many aspects of our morphology, behavior, and life history, one objective of this chapter, and of biological anthropology in general, is to understand our place in nature. This means looking for aspects of human biology that link us to the taxonomic diversity we have discussed. To that end, here we will focus on similarities humans share with other hominoids.
Like other hominoids, humans lack a tail and possess upper-body adaptations for brachiation. While our lower body has been modified for a bipedal gait, we are still able to swing from branches and throw a baseball, all thanks to our mobile shoulder joint. Humans, like other hominoids, also have a Y-5 cusp pattern on our molars. All hominoids, including humans, have an extended life history, taking time to grow and develop, and reproducing slowly over a long life span. Lastly, while humans show a great deal of variation across cultures, many human societies show tendencies for female dispersal (Burton et al. 1996).
Among the hominoids, humans show particular affinities with other members of the African Clade, Pan and Gorilla. Humans share over 96% of our DNA with gorillas (Scally et al. 2012), and over 98% with Pan (Ebersberger et al. 2002). Even without this strong genetic evidence, the African Clade of hominoids share many morphological similarities, including having wide-set eye sockets and backward-sweeping cheekbones. Today, Pan and Gorilla knuckle-walk when on the ground, and it has been suggested the last common ancestor of chimpanzees, bonobos, gorillas, and humans did as well (Richmond, Begun, and Strait 2001). Further, humans, chimpanzees, and bonobos all live in fission-fusion social groups characterized by shared behaviors, like male cooperation in hunting and territoriality, as well as tool use.
Special Topic: Primates in Culture and Religion
One of the best parts of teaching anthropology for me is getting to spend time watching primates at zoos. What I also find interesting is watching people watch primates. I have very often heard a parent and child walk up to a chimpanzee enclosure and exclaim “Look at the monkeys!” The parent and child often don’t know that a chimpanzee is not a monkey, nor are they likely to know that chimpanzees share more than 98% of their DNA with us. What strikes me as significant is that, although most people do not know the difference between a monkey, an ape, and a lemur, they nonetheless recognize something in the animals as being similar to themselves. In fact, recognition of similarities between humans and other primates is very ancient, dating back far earlier than Linnaeus. For many of us, we only ever get to see primates in zoos and animal parks, but in many areas of the world, humans have coexisted with these animals for thousands of years. In areas where humans and primates have a long, shared history, nonhuman primates often play key roles in creation myths and cultural symbolism.
Hamadryas baboons feature significantly in Ancient Egyptian iconography. Ancient Egyptian deities and beliefs transformed over time, as did the role of hamadryas baboons. Early on, baboons were thought to represent dead ancestors, and one monkey deity, called Babi or Baba, was thought to feed off of dead souls. Later, baboons became the totem animal for Thoth, the deity of science, writing, wisdom, and measurement, who also wrote the Book of the Dead. Sunbathing hamadryas baboons led ancient Egyptians to associate them with Ra, the sun god, who was the son of Thoth. During mummification, human organs were removed and put into canopic jars, one of which was topped with the head of the baboon-headed god, Hapi. Hamadryas baboons were also often kept as pets, as depicted in hieroglyphics, and occasionally mummified as well.
On Madagascar, indris and aye-ayes play roles in the creation myths and omens of local people.There are many myths regarding the origins of indris and their relationship to humans, including one where two brothers living in the forest separated, with one brother leaving the forest and becoming a human while the other stayed in the forest to become the indri. Like humans, indris have long legs, no tail, and upright posture. They are considered sacred and are therefore protected. Unfortunately, the aye-aye is not treated with the same reverence. Because of their unusual appearance (see Figure 5.15), aye-ayes are seen as omens of death.They are usually killed when encountered because it is believed that someone will die if an aye-aye points at them.
In India, monkeys play a key role in the Hindu religion. Hanuman, who resembles a monkey, is a key figure in the Ramayana. Hanuman is thought to be a guardian deity, and so local monkeys like Hanuman langurs and macaques are protected in India (Figure 5.47). In Thailand, where Hinduism is also practiced, the Hindu reverence for monkeys extends to “monkey feasts,” where large quantities of food are spread out in gratitude to the monkeys for bringing good fortune.
The people of Japan have coexisted with Japanese macaques for thousands of years, and so monkeys play key roles in both of the major Japanese religions. In the Shinto religion, macaques are thought of as messengers between the spirit world and humans, and monkey symbols are thought to be good luck. The other major religion in Japan is Buddhism, and monkeys play a role in symbolism of this religion as well. The “Three Wise Monkeys” who see no evil, speak no evil, and hear no evil derive from Buddhist iconography of monkeys.
In Central and South America, monkeys feature often in Mayan and Aztec stories. In the Mayan creation story, the Popol Vuh, the “hero brothers,” are actually a howler monkey and a spider monkey, who represent ancestors of humans in the story. In the Aztec religion, spider monkeys are associated with the god of arts, pleasure, and playfulness. A spider monkey is also represented in a Peruvian Nazca geoglyph, a large design made on the ground by moving rocks.
In many of these regions today, the relationships between humans and nonhuman primates are complicated. The bushmeat and pet trades make these animals valuable at the expense of many animals’ lives, and in some areas, nonhuman primates have become pests who raid crop fields and consume valuable foods. All of this has led to the development of a new subarea of anthropology called Ethnoprimatology, which involves studying the political, economic, symbolic, and practical relationships between humans and nonhuman primates.This field highlights the particular challenges for humans of having to coexist with animals with whom we share so much in common. It also provides insight into some of the challenges facing primate conservation efforts (see Appendix B: Primate Conservation).
Conclusion
The Order Primates is a diverse and fascinating group of animals united in sharing a suite of characteristics—visual specialization, grasping hands and feet, large brains, and extended life histories—that differentiates us from other groups of mammals. In this chapter, we surveyed the major taxonomic groups of primates, discussing where humans fit among our close relatives as well as discovering that primates are interesting animals in their own right. We discussed a range of key traits used to distinguish between the many taxa of living primates, including dietary, locomotor, and behavioral characteristics. Because of our long, shared evolutionary history with these animals, nonhuman primates provide a crucial resource for understanding our current biology. In Chapter 6, you will discover the fascinating and complex social behaviors of nonhuman primates, which provide further insight into our evolutionary biology.
Review Questions
- Why does the field of anthropology, a field dedicated to the study of humans, include the study of nonhuman animals? What can we learn from nonhuman primates?
- Why is it important to try to place taxa into a clade classification rather than groupings based on grade? Can you think of an example?
- One of the important goals of an introductory biological anthropology course is to teach you about your place in nature. What is the full taxonomic classification of humans, and what are some of the traits we have of each of these categories?
- When you have seen primates in person, did you observe any facial expressions, behaviors, or physical traits that seemed familiar to you? If so, which ones and why?
- Draw out a tree showing the major taxonomic group of primates described here, making sure to leave room in between each level. Underneath each taxon, list some of the key features of this group so that you can compare traits between groups.
Key Terms
Activity pattern: Refers to the time of day an animal is typically active.
African clade: A grouping that includes gorillas, chimpanzees, bonobos, humans, and their extinct relatives.
Analogy: When two or more taxa exhibit similar traits that have evolved independently, the similar traits evolve due to similar selective pressures. (Also sometimes called convergent evolution, parallel evolution, or homoplasy.)
Ancestral trait: A trait that has been inherited from a distant ancestor.
Arboreal: A descriptor for an organism that spends most of its time in trees.
Asian clade: A grouping that includes orangutans and their extinct relatives.
Bilophodont: Molar pattern of cercopithecoid monkeys in which there are usually four cusps that are arranged in a square pattern and connected by two ridges.
Bipedalism: Walking on two legs.
Brachiation: A form of locomotion in which the organism swings below branches using the forelimbs.
Bunodont: Low, rounded cusps on the cheek teeth.
Canines: In most primates, these are the longest of the teeth, often conical in shape and used as a weapon against predators or others of their species.
Cathemeral: Active throughout the 24-hour period.
Clade: A grouping based on ancestral relationships; a branch of the evolutionary tree.
Cusps: The bumps on the chewing surface of the premolars and molars, which can be quite sharp in some species.
Dental formula: The number of each type of tooth in one quadrant of the mouth, written as number of incisors: canines: premolars: molars.
Derived trait: A trait that has been recently modified, most helpful when assigning taxonomic classification.
Diastema: A space between the teeth, usually for large canines to fit when the mouth is closed.
Dichromatic: Being able to see only blues and greens.
Diurnal: Active during the day.
Dry nose: The nose and upper lip are separated and the upper lip can move independently; sometimes referred to as a “hairy” or “mobile” upper lip.
Ethnoprimatology: A subarea of anthropology that studies the complexities of human-primate relationships in the modern environment.
Evolutionary trade-off: When an organism, which is limited in the time and energy it can put into aspects of its biology and behavior, is shaped by natural selection to invest in one adaptation at the expense of another.
Faunivorous: Having a diet consisting entirely of animal matter: insects, eggs, lizards, etc.
Folivore: Having a diet consisting primarily of leaves.
Fovea: A depressed area in the retina at the back of the eye containing a concentration of cells that allow one to focus on objects very close to one’s face.
Frugivore: Having a diet consisting primarily of fruit.
Generalized trait: A trait that is useful for a wide range of tasks.
Grade: A grouping based on overall similarity in lifestyle, appearance, and behavior.
Grooming claw: A claw present on the second pedal digit in strepsirrhines.
Gummivore: Having a diet consisting primarily of gums and saps.
Heterodont: Having different types of teeth.
Homology: When two or more taxa share characteristics because they inherited them from a common ancestor.
Hone: When primates sharpen their canines by wearing them on adjacent teeth.
Incisors: The spatula-shaped teeth at the front of the mouth.
Insectivore: Having a diet consisting primarily of insects.
Ischial callosities: Modified seat bones of the pelvis that are flattened and over which calluses form; function as seat pads for sitting and resting atop branches.
Knuckle-walking: A form of quadrupedal movement used by Gorilla and Pan when on the ground, wherein the front limbs are supported on the knuckles of the hands.
Life history: Refers to an organism’s pace of growth, reproduction, lifespan, etc.
Locomotion: How an organism moves around.
Male bimaturism: Refers to the alternative reproductive strategies in orangutans in which males can delay maturation, sometimes indefinitely, until a fully mature, “flanged” male disappears.
Molars: The largest teeth at the back of the mouth; used for chewing. In primates, these teeth usually have between three and five cusps.
Monochromatic: Being able to see only in shades of light to dark, no color.
Monomorphic: When males and females of a species do not exhibit significant sexual dimorphism.
Natal coat: Refers to the contrasting fur color of baby leaf monkeys compared to adults.
Nocturnal: Active at night.
Olecranon process: Bony projection at the elbow end of the ulna.
Opposable thumb or opposable big toe: Having thumbs and toes that go in a different direction from the rest of the fingers, allows for grasping with hands and feet.
Pentadactyly: Having five digits or fingers and toes.
Polymorphic color vision: A system in which individuals of a species vary in their abilities to see color. In primates, it refers to males being dichromatic and females being either trichromatic or dichromatic.
Postorbital bar: A bony ring that surrounds the eye socket, open at the back.
Postorbital closure/plate: A bony plate that provides protection to the side and back of the eye.
Prehensile tail: A tail that is able to hold the full body weight of an organism, which often has a tactile pad on the underside of the tip for improved grip.
Premolars: Smaller than the molars, used for chewing. In primates, these teeth usually have one or two cusps.
Quadrupedalism: Moving around on all fours.
Rhinariums: Wet noses; resulting from naked skin of the nose which connects to the upper lip and smell-sensitive structures along the roof of the mouth.
Sagittal crest: A bony ridge along the top/middle of the skull, used for attachment of chewing muscles.
Scent marking: The behavior of rubbing scent glands or urine onto objects as a way of communicating with others.
Semi-brachiation: A form of locomotion in which an organism swings below branches using a combination of forelimbs and prehensile tail.
Sexually dimorphic: When a species exhibits sex differences in morphology, behavior, hormones, and/or coloration.
Shearing crests: Sharpened ridges that connect cusps on a bilophodont molar.
Specialized trait: A trait that has been modified for a specific purpose.
Styloid process of ulna: A bony projection of the ulna at the end near the wrist.
Tactile pads: Sensitive skin at the fingertips for sense of touch. Animals with a prehensile tail have a tactile pad on the underside of the tail as well.
Tapetum lucidum: Reflecting layer at the back of the eye that magnifies light.
Terrestrial: A descriptor for an organism that spends most of its time on the ground.
Tetrachromatic: Having the ability to see reds, yellows, blues, greens, and ultraviolet.
Tooth comb or dental comb: A trait of the front, lower teeth of strepsirrhines in which, typically, the four incisors and canines are long and thin and protrude outward.
Trichromatic color vision: Being able to distinguish yellows and reds in addition to blues and greens.
Vertical clinging and leaping: A locomotor pattern in which animals are oriented upright while clinging to vertical branches, push off with hind legs, and land oriented upright on another vertical branch.
Y-5 molar: Molar cusp pattern in which five molar cusps are separated by a “Y”-shaped groove pattern.
About the Author
Stephanie Etting, Ph.D.
Sacramento City College and Sonoma State University, ettings@scc.losrios.edu
Dr. Etting became hooked on biological anthropology as a freshman at UC Davis when she took the “Introduction to Biological Anthropology” course. She obtained her Ph.D. in anthropology in 2011 from UC Davis, where she studied anti-predator behavior toward snakes in rhesus macaques, squirrel monkeys, and black-and-white ruffed lemurs. While in graduate school, Dr. Etting discovered her love of teaching and, since finishing her dissertation, has taught at UC Berkeley; Sonoma State University; UC Davis; California State University, Sacramento; and Sacramento City College.In addition to her interests in primate behavior, Dr. Etting is also very interested in primate evolution and functional anatomy.
For Further Exploration
Animal Diversity Web. This website is hosted by the Zoology Department at the University of Michigan. It has photographs of skulls, teeth, hands, arms, and feet of many primate species.
eSkeletons. This website is hosted by the Department of Anthropology at University of Texas, Austin. It is an interactive website where you can compare specific bones from different species of primates.
Fleagle, John G. 2013. Primate Adaptation and Evolution. Third edition. San Diego: Academic Press.
Fuentes, Agustín, and Kimberley J. Hockings. 2010. “The Ethnoprimatological Approach in Primatology.” American Journal of Primatology 72 (10): 841–847.
Rowe, Noel. 1996. Pictorial Guide to the Living Primates. Charlestown, RI: Pogonias Press.
Whitehead, Paul F., William K. Sacco, and Susan B. Hochgraf. 2005. A Photographic Atlas for Physical Anthropology. Englewood, CO: Morton Publishing.
References
Balolia, Katharine L., Christophe Soligo, and Bernard Wood. 2017. “Sagittal Crest Formation in Great Apes and Gibbons.” Journal of Anatomy 230 (6): 820–832.
Bininda-Emonds, Olaf R., Marcel Cardillo, Kate E. Jones, Ross D. E. MacPhee, Robin M. D. Beck, Richard Grenyer, Samantha A. Price, Rutger A. Vos, John L. Gittleman, and Andy Purvis. 2007. “The Delayed Rise of Present-Day Mammals.” Nature 446 (7135): 507–512.
Burton, Michael L., Carmella C. Moore, John W. M. Whiting, A. Kimball Romney, David F. Aberle, Juan A. Barcelo, Malcolm M. Dow, et al. 1996. “Regions Based on Social Structure.” Current Anthropology 37 (1): 87–123.
Chivers, David J., and C. M. Hladik. 1980. “Morphology of the Gastrointestinal Tract in Primates: Comparisons with Other Mammals in Relation to Diet.” Journal of Morphology 166 (3): 337–386.
Clutton-Brock, T. H., and Paul H. Harvey. 1980. “Primates, Brains, and Ecology.” Journal of Zoology 190 (3): 309–323.
Dunbar, Robin I. M. 1998. “The Social Brain Hypothesis.” Evolutionary Anthropology 6 (5): 178–190.
Ebersberger, Ingo, Dirk Metzler, Carsten Schwarz, and Svante Pääbo. 2002. “Genomewide Comparison of DNA Sequences Between Humans and Chimpanzees.” American Journal of Human Genetics 70 (6): 1490–1497.
Jameson, Natalie M., Zhuo-Cheng Hou, Kirstin N. Sterner, Amy Weckle, Morris Goodman, Michael E. Steiper, and Derek E. Wildman. 2011. “Genomic Data Reject the Hypothesis of a Prosimian Primate Clade.” Journal of Human Evolution 61 (3): 295–305.
Kawamura, Shoji, Chihiro Hiramatsu, Amanda D. Melin, Colleen M. Schaffner, Filippo Aureli, and Linda M. Fedigan. 2012. “Polymorphic Color Vision in Primates: Evolutionary Considerations.” In Post-Genome Biology of Primates, edited by H. Irai, H. Imai, and Y. Go, 93–120. Tokyo: Springer.
Matsui, Atsushi, Felix Rakotondraparany, Isao Munechika, Masami Hasegawa, and Satoshi Horai. 2009. “Molecular Phylogeny and Evolution of Prosimians Based on Complete Sequences of Mitochondrial DNAs.” Gene 441 (1–2): 53–66.
Nater, Alexander, Maja P. Mattle-Greminger, Anton Nurcahyo, Matthew G. Nowak, Marc de Manuel, Tariq Desai, Colin Groves, et al. 2017. “Morphometric, Behavioral, and Genomic Evidence for a New Orangutan Species.” Current Biology 27 (22): 3487–3498.
Pozzi, Luca, Jason A. Hodgson, Andrew S. Burrell, Kirstin N. Sterner, Ryan L. Raaum, and Todd R. Disotell. 2014. “Primate Phylogenetic Relationships and Divergence Dates Inferred from Complete Mitochondrial Genomes.” Molecular Phylogenetics and Evolution 75: 165–183.
Remis, Melissa J. 1997. “Western Lowland Gorillas (Gorilla gorilla gorilla) as Seasonal Frugivores: Use of Variable Resources.” American Journal of Primatology 43 (2): 87–109.
Richmond, Brian G., David R. Begun, and David S. Strait. 2001. “Origin of Human Bipedalism: The Knuckle‐Walking Hypothesis Revisited.” American Journal of Physical Anthropology 116 (S33): 70–105.
Robson, Shannen L., Carel P. van Schaik, and Kristen Hawkes. 2006. “The Derived Features of Human Life History.” In The Evolution of Human Life History, edited by Kristen Hawkes and Richard R. Paine, 17–44. Santa Fe: SAR Press.
Scally, Aylwyn, Julien Y. Dutheil, LaDeana W. Hillier, Gregory E. Jordan, Ian Goodhead, Javier Herrero, Asger Hobolth, et al. 2012. “Insights into Hominid Evolution from the Gorilla Genome Sequence.” Nature 483 (7388): 169–175.
Schneider, Horacio, and Iracilda Sampaio. 2015. “The Systematics and Evolution of New World Primates: A Review.” Molecular Phylogenetics and Evolution 82 (B): 348–357.
Setchell, Joanna M., Phyllis C. Lee, E. Jean Wickings, and Alan F. Dixson. 2001. “Growth and Ontogeny of Sexual Size Dimorphism in the Mandrill (Mandrillus sphinx).” American Journal of Physical Anthropology 115 (4): 349–360.
Utami, Sri Suci, Benoît Goossens, Michael W. Bruford, Jan R. de Ruiter, and Jan A. R. A. M. van Hooff. 2002. “Male Bimaturism and Reproductive Success in Sumatran Orang-utans.” Behavioral Ecology 13 (5): 643–652.
Vasey, Natalie. 2006. “Impact of Seasonality and Reproduction on Social Structure, Ranging Patterns, and Fission–Fusion Social Organization in Red Ruffed Lemurs.” In Lemurs: Ecology and Adaptation, edited by Lisa Gould and Michelle L. Sauther, 275–304. New York: Springer.
Wright, Patricia C. 1999. “Lemur Traits and Madagascar Ecology: Coping with an Island Environment.” American Journal of Physical Anthropology 110 (S29): 31–72.
Acknowledgments
The author would very much like to thank the editors for the opportunity to contribute to this textbook, along with anonymous reviewers who provided useful feedback on earlier drafts of this chapter. She would particularly like to thank Karin Enstam Jaffe for her support and encouragement during the writing of this chapter and its revision. Most of all, the author would like to thank all of the Introduction to Biological Anthropology students that she has had over the years who have listened to her lecture endlessly on these animals that she finds so fascinating and who have helped her to hone her pedagogy in a field that she loves.
Sarah S. King, Ph.D., Cerro Coso Community College
Kara Jones, M.A., Ph.D. student, University of Nevada Las Vegas
This chapter is a revision from "Chapter 7: Understanding the Fossil Context” by Sarah King and Lee Anne Zajicek. In Explorations: An Open Invitation to Biological Anthropology, first edition, edited by Beth Shook, Katie Nelson, Kelsie Aguilera, and Lara Braff, which is licensed under CC BY-NC 4.0.
Learning Objectives
- Identify the different types of fossils and describe how they are formed.
- Discuss relative and chronometric dating methods, the type of material they analyze, and their applications.
- Describe the methods used to reconstruct past environments.
- Interpret a site using the methods described in this chapter.
Fossil Study: An Evolving Process
Mary Anning and the Age of Wonder
Mary Anning (1799–1847) is likely the most famous fossil hunter you’ve never heard of (Figure 7.1). Anning lived her entire life in Lyme Regis on the Dorset coast in England. As a woman, born to a poor family, with minimal education (even by 19th-century standards), the odds were against Anning becoming a scientist (Emling 2009, xii). It was remarkable that Anning was eventually able to influence the great scientists of the day with her fossil discoveries and her subsequent hypotheses regarding evolution.
The time when Anning lived was a remarkable period in human history because of the Industrial Revolution in Britain. Moreover, the scientific discoveries of the 18th and 19th centuries set the stage for great leaps of knowledge and understanding about humans and the natural world. Barely a century earlier, Sir Isaac Newton had developed his theories on physics and become the president of the Royal Society of London (Dolnick 2011, 5). In this framework, the pursuit of intellectual and scientific discovery became a popular avocation for many individuals, the vast majority of whom were wealthy men (Figure 7.2).
In spite of the expectations of Georgian English society to the contrary, Anning became a highly successful fossil hunter as well as a self-educated geologist and anatomist. The geology of Lyme Regis, with its limestone cliffs, provided a fortuitous backdrop for Anning’s lifework. Now called the “Jurassic Coast,” Lyme Regis has always been a rich source for fossilized remains (Figure 7.3). Continuing her father’s passion for fossil hunting, Anning scoured the crumbling cliffs after storms for fossilized remains and shells. The work was physically demanding and downright dangerous. In 1833, while searching for fossils, Anning lost her beloved dog in a landslide and nearly lost her own life in the process (Emling 2009).
Around the age of ten, Anning located and excavated a complete fossilized skeleton of an ichthyosaurus (“fish lizard”). She eventually found Pterodactylus macronyx and a 2.7-meter Plesiosaurus, considered by many to be her greatest discovery (Figure 7.4). These discoveries proved that there had been significant changes in the way living things appeared throughout the history of the world. Like many of her peers, including Darwin, Anning had strong religious convictions. However, the evidence that was being found in the fossil record was contradictory to the Genesis story in the Bible. In The Fossil Hunter: Dinosaurs, Evolution, and the Woman Whose Discoveries Changed the World, Anning’s biographer Shelley Emling (2009, 38) notes, “the puzzling attributes of Mary’s fossil [ichthyosaurus] struck a blow at this belief and eventually helped pave the way for a real understanding of life before the age of humans.”
Intellectual and scientific debate now had physical evidence to support the theory of evolution, which would eventually result in Darwin’s seminal work, On the Origin of Species (1859). Anning’s discoveries and theories were appreciated and advocated by her friends, intellectual men who were associated with the Geological Society of London. Regrettably, this organization was closed to women, and Anning received little official recognition for her contributions to the fields of natural history and paleontology. It is clear that Anning’s knowledge, diligence, and uncanny luck in finding magnificent specimens of fossils earned her unshakeable credibility and made her a peer to many antiquarians (Emling 2009).
Fossil hunting is still providing evidence and a narrative of the story of Earth. Mary Anning recognized the value of fossils in understanding natural history and relentlessly championed her theories to the brightest minds of her day. Anning’s ability to creatively think “outside the box”—skillfully assimilating knowledge from multiple academic fields—was her gift to our present understanding of the fossil record. Given how profoundly Anning has shaped how we, in the modern day, think about the origins of life, it is surprising that her contributions have been so marginalized. Anning’s name should be on the tip of everyone’s tongue. Fortunately, at least in one sense of the word, it is. The well-known tongue twister, below, may have been written about Mary Anning:
She sells sea-shells on the sea-shore.
The shells she sells are sea-shells, I’m sure.
For if she sells sea-shells on the sea-shore
Then I’m sure she sells sea-shore shells.
—T. Sullivan (1908)
Developing Modern Methods
As Mary Anning’s story suggests, scientists in Europe were working at a time dominated by western Christian tradition. Literal interpretations of the bible did not allow for the long, slow processes of geological or evolutionary change to operate. However, many scientists were making observations that did not fit the biblical narrative. During the 18th century, Scotsman James Hutton’s work on the formation of Earth provided a much longer timeline of events than previous biblical interpretations would allow. Hutton’s theory of Deep Time was crucial to the understanding of fossils. Deep Time gave the history of Earth enough time—4.543 billion years—to encompass continental drift, the evolution of species, and the fossilization process. A second Scotsman, Charles Lyell, propelled Hutton’s work into his own theory of uniformitarianism, the doctrine that Earth’s geologic formations are the work of slow geologic forces. Lyell’s three-volume work, Principles of Geology (1830–1833), was influential to naturalist Charles Darwin (see Chapter 2 for more information on Darwin’s work). In fact, Lyell’s first volume accompanied Darwin on his five-year voyage around the world on the HMS Beagle (1831–1836). The concepts proposed by Lyell gave Darwin an opportunity to apply his working theories of evolution by natural selection and a greater length of time with which to work. These resulting theories were important scientific discoveries and paved the way for the “Age of Wonder” (Holmes 2010, xvi).
The work of Anning, Darwin, Lyell, and many others laid the foundation for the modern methods we use today. Though anthropology is focused on humans and our primate relatives (and not on dinosaurs, as many people wrongly assume), you will see that methods developed in paleontology, geology, chemistry, biology, and physics are often applied in anthropological research. In this chapter, you will learn about the primary methods and techniques employed by biological anthropologists to answer questions about fossils, the mineralized copies of once-living organisms (Figure 7.5). Ultimately, these answers provide insights into human evolution. Pay close attention to ways in which modern biological anthropologists use other disciplines to analyze evidence and reconstruct past activities and environments.
Earth: It's Older than Dirt
Scientists have developed precise and accurate dating methods based on work in the fields of physics and chemistry. Using these methods, scientists are able to establish the age of Earth as well as approximate ages of the organisms that have lived here. Earth is roughly 4.6 billion years old, give or take a few hundred million years. The first evidence for a living organism appeared around 3.5 billion years ago (bya). The scale of geologic time can seem downright overwhelming. In order to organize and make sense of Earth’s past, geologists break up that time into subunits, which are human-made divisions along Earth’s timeline. The largest subunit is the eon. An eon is further divided into eras, and eras are divided into periods. Finally, periods are divided into epochs (see Figure 7.6; Williams 2004, 37). Currently, we are living in the Phanerozoic eon, Cenozoic era, Quaternary period, and probably the Holocene epoch—though there is academic debate about the current epoch (see below).
These divisions are based on major changes and events recorded in the geologic record. Events like significant shifts in climate or mass extinctions can be used to mark the end of one geologic time unit and the beginning of another. However, it is important to remember that these borders are not real in a physical sense; they are helpful organizational guidelines for scientific research. There can be debate regarding how the boundaries are defined. Additionally, the methods we use to establish these dates are refined over time, occasionally leading to shifts in established chronology (see the discussion on calibration in the radiocarbon dating section below). For instance, the current epoch has been traditionally known as the Holocene. It began almost twelve thousand years ago (kya) during the warming period after that last major ice age. Today, there is evidence to indicate human-driven climate change is warming the world and changing the environmental patterns faster than the natural cyclical processes. This has led some scientists within the stratigraphic community to argue for a new epoch beginning around 1950 with the Nuclear Age called the Anthropocene (Monastersky 2015; Waters et al. 2016). Nobel Laureate Paul Crutzen places the beginning of the Anthropocene much earlier—at the dawn of the Industrial Revolution, with its polluting effects of burning coal (Crutzen and Stoermer 2000, 17–18). Geologist William Ruddiman argues that the epoch began 5,000–8,000 years ago with the advent of agriculture and the buildup of early methane gasses (Ruddiman et al. 2008). Regardless of when the Anthropocene started, the major event that marks the boundary is the warming temperatures and mass extinction of nonhuman species caused by human activity (Figure 7.7). Researchers now declare that “human activity now rivals geologic forces in influencing the trajectory of the Earth System” (Steffen et al. 2018, 1).
Fossils: The Taphonomic Process
Most of the evidence of human evolution comes from the study of the dead. To obtain as much information as possible from the remains of once-living creatures, one must understand the processes that occur after death. This is where taphonomy comes in (Figure 7.8). Taphonomy is the study of what happens to an organism after death (Komar and Buikstra 2008, 189; Stodder 2008). It includes the study of how an organism becomes a fossil. However, as you’ll see throughout this book, the majority of organisms never make it through the full fossilization process.
Taphonomy is important in biological anthropology, especially in subdisciplines like bioarchaeology (the study of human remains in the archaeological record) and zooarchaeology (the study of faunal remains from archaeological sites). It is so important that many scientists have recreated a variety of burial and decay experiments to track taphonomic change in modern contexts. These contexts can then be used to understand the taphonomic patterns seen in the fossil record (see Reitz and Wing 1999, 122–141).
Taphonomic analysis can also give us important insights into the development of complex thought and ritual in human evolution. In Chapter 11, you will see the first evidence of recognized burial practices in hominins. Taphonomy helped to establish whether these burials were simply the result of natural processes or intentionally constructed by humans (Klein 1999, 395; Straus 1989). Deliberate burials often include the body placed in a specific position, such as supine (on the back) with arms crossed over the chest or in a flexed position (think fetal position) facing a particular direction. If bones have evidence of a carnivore or rodent gnawing on them, it can be inferred that the remains were exposed to scavengers after death.
Going back further in time, taphonomic evidence may tell us how our ancestors died. For instance, several australopithecine fossils show evidence of carnivore tooth marks and even punctures from saber-toothed cats, indicating that we weren’t always the top of the food chain. The Bodo Cranium, a Homo erectus cranium from Middle Awash Valley, Ethiopia, shows cut marks made by stone tools, indicating an early example of possible defleshing activity in our human ancestors (White 1986). At the archaeological site of Zhoukoudian, researchers used taphonomy to show that the highly fragmented remains of at least 51 Homo erectus individuals were scavenged by Pleistocene cave hyenas (Boaz et al. 2004). The damage on Skull VI was described as “elongated, raking bite marks, isolated puncture bite marks, and perimortem breakage consistent with patterns of modern hyaenid bone modification” (Boaz et al. 2004). Additionally, a fresh burnt equid cranium was discovered which supports the theory of mobile hominid scavenging and fire use at the site (Boaz et al. 2004).
Special Topic: Bog Bodies and Mummies (Add a picture to show how well preserved they are!)
Preservation is a key topic in anthropological research, since we can only study the evidence that gets left behind in the fossil and archaeological record. This chapter is concerned with the fossil record; however, there are other forms of preserved remains that provide anthropologists with information about the past. You’ve undoubtedly heard of mummification, likely in the context of Egyptian or South American mummies. However, bog bodies and ice mummies are further examples of how remains can be preserved in special circumstances. It is important to note that fossilization is a process that takes much longer than the preservation of bog bodies or mummies.
Bog bodies are good examples of wetland preservation. Peat bogs are formed by the slow accumulation of vegetation and silts in ponds and lakes. Individuals were buried in bogs throughout Europe as far back as 10 kya, with a proliferation of activity from 1,600 to 3,200 years ago (Giles 2020; Ravn 2010). When they were found thousands of years later, they resembled recent burials. Their hair, skin, clothing, and organs were exceptionally well preserved, in addition to their bones and teeth (Eisenbeiss 2016; Ravn 2010). Preservation was so good in fact that archaeologists could identify the individuals’ last meals and re-create tattoos found on their skin.
Extreme cold can also halt the natural decay process. A well-known ice mummy is Ötzi, a Copper Age man dating to around 5,200 years ago found in the Alps (Vanzetti et al. 2012; Vidale et al. 2016). As with the bog bodies, his hair, skin, clothing, and organs were all well preserved. Recently, archaeologists were able to identify his last meal (Maixner et al. 2018). It was high in fat, which makes sense considering the extremely cold environment in which he lived, as meals high in fat assist in cold tolerance (Fumagalli et al. 2015).
In the Andes, ancient peoples would bury human sacrifices throughout the high peaks in a sacred ritual called Capacocha (Wilson et al. 2007). The best-preserved mummy to date is called the “Maiden” or “Sarita” because she was found at the summit of Sara Sara Volcano. Her remains are over 500 years old, but she still looks like the 15-year-old girl she was at the time of her death, as if she had just been sleeping for 500 years (Reinhard 2006).
Finally, arid environments can also contribute to the preservation of organic remains. As discussed with waterlogged sites, much of the bacteria that is active in breaking down bodies is already present in our gut and begins the putrefaction process shortly after death. Arid environments deplete organic material of the moisture that putrefactive bacteria need to function (Booth et al. 2015). When that occurs, the soft tissue like skin, hair, and organs can be preserved. It is similar to the way a food dehydrator works to preserve meat, fruit, and vegetables for long-term storage. There are several examples of arid environments spontaneously preserving human remains, including catacomb burials in Austria and Italy (Aufderheide 2003, 170, 192–205).
Fossilization
Fossils only represent a tiny fraction of creatures that existed in the past. It is extremely difficult for an organism to become a fossil. After all, organisms are designed to deteriorate after they die. Bacteria, insects, scavengers, weather, and environment all aid in the process that breaks down organisms so their elements can be returned to Earth to maintain ecosystems (Stodder 2008). Fossilization, therefore, is the preservation of an organism against these natural decay processes (Figure 7.9).
For fossilization to occur, several important things must happen. First, the organism must be protected from things like bacterial activity, scavengers, and temperature and moisture fluctuations. A stable environment is important. This means that the organism should not be exposed to significant fluctuations in temperature, humidity, and weather patterns. Changes to moisture and temperature cause the organic tissues to expand and contract repeatedly, which will eventually cause microfractures and break down (Stodder 2008). Soft tissue like organs, muscle, and skin are more easily broken down in the decay process; therefore, they are less likely to be preserved. Bones and teeth, however, last much longer and are more common in the fossil record (Williams 2004, 207).
Wetlands are a particularly good area for preservation because they allow for rapid permanent burial and a stable moisture environment. That is why many fossils are found in and around ancient lakes and river systems. Waterlogged sites can also be naturally anaerobic (without oxygen). Much of the bacteria that causes decay is already present in our gut and can begin the decomposition process shortly after death during putrefaction (Booth et al. 2015). Since oxygen is necessary for the body’s bacteria to break down organic material, the decay process is significantly slowed or halted in anaerobic conditions.
The next step in the fossilization process is sediment accumulation. The sediments cover and protect the organism from the environment. They, along with water, provide the minerals that will eventually become the fossil (Williams 2004, 31). Sediment accumulation also provides the pressure needed for mineralization to take place. Lithification is when the weight and pressure of the sediments squeeze out extra fluids and replace the voids that appear with minerals from the surrounding sediments. Finally, we have permineralization. This is when the organism is fully replaced by minerals from the sediments. A fossil is really a mineral copy of the original organism (Williams 2004, 31).
Types of Fossils
Plants
Plants make up the majority of fossilized materials. One of the most common plants existing today, the fern, has been found in fossilized form many times. Other plants that no longer exist or the early ancestors of modern plants come in fossilized forms as well. It is through these fossils that we can discover how plants evolved and learn about the climate of Earth over different periods of time.
Another type of fossilized plant is petrified wood. This fossil is created when actual pieces of wood—such as the trunk of a tree—mineralize and turn into rock. Petrified wood is a combination of silica, calcite, and quartz, and it is both heavy and brittle. Petrified wood can be colorful and is generally aesthetically pleasing because all the features of the original tree’s composition are illuminated through mineralization (Figure 7.10). There are a number of places all over the world where petrified wood “forests” can be found, but there is an excellent assemblage in Arizona, at the Petrified Forest National Park. At this site, evidence relating to the environment of the area some 225 mya is on display.
Human/Animal Remains
We are more familiar with the fossils of early animals because natural history museums have exhibits of dinosaurs and extinct mammals. However, there are a number of fossilized hominin remains that provide a picture of the fossil record over the course of our evolution from primates. The term hominins includes all human ancestors who existed after the evolutionary split from chimpanzees and bonobos, some six to seven mya. Modern humans are Homo sapiens, but hominins can include much earlier versions of humans. One such hominin is “Lucy” (AL 288-1), the 3.2 million-year-old fossil of Australopithecus afarensis that was discovered in Ethiopia in 1974 (Figure 7.11). Until recently, Lucy was the most complete and oldest hominin fossil, with 40% of her skeleton preserved (see Chapter 9 for more information about Lucy). In 1994, an Australopithecus fossil nicknamed “Little Foot” (Stw 573) was located in the World Heritage Site at Sterkfontein Caves (“the Cradle of Humankind”) in South Africa. Little Foot is more complete than Lucy and possibly the oldest fossil that has so far been found, dating to at least 3.6 million years (Granger et al. 2015). The ankle bones of the fossil were extricated from the matrix of concrete-like rock, revealing that the bones of the ankles and feet indicate bipedalism (University of Witwatersrand 2017).
Both the Lucy and Little Foot fossils date back to the Pliocene (5.8 to 2.3 mya). Older hominin fossils from the late Miocene (7.25 to 5.5 mya) have been located, although they are much less complete. The oldest hominin fossil is a fragmentary skull named Sahelanthropus tchadensis, found in Northern Chad and dating to circa seven mya (Lebatard et al. 2008). It is through the discovery, dating, and study of primate and early hominin fossils that we find physical evidence of the evolutionary timeline of humans.
Asphalt
Asphalt, a form of crude oil, can also yield fossilized remains. Asphalt is commonly referred to in error as tar because of its viscous nature and dark color. A famous fossil site from California is La Brea Tar Pits in downtown Los Angeles (Figure 7.12). In the middle of the busy city on Wilshire Boulevard, asphalt (not tar) bubbles up through seeps (cracks) in the sidewalk. The La Brea Tar Pits Museum provides an incredible look at the both extinct and extant animals that lived in the Los Angeles Basin 40,000–11,000 years ago. These animals became entrapped in the asphalt during the Pleistocene and perished in place. Ongoing excavations have yielded millions of fossils, including megafauna such as American mastodons and incomplete skeletons of extinct species of dire wolves, Canis dirus, and the saber-toothed cat, Smilodon fatalis (Figure 7.13). Fossilized remains of plants have also been found in the asphalt. The remains of one person have also been found at the tar pits. Referred to as La Brea Woman, the remains were found in 1914 and were subsequently dated to around 10,250 years ago. The La Brea Woman was a likely female individual who was 17–28 years old at the time of her death, with a height of under five feet (Spray 2022). She is thought to have died from blunt force trauma to her head, famously making her Los Angeles’s first documented homicide victim (Spray 2022). (Learn more about her in the Special Topic box, “Necropolitics,” below.) Between the fossils of animals and those of plants, paleontologists have a good idea of the way the Los Angeles Basin looked and what the climate in the area was like many thousands of years ago.
Igneous Rock
Most fossils are found in sedimentary rock. This type of rock has been formed from deposits of minerals over millions of years in bodies of water on Earth’s surface. Some examples include shale, limestone, and siltstone. Sedimentary rock typically has a layered appearance. However, fossils have been found in igneous rock as well. Igneous rock is volcanic rock that is created from cooled molten lava. It is rare for fossils to survive molten lava, and it is estimated that only 2% of all fossils have been found in igneous rock (Ingber 2012). Part of a giant rhinocerotid skull dating back 9.2 mya to the Miocene was discovered in Cappadocia, Turkey, in 2010. The fossil was a remarkable find because the eruption of the Çardak caldera was so sudden that it simply dehydrated and “baked” the animal (Antoine et al. 2012).
Trace Fossils
Depending on the specific circumstances of weather and time, even footprints can become fossilized. Footprints fall into the category of trace fossils, which includes other evidence of biological activity such as nests, burrows, tooth marks, and shells. A well-known example of trace fossils are the Laetoli footprints in Tanzania (Figure 7.14). More recently, archaeological investigations in North America have revealed fossil footprints which rewrite the history of people in the Americas at White Sands, New Mexico. You can read more about the Laetoli and White Sands footprints in the Dig Deeper box below.
Other fossilized footprints have been discovered around the world. At Pech Merle cave in the Dordogne region of France, archaeologists discovered two fossilized footprints. They then brought in indigenous trackers from Namibia to look for other footprints. The approach worked, as many other footprints belonging to as many as five individuals were discovered with the expert eyes of the trackers (Pastoors et al. 2017). These footprints date back 12,000 years (Granger Historical Picture Archive 2018).
Some of the more unappealing but still-fascinating trace fossils are bezoars and coprolite. Bezoars are hard, concrete-like substances found in the intestines of fossilized creatures. Bezoars start off like the hair balls that cats and rabbits accumulate from grooming, but they become hard, concrete-like substances in the intestines. If an animal with a hairball dies before expelling the hair ball mass and the organism becomes fossilized, that mass becomes a bezoar.
Coprolite is fossilized dung. One of the best collections of coprolites is affectionately known as the “Poozeum.” The collection includes a huge coprolite named “Precious” (Figure 7.15). Coprolite, like all fossilized materials, can be in matrix—meaning that the fossil is embedded in secondary rock. As unpleasant as it may seem to work with coprolites, remember that the organic material in dung has mineralized or has started to mineralize; therefore, it is no longer soft and is generally not smelly. Also, just as a doctor can tell a lot about health and diet from a stool sample, anthropologists can glean a great deal of information from coprolite about the diets of ancient animals and the environment in which the food sources existed. For instance, 65 million-year-old grass phytoliths (microscopic silica in plants) found in dinosaur coprolite in India revealed that grasses had been in existence much earlier than scientists initially believed (Taylor and O’Dea 2014, 133).
Pseudofossils
Pseudofossils are not to be mistaken for fake fossils, which have vexed scientists from time to time. A fake fossil is an item that is deliberately manipulated or manufactured to mislead scientists and the general public. In contrast, pseudofossils are not misrepresentations but rather misinterpretations of rocks that look like true fossilized remains (S. Brubaker, personal communication, March 9, 2018). Pseudofossils are the result of impressions or markings on rock, or even the way other inorganic materials react with the rock. A common example is dendrites, the crystallized deposits of black minerals that resemble plant growth (Figure 7.16). Other examples of pseudofossils are unusual or odd-shaped rocks that include various concretions and nodules. An expert can examine a potential fossil to see if there is the requisite internal structure of organic material such as bone or wood that would qualify the item as a fossil.
Dig Deeper: Trace Fossils The Power of Poop
Coprolites found in Paisley Caves, Oregon, in the United States are shedding new light on some of the earliest occupants in North America. Human coprolites are distinguished from animal coprolites through the identification of fecal biomarkers using lipids, or fats, and bile acids (Shillito et al. 2020a). Paisley Caves have 16,000 years of anthropogenic, or human-caused, deposition, with some coprolites having been dated as old as 12.8kya (Blong et al. 2020). Over 285 radiocarbon dates have been recorded from the site (Shillito et al. 2020a), making Paisley Caves one of the most well-dated archaeological sites in the United States. Coprolite analysis can be summarized in three levels, macroscopic, microscopic, and molecular. This can also be understood as analyzing the morphology (macroscopic), contents (microscopic), and residues (molecular) (Shillito et al. 2020b). Each of these levels adds a different layer of information. Coprolite shape is informative through what can be seen macroscopically, such as ingestions of basketry or cordage, small gravels and grains, and general shape. The contents of coprolites may be of the most interest to scientists because certain plants and animals can signal past environments as well as food procurement methods. Coprolites from Paisley Caves have included small pebbles and obsidian chips from butchering game, grinding plants, and general food preparation as well as small bits of fire cracked rock likely from cooking in hearths (Blong 2020). Additionally, rodent bones in coprolites included crania and vertebrae, which suggests whole consumption (Taylor et al. 2020). Insect remains are present in the coprolites as well, such as ants, Jerusalem crickets, June beetles, and darkling beetles (Blong 2020). In all, the coprolites of Paisley Caves have provided an invaluable resource to anthropologists to study the past climate and lifeways of early humans in the Americas.
Coprolites can also signal past health, which is a study known as paleopathology. A study by Katelyn McDonough and colleagues (2022) focused on the identification of parasites in coprolites at Bonneville Estates Rockshelter in eastern Nevada and their link to the greater Great Basin during the Archaic, a period of time spanning 8,000–5,000 years ago. According to the study, parasites such as Acanthocephalans (thorny-headed worms) have been affecting the Great Basin for at least the last 10,000 years. Acanthocephalans are endoparasites, meaning parasites that live inside of their hosts. They are found worldwide and seem to have been concentrated in the Great Basin in the past. Bonneville Estates Rockshelter has been visited by humans for over 13,000 years, with parasite identification going back to nearly 7,000 years. The species identified at Bonneville Estates is Moniliformis clarki. This species parasitizes crickets and insects, a popular food source during the Archaic in the Great Basin. The parasite uses intermediate hosts to get to mammals and birds as definitive hosts. Crickets and beetles have been recorded as food materials in Paisley Caves as well. Insects have remained an important dietary staple for people of the Great Basin and are consumed raw, dried, brined, or ground into flour. Insects that remain uncooked or undercooked have a higher risk for transmission of parasites. Symptoms associated with Acanthocephalans infection are intense intestinal discomfort, anemia, and anorexia, leading to death. It is hypothesized that the consumption of basketry, cordage, and charcoal (which was also identified at Paisley Caves), sometimes associated with parasite-infected coprolites, may have been a method of treatment for the infection. Interestingly, present day infections from this parasite are rising after remaining quite rare, as detection of the parasite is occurring in insect farms.
Walking to the Past
In 1974, British anthropologist Mary Leakey discovered fossilized animal tracks at Laetoli (Figure 7.17), not far from the important paleoanthropological site at Olduvai Gorge in Tanzania. A few years later, a 27-meter trail of hominin footprints were discovered at the same site. These 70 footprints, now referred to as the Laetoli Footprints, were created when early humans walked in wet volcanic ash. Before the impressions were obscured, more volcanic ash and rain fell, sealing the footprints. These series of environmental events were truly extraordinary, but they fortunately resulted in some of the most famous and revealing trace fossils ever found. Dating of the footprints indicate that they were made 3.6 mya (Smithsonian National Museum of Natural History 2018).
Just as forensic scientists can use footprints to identify the approximate build of a potential suspect in a crime, archaeologists have read the Laetoli Footprints for clues to these early humans. The footprints clearly indicate bipedal hominins who had similar feet to those of modern humans. Analysis of the gait through computer simulation revealed that the hominins at Laetoli walked similarly to the way we walk today (Crompton 2012). More recent analyses confirm the similarity to modern humans but also indicate a gait that involved more of a flexed limb than that of modern humans (Hatala et al. 2016; Raichlen and Gordon 2017). The relatively short stride implies that these hominins had short legs—unlike the longer legs of later early humans who migrated out of Africa (Smithsonian National Museum of Natural History 2018). In the context of Olduvai Gorge, where fossils of Australopithecus afarensis have been located and dated to the same timeframe as the footprints, it is likely that these newly discovered impressions were left by these same hominins.
The footprints at Laetoli were made by a small group of as many as three Australopithecus afarensis, walking in close proximity, not unlike what we would see on a modern street or sidewalk. Two trails of footprints have been positively identified with the third set of prints appearing smaller and set in the tracks left by one of the larger individuals. While scientific methods have given us the ability to date the footprints and understand the body mechanics of the hominin, additional consideration of the footprints can lead to other implications. For instance, the close proximity of the individuals implies a close relationship existed between them, not unlike that of a family. Due to the size variation and the depth of impression, the footprints seem to have been made by two larger adults and possibly one child. Scientists theorize that the weight being carried by one of the larger individuals is a young child or a baby (Masao et al. 2016). Excavation continues at Laetoli today, resulting in the discovery of two more footprints in 2015, also believed to have been made by Au. afarensis (Masao et al. 2016).
But it is not just human evolution studies that can benefit from the analysis of fossil footprints. A recent discovery of fossilized footprints has rewritten what we know about the peopling of the Americas. It was originally thought that humans had been in the Americas for at least the last 15,000 years by crossing through the ice-free corridor (IFC) between the Cordilleran and Laurentide ice sheets in present-day Alaska and Canada. However, fossil footprints from the Tularosa Basin of New Mexico (see Figure 7.18) discovered in 2021 have challenged this theory. The footprints, dated between 22,860 (∓320) and 21,130 (∓250) years ago (nps.gov) based on Ruppia cirrhosa grass seeds located above and below the footprints, have shown humans have been in the Americas for much longer than previously thought. These footprints represent an adolescent individual and toddler walking through the lakebed at White Sands (see Figure 7.19), New Mexico, alongside both giant ground sloths and mammoths (Barras 2022; Wade 2021). Also present in the lakebed are footprints of camels and dire wolves (nps.gov 2022; Wade 2021).
The IFC model was upheld by a group of theorists known as “Clovis First,” who believed the migration of people into the Americas was recent and was represented archaeologically through the Clovis projectile point toolkit. Subsequent discoveries at sites such as Cactus Hill on the east coast of the United States and Monte Verde, Chile, have demonstrated that this model wouldn’t have worked. Because these sites are as old as 20,000 years and 18,500 years respectively, the IFC would have been frozen over and impassable (Gruhn 2020). Other models have been adopted to account for this, such as the coastal migration model down the west coast of North America. The more-likely migration scenario seems to be neither of these as more discoveries or antiquity continue to emerge. People may instead have migrated into the Americas before the last glacial maximum began, around 25,500–19,000 years ago. According to Indigenous knowledge, they have always been here. With the discovery of the White Sands footprints, it is known that humans have been in the Americas for at least 20,000 years.
This discovery also reveals the importance of recognizing knowledge beyond that which is produced by the European scientific tradition. Rather than framing science in a way that runs counter to Indigenous knowledge, it can be thought that science is catching up with it. For instance, the Acoma Pueblo people have the word for camel in their vocabulary. This was dismissed by scientists who assumed the word was for describing camels that were introduced to the United States in the past 100 years. However, the discovery of the White Sands footprints also included the footprints of Pleistocene camels in the same strata. Therefore, the fact that the Acoma Pueblo people have had a word for camel likely refers the Pleistocene-age megafauna camel, Camelops hesternus, rather than Camelus dromedarius or Camelus bactrianus, two present-day camel species (which are actually descendants of Camelops hesternus). Therefore, the existence of the Acoma Pueblo word for camel is not like an anomaly but rather a testament to the fact that Acoma Pueblo ancestors walked beside C. hesternus on this continent 20,000 years ago. These footprints challenge the “ice-free corridor” expansion model, as the bridge connecting present-day Alaska and Russia into Canada would have been covered in an impenetrable ice sheet at this time. The discovery of these footprints urges scientists to reconsider further investigations at well-known Terminal Pleistocene/Early Holocene dry lake beds in the Southwestern and Mojave deserts—and to include Indigenous knowledge in their work rather than ignore it.
Special Topic: Necropolitics
What are necropolitics? Necropolitics is an application of critical theory that describes how “governments assign differential value to human life” and similarly how someone is treated after they die (Verghese 2021). How is someone’s death political?
Consider the La Brea Woman example from the section on asphalt above. The La Brea Woman’s discovery was controversial, not because she is the only person to be found in the tar pits or because of her age but also because of necropolitics. The La Brea Woman was collected in 1914 and her body was housed on display at the George C. Page Museum in Los Angeles against the wishes of the Chumash and the Tongva, two tribes whose ancestral lands include Los Angeles. The museum decided to display a skull cast instead to meet the request of the tribes which included a separate postcranial skeleton from a different individual. The updated display itself was wrought with other ethical issues, as a cast of her skull was “attached to the ancient remains of a Pakistani female that was dyed dark bronze, the femurs shortened to approximate the stature of native people” (Cooper 2010). In both cases, neither the individuals or their descendent communities consented to the display or grotesque modification of human remains. According to an interview conducted by LA Weekly (Cooper 2010) with Cindi Alvitre, former chair of the Gabrielino-Tongva Tribal Council, the display of Indigenous human remains is akin to voyeurism. She states “It's disheartening to me because it's very inappropriate to display any human remains. The things we do to fill the imagination of visitors. It violates human rights.” It is important to listen to the wishes of Indigenous people and center their values when conducting work with their ancestors. A good source for considering places to look for archaeological research ethics before conducting fieldwork (and ideally during your research design) is the Society for American Archaeology’s ethics principle list, as well as following the Indigenous Archaeology Collective.
Indigenous remains are now protected in the United States due to legislation such as Native American Graves Protection and Repatriation Act (NAGPRA). You can read more about this in Chapter 15: Bioarchaeology and Forensic Anthropology. Before the passing of NAGPRA, tribes had little agency over how the bodies of their ancestors were treated by anthropologists and museums, including decisions about sampling and destructive tests. Now when archaeological field work is conducted on federal land, tribes must be consulted before work begins. This consultation process often includes what to do if human remains are encountered. Indigenous tribes are multifaceted and multivocal; each has its own rules about how to handle the remains of their ancestors. In some cases, all work on the project must be halted after the discovery of human remains. Other tribes allow for work to continue if the remains are moved and reburied. Some tribes are open to radiometric dating if it aligns with their beliefs in the afterlife. Each tribe is different, and each tribe deserves to have its wishes respected.
Voices From the Past: What Fossils Can Tell Us
Given that so few organisms ever become fossilized, any anthropologist or fossil hunter will tell you that finding a fossil is extremely exciting. But this is just the beginning of a fantastic mystery. With the creative application of scientific methods and deductive reasoning, a great deal can be learned about the fossilized organism and the environment in which it lived, leading to enhanced understanding of the world around us.
Dating Methods
Context is a crucial concept in paleoanthropology and archaeology. Objects and fossils are interesting in and of themselves, but without context there is only so much we can learn from them. One of the most important contextual pieces is the dating of an object or fossil. By being able to place it in time, we can compare it more accurately with other contemporary fossils and artifacts or we can better analyze the evolution of a fossil species or artifacts. To answer the question “How do we know what we know?,” you have to know how archaeologists and paleoanthropologists establish dates for artifacts, fossils, and sites.
Though accurate dating is important for context and analysis, we must consider the impact. Many of the chronometric dating methods used by anthropologists require the removal of small samples from artifacts, bones, soils, and rock. Thus these techniques are considered destructive. How much of an artifact are you willing to destroy to get your date? Sharon Clough, a Senior Environmental Officer at Cotswold Archaeology, addressed this issue in a case study from her research. She stated that “the benefit of a date did not outweigh the destruction of a valuable and finite resource” (Clough 2020). The resource in question was human remains. When considering our dating options, we want to be sure that we do as little harm as possible, especially in the case of human remains (read more about this issue in the Special Topic box, “Necropolitics”).
Dating techniques are divided into two broad categories: relative dating methods and chronometric (sometimes called absolute) dating methods.
Relative Dating
Relative dating methods are used first because they rely on simple observational skills. In the 1820s, Christian Jürgensen Thomsen at the National Museum of Denmark in Copenhagen developed the “three-age” system still used in European archaeology today (Feder 2017, 17). He categorized the artifacts at the museum based on the idea that simpler tools and materials were most likely older than more complex tools and materials. Stone tools must predate metal tools because they do not require special technology to develop. Copper and bronze tools must predate iron because they can be smelted or worked at lower temperatures, etc. Based on these observations, he categorized the artifacts into Stone Age, Bronze Age, and Iron Age.
The restriction of relative dating is that you don’t know specific dates or how much time passed between different sites or artifacts. You simply know that one artifact or fossil is older than another. Thomsen knew that Stone Age artifacts were older than Bronze Age artifacts, but he couldn’t tell if they were hundreds of years older or thousands of years older. The same is true with fossils that have differences of ages into the hundreds of millions of years.
The first relative dating technique is stratigraphy (Figure 7.20). You might have already heard this term if you have watched documentaries on archaeological excavations. That’s because this method is still being used today. It provides a solid foundation for other dating techniques and gives important context to artifacts and fossils found at a site.
Stratigraphy is based on the Law of Superposition first proposed by Nicholas Steno in 1669 and further explored by James Hutton (the previously mentioned “Father” of Deep Time). Essentially, superposition tells us that things on the bottom are older than things on the top (Williams 2004, 28). Notice on Figure 7.20 that there are distinctive layers piled on top of each other. It stands to reason that each layer is older than the one immediately on top of it (Hester et al. 1997, 338). Think of a pile of laundry on the floor. Over the course of a week, as dirty clothes get tossed on that pile, the shirt tossed down on Monday will be at the bottom of the pile while the shirt tossed down on Friday will be at the top. Assuming that the laundry pile was undisturbed throughout the week, if the clothes were picked up layer by layer, the clothing choices that week could be reconstructed in the order that they were worn.
Another relative dating technique is biostratigraphy. This form of dating looks at the context of a fossil or artifact and compares it to the other fossils and biological remains (plant and animal) found in the same stratigraphic layers. For instance, if an artifact is found in the same layer as wooly mammoth remains, you know that it must date to around the last ice age, when wooly mammoths were still abundant on Earth. In the absence of more specific dating techniques, early archaeologists could prove the great antiquity of stone tools because of their association with extinct animals. The application of this relative dating technique in archaeology was used at the Folsom site in New Mexico. In 1927, a stone spear point was discovered embedded in the rib of an extinct species of bison. Because of the undeniable association between the artifact and the ancient animal, there was scientific evidence that people had occupied the North American continent since antiquity (Cook 1928).
Similar to biostratigraphic dating is cultural dating (Figure 7.21). This relative dating technique is used to identify the chronological relationships between human-made artifacts. Cultural dating is based on artifact types and styles (Hester et al. 1997, 338). For instance, a pocket knife by itself is difficult to date. However, if the same pocket knife is discovered surrounded by cassette tapes and VHS tapes, it is logical to assume that the artifact came from the late 20th century like the cassette and VHS tapes. The pocket knife could not be dated earlier than the late 20th century because the tapes were made no earlier than 1977. In the Thomsen example above, he was able to identify a relative chronology of ancient European tools based on the artifact styles, manufacturing techniques, and raw materials. Cultural dating can be used with any human-made artifacts. Both cultural dating and biostratigraphy are most effective when researchers are already familiar with the time periods for the artifacts and animals. They are still used today to identify general time periods for sites.
Chemical dating was developed in the 19th century and represents one of the early attempts to use soil composition and chemistry to date artifacts. A specific type of chemical dating is fluorine dating, and it is commonly used to compare the age of the soil around bone, antler, and teeth located in close proximity (Cook and Ezra-Cohn 1959; Goodrum and Olson 2009). While this technique is based on chemical dating, it only provides the relative dates of items rather than their absolute ages. For this reason, fluorine dating is considered a hybrid form of relative and chronometric dating methods (which will be discussed next).
Soils contain different amounts of chemicals, and those chemicals, such as fluorine, can be absorbed by human and animal bones buried in the soil. The longer the remains are in the soil, the more fluorine they will absorb (Cook and Ezra-Cohn 1959; Goodrum and Olson 2009). A sample of the bone or antler can be processed and measured for its fluorine content. Unfortunately, this absorption rate is highly sensitive to temperature, soil pH, and varying fluorine levels in local soil and groundwater (Goodrum and Olson 2009; Haddy and Hanson 1982). This makes it difficult to get an accurate date for the remains or to compare remains between two sites. However, this technique is particularly useful for determining whether different artifacts come from the same burial context. If they were buried in the same soil for the same length of time, their fluorine signatures would match.
Chronometric Dating
Unlike relative dating methods, chronometric dating methods provide specific dates and time ranges. Many of the chronometric techniques we will discuss are based on work in other disciplines such as chemistry and physics. The modern developments in studying radioactive materials are accurate and precise in establishing dates for ancient sites and remains.
Many of the chronometric dating methods are based on the measurement of radioactive decay of particular elements. Elements are materials that cannot be broken down into more simple materials without losing their chemical identity (Brown et al. 2018, 48). Each element consists of an atom that has a specific number of protons (positively charged particles) and electrons (negatively charged particles) as well as varying numbers of neutrons (particles with no charge). The protons and neutrons are located in the densely compacted nucleus of the atom, but the majority of the volume of an atom is space outside the nucleus around which the electrons orbit (see Figure 7.22).
Elements are classified based on the number of protons in the nucleus. For example, carbon has six protons, giving it an atomic number 6. Uranium has 92 protons, which means that it has an atomic number 92. While the number of protons in the atom of an element do not vary, the number of neutrons may. Atoms of a given element that have different numbers of neutrons are known as isotopes.
The majority of an atom’s mass is determined by the protons and neutrons, which have more than a thousand times the mass of an electron. Due to the different numbers of neutrons in the nucleus, isotopes vary by nuclear/atomic weight (Brown et al. 2018, 94). For instance, isotopes of carbon include carbon 12 (12C), carbon 13 (13C), and carbon 14 (14C). Carbon always has six protons, but 12C has six neutrons whereas 14C has eight neutrons. Because 14C has more neutrons, it has a greater mass than 12C (Brown et al. 2018, 95).
Most isotopes in nature are considered stable isotopes and will remain in their normal structure indefinitely. However, some isotopes are considered unstable isotopes (sometimes called radioisotopes) because they spontaneously release energy and particles, transforming into stable isotopes (Brown et al. 2018, 946; Flowers et al. 2018, section 21.1). The process of transforming the atom by spontaneously releasing energy is called radioactive decay. This change occurs at a predictable rate for nearly all radioisotopes of elements, allowing scientists to use unstable isotopes to measure time passage from a few hundred to a few billion years with a large degree of accuracy and precision.
The leading chronometric method for archaeology is radiocarbon dating (Figure 7.23). This method is based on the decay of 14C, which is an unstable isotope of carbon. It is created when nitrogen 14 (14N) interacts with cosmic rays, which causes it to capture a neutron and convert to 14C. Carbon 14 in our atmosphere is absorbed by plants during photosynthesis, a process by which light energy is turned into chemical energy to sustain life in plants, algae, and some bacteria. Plants absorb carbon dioxide from the atmosphere and use the energy from light to convert it into sugar that fuels the plant (Campbell and Reece 2005, 181–200). Though 14C is an unstable isotope, plants can use it in the same way that they use the stable isotopes of carbon.
Animals get 14C by eating the plants. Humans take it in by eating plants and animals. After death, organisms stop taking in new carbon, and the unstable 14C will begin to decay. Carbon 14 has a half-life of 5,730 years (Hester et al. 1997, 324). That means that in 5,730 years, half the amount of 14C will convert back into 14N. Because the pattern of radioactive decay is so reliable, we can use 14C to accurately date sites up to 55,000 years old (Hajdas et al. 2021). However, 14C can only be used on the remains of biological organisms. This includes charcoal, shell, wood, plant material, and bone. This method involves destroying a small sample of the material. Earlier methods of radiocarbon dating required at least 1 gram of material, but with the introduction of accelerator mass spectrometry (AMS), sample sizes as small as 1 milligram can now be used (Hajdas et al. 2021). This significantly reduces the destructive nature of this method.
The use of radiocarbon dating at Denisova Cave in modern-day Russia revealed an astounding find, the first dated first-generation individual with a Neanderthal mother and Denisovan father. Vivian Slon and colleagues (2018) sequenced the genome, which revealed the individual's hybrid genetic background, and radiocarbon dated the remains, revealing the sub-adult was over 50,000 years old (Slon et al. 2018).
As mentioned before, 14C is unstable and ultimately decays back into 14N. This decay is happening at a constant rate (even now, inside your own body!). However, as long as an organism is alive and taking in food, 14C is being replenished in the body. As soon as an organism dies, it no longer takes in new 14C. We can then use the rate of decay to measure how long it has been since the organism died (Hester et al. 1997, 324). However, the amount of 14C in the atmosphere is not stable over time. It fluctuates based on changes to the earth’s magnetic field and solar activity. In order to turn 14C results into accurate calendar years, they must be calibrated using data from other sources. Annual tree rings (see discussion of dendrochronology below), foraminifera from stratified marine sediments, and microfossils from lake sediments can be used to chart the changes in 14C as “calibration curves.” The radiocarbon date obtained from the sample is compared to the established curve and then adjusted to reflect a more accurate calendar date (see Figure 7.24). The curves are updated over time with more data so that we can continue to refine radiocarbon dates (Törnqvist et al. 2016). The most recent calibration curves were released in 2020 and may change the dates for some existing sites by hundreds of years (Jones 2020).
As you will see in the hominin chapters (Chapters 9–12), 55,000 years is only a tiny fragment of human evolutionary history. It is insignificant in the context of the age of our planet. In order to date even older fossils, other methods are necessary.
Potassium-argon (K-Ar) dating and argon-argon (Ar-Ar) dating can reach further back into the past than radiocarbon dating. Used to date volcanic rock, these techniques are based on the decay of unstable potassium 40 (40K) into argon 40 (40Ar) gas, which gets trapped in the crystalline structures of volcanic material. It is a method of indirect dating. Instead of dating the fossil itself, K-Ar and Ar-Ar dates volcanic layers around the fossil. It will tell you when the volcanic eruption that deposited the layers occurred. This is where stratigraphy becomes important. The date of the surrounding layers can give you a minimum and maximum age of the fossil based on where it is in relation to those layers. This technique was used at Gesher Benot Ya'aqo in the Jordan Valley, dating early stratigraphic deposits of basalt flows to 100,000 years old (Bar-Yosef and Belmaker 2011). The site is unique because early layers of occupation with an Acheulean handaxe industry were made primarily of basalt, which is an uncommon material for this tool technology (see Chapter 10 for a full discussion of this tool technology). The benefit of this dating technique is that 40K has a half-life of circa 1.3 billion years, so it can be used on sites as young as 100 kya and as old as the age of Earth. As you will see in later chapters, it is particularly useful in dating early hominin sites in Africa (Michels 1972, 120; Renfrew and Bahn 2016, 155). Another benefit to this technique is that it does not damage precious fossils because the samples are taken from the surrounding rock instead. However, this method is not without its flaws. A study by J. G. Funkhouser and colleagues (1966) and Raymond Bradley (2015) demonstrated that igneous rocks with fluid inclusions, such as those found in Hawai‘i, can release gasses including radiogenic argon when crushed, leading to incorrectly older dates. This is an example of why it is important to use multiple dating methods in research to detect anomalies.
Uranium series dating is based on the decay chain of unstable isotopes of uranium. It uses mass spectrometry to detect the ratios of uranium 238 (238U), uranium 234(234U), and thorium 230 (230Th) in carbonates (Wendt et al. 2021). Thorium accumulates in the carbonate sample through radiometric decay. Thus, the age of the sample is calculated from the difference between a known initial ratio and the ratio present in the sample to be dated. This makes uranium series ideal for dating carbonate rich deposits such as carbonate cements from glacial moraine deposits, speleothems (deposits of secondary minerals that form on the walls, floors, and ceilings of caves, like stalactites and stalagmites), marine and lacustrine carbonates from corals, caliche, and tufa, as well as bones and teeth (University of Arizona, n.d.; van Calsteren and Thomas 2006). Due to the timing of the decay process, this dating technique can be used from a few years up to 650k (Wendt et al. 2021). Since many early hominin sites occur in cave environments, this dating technique can be very powerful. This method has also been used to develop more accurate calibration curves for radiocarbon dating. However, the accuracy of this method depends on knowing the initial ratios of the elements and ruling out possible contamination (Wendt et al. 2021). It also involves the destruction of a small sample of material.
Fission track dating is another useful dating technique for sites that are millions of years old. This is based on the decay of radioactive uranium 238 (238U). The unstable atom of 238U fissions at a predictable rate. The fission takes a lot of energy and causes damage to the surrounding rock. For instance, in volcanic glasses we can see this damage as trails in the glass. Researchers in the lab take a sample of the glass and count the number of fission trails using an optical microscope. As 238U has a half-life of 4,500 million years, it can be used to date rock and mineral material starting at just a few decades and extending back to the age of Earth. As with K-Ar, archaeologists are not dating artifacts directly. They are dating the layers around the artifacts in which they are interested (Laurenzi et al. 2007).
Luminescence dating, which includes thermoluminescence and a related technique called optically stimulated luminescence, is based on the naturally occurring background radiation in soils. Pottery, baked clay, and sediments that include quartz and feldspar are bombarded by radiation from the soils surrounding it. Electrons in the material get displaced from their orbit and trapped in the crystalline structure of the pottery, rock, or sediment. When a sample of the material is heated to 500°C (thermoluminescence) or exposed to particular light wavelengths (optically stimulated luminescence) in the laboratory, this energy gets released in the form of light and heat and can be measured (Cochrane et al. 2013; Renfrew and Bahn 2016, 160). You can use this method to date artifacts like pottery and burnt flint directly. When attempting to date fossils, you may use this method on the crystalline grains of quartz and feldspar in the surrounding soils (Cochrane et al. 2013). The important thing to remember with this form of dating is that heating the artifact or soils will reset the clock. The method is not necessarily dating when the object was last made or used but when it was last heated to 500°C or more (pottery) or exposed to sunlight (sediments). Luminescence dating can be used on sites from less than 100 years to over 100,000 years (Duller 2008, 4). As with all archaeological data, context is crucial to understanding the information.
Like thermoluminescence dating, electron spin resonance dating is based on the measurement of accumulated background radiation from the burial environment. It is used on artifacts and rocks with crystalline structures, including tooth enamel, shell, and rock—those for which thermoluminescence would not work. The radiation causes electrons to become dislodged from their normal orbit. They become trapped in the crystalline matrix and affect the electromagnetic energy of the object. This energy can be measured and used to estimate the length of time in the burial environment. This technique works well for remains as old as two million years (Carvajal et al. 2011, 115–116). It has the added benefit of being nondestructive, which is an important consideration when dealing with irreplaceable material.
Not all chronometric dating methods are based on unstable isotopes and their rates of decay. There are several other methods that make use of other natural biological and geologic processes. One such method is known as dendrochronology (Figure 7.25), which is based on the natural growth patterns of trees. Trees create concentric rings as they grow; the width of those rings depends on environmental conditions and season. The age of a tree can be determined by counting its rings, which also show records of rainfall, droughts, and forest fires.
Tree rings can be used to date wood artifacts and ecofacts from archaeological sites. This first requires the creation of a profile of trees in a particular area. The Laboratory of Tree-Ring Research at the University of Arizona has a comprehensive and ongoing catalog of tree profiles (see University of Arizona n.d.). Archaeologists can then compare wood artifacts and ecofacts with existing timelines, provided the tree rings are visible, and find where their artifacts fit in the pattern. Dendrochronology has been in use since the early 20th century (Dean 2009, 25). The Northern Hemisphere chronology stretches back nearly 14,000 years (Reimer et al. 2013, 1870) and has been used successfully to date southwestern U.S. sites such as Pueblo Bonito and Aztec Ruin (Dean 2009, 26). Dendrochronological evidence has helped calibrate radiocarbon dates and even provided direct evidence of global warming (Dean 2009, 26–27).
In Australia, dendrochronology, along with other environmental reconstruction methods, has been used to show that the Indigenous people had sophisticated land management systems before the arrival of British invaders. According to the work of Michael-Shawn Fletcher and colleagues (2021), there was a significant encroachment of the rainforests and tree species into grasslands after the British invasion. Prior to this time, Indigenous people managed the landscape through controlled burns at regular intervals. This practice created climate-resistant grasslands that were biodiverse and provided predictable food supplies for humans and other animals. Under European land management, there have been negative impacts on biodiversity and climate resilience and an increase in catastrophic wildfires (Fletcher et al. 2021).
This dating method does have its difficulties. Some issues are interrupted ring growth, microclimates, and species growth variations. This is addressed through using multiple samples, statistical analysis, and calibration with other dating methods. Despite these limitations, dendrochronology can be a powerful tool in dating archaeological sites (Hillam et al. 1990; Kuniholm and Striker 1987).
Environmental Reconstruction
As you read in Chapter 2, Charles Darwin, Jean-Baptiste Lamarck, Alfred Russel Wallace, and others recognized the importance of the environment in shaping the evolutionary course of animal species. To understand what selective processes might be shaping evolutionary change, we must be able to reconstruct the environment in which the organism was living.
One of the ways to do that is to look at the plant species that lived in the same time range as the species in which you are interested. One way to identify ancient flora is to analyze sediment cores from water and other protected sources. Pollen gets released into the air and some of that pollen will fall on wetlands, lakes, caves, and so forth. Eventually it sinks to the bottom of the lake and forms part of the sediment. This happens year after year, so subsequent layers of pollen build up in an area, creating strata. By taking a core sample and analyzing the pollen and other organic material, an archaeologist can build a timeline of plant types and see changes in the vegetation of the area (Hester et al. 1997, 284). This can even be done over large areas by studying ocean bed cores, which accumulate pollen and dust from large swaths of neighboring continents.
While sediment coring is one of the more common ways to reconstruct past environments, there are a few other methods. These have been recently employed at Holocene Lake Ivanpah, a paleolake that straddles the California and Nevada border in the United States. This lake was originally thought to have been completely dry around 9,300–7,800 kya (Sims and Spaulding 2017). However, analyzing core samples using soil identification, sediment chemistry, subsurface stratigraphy, and geomorphology (the study of the physical characteristics of the Earth’s surface) revealed deposition of three recent lake fillings during this period in the forms of additional hardpan, or lake bottom, playas, bedded or layered fine-grained (wetland) sediments, and buried beaches below the surface (Sims and Spaulding 2017; Spaulding and Sims 2018). These discoveries are important because they have not been integrated into interpretation of the local archaeological record, as it was assumed that the lake had been dry for thousands of years. Sedimentological analyses such as coring and those listed above can provide great insight into past climates and are accomplished in a minimally destructive way.
Another way of reconstructing past environments is by using stable isotopes. Unlike unstable isotopes, stable isotopes remain constant in the environment throughout time. Plants take in the isotopes through photosynthesis and ground water absorption. Animals take in isotopes by drinking local water and eating plants. Stable isotopes can be powerful tools for identifying where an organism grew up and what kind of food the organism ate throughout its life. They can even be used to identify global temperature fluctuations.
Global Temperature Reconstruction
Oxygen isotopes are a powerful tool in tracking global temperature fluctuations throughout time. The isotopes of Oxygen 18 (18O) and Oxygen 16 (16O) occur naturally in Earth’s water. Both are stable isotopes, but 18O has a heavier atomic weight. In the normal water cycle, evaporation takes water molecules from the surface to the atmosphere. Because 16O is lighter, it is more likely to be part of this evaporation process. The moisture gathers in the atmosphere as clouds that eventually may produce rain or snow and release the water back to the surface of the planet. During cool periods like glacial periods (ice ages), the evaporated water often comes down to Earth’s surface as snow. The snow piles up in the winter but, because of the cooler summers, does not melt off. Instead, it gets compacted and layered year after year, eventually resulting in large glaciers or ice sheets covering parts of Earth. Since 16O, with the lighter atomic weight, is more likely to be absorbed in the evaporation process, it gets locked up in glacier formation. The waters left in oceans would have a higher ratio of 18O during these periods of cooler global temperatures (Potts 2012, 154–156; see Figure 7.26).
The microorganisms that live in the oceans, foraminifera, absorb the water from their environment and use the oxygen isotopes in their body structures. When these organisms die, they sink to the ocean floor, contributing to the layers of sediment. Scientists can extract these ocean cores and sample the remains of foraminifera for their 18O and 16O ratios. These ratios give us a good approximation of global temperatures deep into the past. Cooler temperatures indicate higher ratios of 18O (Potts 2012, 154–156).
Diet Reconstruction
You may be familiar with the saying “you are what you eat.” When it comes to your teeth and bones, this adage is literal. Stable isotopes can also be used to reconstruct animal diet and migration patterns. Living organisms absorb elements from ingested plants and water. These elements are used in tissues like bones, teeth, skin, hair, and so on. By analyzing the stable isotopes in the bones and teeth of humans and other animals, we can identify the types of food they ate at different stages of their lives.
Plants take in carbon dioxide from the atmosphere during photosynthesis. We’ve already discussed this using the example of the unstable isotope 14C; however, this absorption also takes place with the stable isotopes of 12C and 13C. During photosynthesis, some plants incorporate carbon dioxide as a three-carbon molecule (C3 plants) and some as a four-carbon molecule (C4 plants). On the one hand, C3 plants include certain types of trees and shrubs that are found in relatively wet environments and have lower ratios of 13C compared to 12C. C4 plants, on the other hand, include plants from drier environments like savannahs and grasslands. C4 plants have higher ratios of 13C to 12C than C3 plants (Renfrew and Bahn 2016, 312). These ratios remain stable as you go up the food chain. Therefore, you can analyze the bones and teeth of an animal to identify the 13C/12C ratios and identify the types of plants that animal was eating.
The ratios of stable nitrogen isotopes 15N and 14N can also give information about the diet of fossilized or deceased organisms. Though initially absorbed from water and soils by plants, the nitrogen ratios change depending on the primary diet of the organism. An animal who has a mostly vegetarian diet will have lower ratios of 15N to 14N, while those further up the food chain, like carnivores, will have higher ratios of 15N. Interestingly, breastfeeding infants have a higher nitrogen ratio than their mothers, because they are getting all of their nutrients through their mother’s milk. So nitrogen can be used to track life events like weaning (Jay et al. 2008, 2). A marine versus terrestrial diet will also affect the nitrogen signatures. Terrestrial diets have lower ratios of 15N than marine diets. In the course of human evolution, this type of analysis can help us identify important changes in human nutrition. It can help anthropologists figure out when meat became a primary part of the ancient human diet or when marine resources began to be used. The ratios of stable nitrogen isotopes can also be used to determine a change in status, as in the case of the Llullaillaco children (the “ice mummies”) found in the Andes Mountains. For instance, the nitrogen values in hair from the Llullaillaco Maiden showed a significant positive shift that is associated with increased meat consumption in the last 12 months of her life (Wilson et al. 2007). Although the two younger children had little changes in their diets in the last year of their short lives, the changes in their nitrogen values were significant enough to suggest that the improvement in their diets may have been attributed to the Incas’ desire to sacrifice healthy, high-status children” (Faux 2012, 6).
Migration
Stable isotopes can also tell us a great deal about where an individual lived and whether they migrated during their lifetime. The geology of Earth varies because rocks and soils have different amounts or ratios of certain elements in them. These variations in the ratios of isotopes of certain elements are called isotopic signatures. They are like a chemical fingerprint for a geographical region. These isotopes get into the groundwater and are absorbed by plants and animals living in that area. Elements like strontium, oxygen, and nitrogen, among others, are then used by the body to build bones and teeth. If you ate and drank local water all of your life, your bones and teeth would have the same isotopic signature as the geographical region in which you lived.
However, many people (and animals) move around during their lifetimes. Isotopic signatures can be used to identify migration patterns in organisms (Montgomery et al. 2005). Teeth develop in early childhood. If the isotopes of teeth are analyzed, these isotopes would resemble those found in the geographic area where an individual lived as a child. Bones, however, are a different story. Bones are constantly changing throughout life. Old cells are removed and new cells are deposited to respond to growth, healing, activity change, and general deterioration. Therefore, the isotopic signature of bones will reflect the geographical area in which an individual spent the last seven to ten years of life. If an individual has different isotopic signatures for their bones and teeth, it could indicate a migration some time during their life after childhood.
Recent work involving stable isotope analysis has been done on the cremation burials from Stonehenge, in Wessex, England (Figure 7.27). Much of the archaeological work at Stonehenge in the past focused on the building and development of the monument itself. That is partly because most of the burials at the monument were cremated remains, which are difficult to study because of their fragmentary nature and the chemical alterations that bone and teeth undergo when heated. The cremation process complicates the oxygen and carbon isotopes. However, the researchers determined that strontium would not be affected by heating and could still be analyzed in cranial fragments. Using the remains of 25 individuals, they compared their strontium signatures to the geology of Wessex and other regions of the UK. Fifteen of those individuals had strontium signatures that matched the local geology. This means that in the last ten or so years of their lives, they lived and ate food from around Stonehenge. However, ten of the individuals did not match the local geologic signature. These individuals had strontium ratios more closely aligned with the geology of west Wales. Archaeologists find this particularly interesting because in the early phases of Stonehenge’s construction, the smaller “blue stones” were brought 200 km from Wales in a feat of early engineering. These larger regional connections show that Stonehenge was not just a site of local importance. It dominated a much larger region of influence and drew people from all over ancient Britain (Snoeck et al. 2018).
Special Topic: Cold Case Naia
In 2007, cave divers exploring the Sistema Sac Actun in the Yucatán Peninsula in Mexico (see Figure 7.28 and 7.29) discovered the bones of a 15- to 16-year-old female human along with the bones of various extinct animals from the Pleistocene (Collins et al. 2015). The site was named Hoyo Negro (“Black Hole”). The human bones belonged to a Paleo-American, later named “Naia” after a Greek water nymph. Examination of the partially fossilized remains revealed a great deal about Naia’s life, and the radiocarbon dating of her tooth enamel indicated that she lived some 13,000 years ago (Chatters et al. 2014). Naia’s arms were not overly developed, so her daily activities did not involve heavy carrying or grinding of grain or seeds. Her legs, however, were quite muscular, implying that Naia was used to walking long distances. Naia’s teeth and bones indicate habitually poor nutrition. There is evidence of violent injury during the course of Naia’s life from a healed spiral fracture of her left forearm. Naia also suffered from tooth decay and osteoporosis even though she appeared young and undersized. Dr. Jim Chatters hypothesizes that Naia entered the cave at a time when it was not flooded, probably looking for water. She may have become disoriented and fell off a high ledge to her death. The trauma to her pelvis is consistent with such an injury (Watson 2017).
Naia’s skeleton is remarkably complete given its age. As divers were able to locate her skull, Naia’s physical appearance in life could be interpreted. Surprisingly, in examining the skull, it was determined that Naia did not resemble modern Indigenous peoples in the region. However, the mitochondrial DNA (mtDNA) recovered from a tooth indicates that Naia shares her DNA with modern Indigenous peoples (Chatters et al. 2014). Though Naia’s burial environment made chemical analysis difficult, researchers were able to recover carbon isotopes from her remains. The isotopes from Naia’s tooth enamel suggest a diet of “cool-season grasses and/or broad-leaf vegetation” (Chatters et al. 2022, 68). Naia’s teeth also displayed numerous dental caries and only light dental wear. Coupled with the isotopic data, she likely had a “softer, more sugar-rich diet” (Chatters et al. 2022, 68).
Summary
With a timeline that extends back some 4.6 billion years, Earth has witnessed continental drift, environmental changes, and a growing complexity of life. Fossils, the mineralized remains of living organisms, provide physical evidence of life and the environment on the planet over the course of billions of years. In order to better understand the fossil record, anthropologists rely on the collaboration of numerous academic fields and disciplines. Anthropologists use a variety of scientific methods, both relative and chronometric, to analyze fossils to determine age, origins, and migration patterns as well as to provide insight into the health and diet of the fossilized organism. While each method has its advantages, disadvantages, and limited applications, these tools enable anthropologists to theorize how all living organisms evolved, including the evolution of early humans into modern humans, H. sapiens. The fossil record is far from complete, but our expanding understanding of the fossil context, with exciting new discoveries and improved scientific methods, enables us to document the history of our planet and the evolution of life on Earth.
Dating Methods Quick Guide
Method |
Material |
Effective date range |
Stratigraphy |
Soil layers |
Relative |
Biostratigraphy |
Plant and animal remains |
Relative |
Cultural dating |
Human-made objects |
Relative |
Fluorine |
Bone, antler, teeth |
Relative |
Radiocarbon |
Organic carbon bearing material (bones, teeth, antler, plant material, shell, charcoal) |
Younger than 55,000 years |
Potassium-argon and argon-argon |
Volcanic rock |
Older than 100,000 years |
Uranium series |
Carbonates such as stalactites, stalagmites, corals, caliche, and tufa |
Younger than 650,000 years |
Fission track |
Volcanic glasses and crystalline minerals |
Spans age of Earth |
Luminescence |
Pottery, baked clay, sediments |
100 to older than 100,000 years |
Electron spin resonance dating |
Tooth enamel, shell, rock with crystalline structures |
Younger than 2 million years |
Dendrochronology |
Wood (where tree rings are identifiable) |
Dependent on location and available chronologies |
Review Questions
- How do remains become fossils? What conditions are necessary for the fossilization process?
- What kind of information could you acquire from a single fossil? What could it tell you about the broader environment?
- What factors would you take into consideration when deciding which dating method to use for a particular artifact?
- What methods do anthropologists use to reconstruct past environments and lifestyles?
Key Terms
Anaerobic: An oxygen-free environment.
Anthropocene: The proposed name for our current geologic epoch based on human-driven climate change.
Argon-argon (Ar-Ar) dating: A chronometric dating method that measures the ratio of argon gas in volcanic rock to estimate time elapsed since the volcanic rock cooled and solidified. See also potassium-argon dating.
Atom: A small building block of matter.
Bezoars: Hard, concrete-like substances found in the intestines of fossil creatures.
Biostratigraphy: A relative dating method that uses other plant and animal remains occurring in the stratigraphic context to establish time depth.
Bya: Billion years ago.
Chronometric dating: Dating methods that give estimated numbers of years for artifacts and sites.
Continental drift: The slow movement of continents over time.
Coprolite: Fossilized poop.
Cultural dating: The relative dating method that arranges human-made artifacts in a time frame from oldest to youngest based on material, production technique, style, and other features.
Deep Time: James Hutton’s theory that the world was much older than biblical explanations allowed. This age could be determined by gradual natural processes like soil erosion.
Dendrochronology: A chronometric dating method that uses the annual growth of trees to build a timeline into the past.
Electron spin resonance dating: A chronometric dating method that measures the background radiation accumulated in material over time.
Element: Matter that cannot be broken down into smaller matter.
Eon: The largest unit of geologic time, spanning billions of years and divided into subunits called eras, periods, and epochs.
Epochs: The smallest units of geologic time, spanning thousands to millions of years.
Eras: Units of geologic time that span millions to billions of years and that are subdivided into periods and epochs.
Fission track dating: A chronometric dating method that is based on the fission of 283U.
Fluorine dating: A relative dating method that analyzes the absorption of fluorine in bones from the surrounding soils.
Foraminifera: Single-celled marine organisms with shells.
Fossilization: The process by which an organism becomes a fossil.
Fossils: Mineralized copies of organisms or activity imprints.
Geomorphology: The study of the physical characteristics of the Earth’s surface.
Glacial periods: Periods characterized by low global temperatures and the expansion of ice sheets on Earth’s surface.
Holocene: The geologic epoch from 10 kya to present. (See the discussion on “the Anthropocene” for the debate regarding the current epoch name.)
Hominin: The term used for humans and their ancestors after the split with chimpanzees and bonobos.
In matrix: When a fossil is embedded in a substance, such as igneous rock.
Isotopes: Variants of elements.
Kya: Thousand years ago.
Law of Superposition: The scientific law that states that rock and soil are deposited in layers, with the youngest layers on top and the oldest layers on the bottom.
Lithification: The process by which the pressure of sediments squeeze extra water out of decaying remains and replace the voids that appear with minerals from the surrounding soil and groundwater.
Luminescence dating: The chronometric dating method based on the buildup of background radiation in pottery, clay, and soils.
Megafauna: Large animals such as mammoths and mastodons.
Mitochondrial DNA: DNA located in the mitochondria of a cell that is only passed down from biological mother to child.
Mya: Million years ago.
Paleopathology: Study of ancient diseases and injuries identified through examining remains.
Periods: Geologic time units that span millions of years and are subdivided into epochs.
Permineralization: When minerals from water impregnate or replace organic remains, leaving a fossilized copy of the organism.
Petrified wood: A fossilized piece of wood in which the original organism is completely replaced by minerals through petrifaction.
Potassium-argon (K-Ar) dating: A chronometric dating method that measures the ratio of argon gas in volcanic rock to estimate time elapsed since the volcanic rock cooled and solidified. See also argon-argon dating.
Pseudofossils: Natural rocks or mineral formations that can be mistaken for fossils.
Radioactive decay: The process of transforming the atom by spontaneously releasing energy.
Radiocarbon dating: The chronometric dating method based on the radioactive decay of 14C in organic remains.
Relative dating: Dating methods that do not result in numbers of years but, rather, in relative timelines wherein some organisms or artifacts are older or younger than others.
Sediment cores: Core samples taken from lake beds or other water sources for analysis of their pollen.
Stable isotopes: Variants of elements that do not change over time without outside interference.
Stratigraphy: A relative dating method that is based on ordered layers or (strata) that build up over time.
Taphonomy: The study of what happens to an organism after death.
Trace fossils: Fossilized remains of activity such as footprints.
Uniformitarianism: The theoretical perspective that the geologic processes observed today are the same as the processes operating in the past.
Unstable isotopes: Variants of elements that spontaneously change into stable isotopes over time.
Uranium series dating: A radiometric dating method based on the decay chain of unstable isotopes of 238U and 235U.
About the Authors
Sarah S. King, Ph.D.
Cerro Coso Community College, sarah.king1@cerrocoso.edu
Dr. Sarah S. King is an anthropology/sociology professor at Cerro Coso Community College in California. She completed her Ph.D. work at the Division of Archaeological, Geographical and Environmental Sciences at the University of Bradford in West Yorkshire, England. Her thesis was entitled “What Makes War?: Assessing Iron Age Warfare through Mortuary Behavior and Osteological Patterns of Violence.” She also holds anthropology degrees from the University of California, Santa Cruz (B.A. hons., 2004), and the University of New Mexico, Albuquerque (M.A., 2006).
Kara Jones, M.A.
PhD student at University of Nevada, Las Vegas, jonesk44@unlv.nevada.edu
Kara Jones received their B.A. in anthropology at California State University, Bakersfield (2018) and their M.A. from University of Nevada, Las Vegas (2023, summer). Their master’s thesis is titled “Rockin’ at the Lake: Toolstone Use and Procurement along Holocene Lake Ivanpah, CA.” Mx Jones is a Mojave Desert archaeologist specializing in stone tool use and manufacture, focusing further on Holocene lakeshore adaptations.
For Further Exploration
Books
Bjornerud, Marcia. 2006. Reading the Rocks: The Autobiography of the Earth. New York: Basic Books.
Hazen, Robert M. 2013. The Story of Earth: The First 4.5 Billion Years, From Stardust to Living Planet. New York: Viking Penguin.
Holmes, Richard. 2010. The Age of Wonder: The Romantic Generation and the Discovery of the Beauty and Terror of Science. New York: Vintage.
Palmer, Douglas. 2005. Earth Time: Exploring the Deep Past from Victorian England to the Grand Canyon. New York: John Wiley & Sons.
Prothero, Donald R. 2015. The Story of Life in 25 Fossils: Tales of Intrepid Fossil Hunters and the Wonder of Evolution. New York: Columbia University Press.
Pyne, Lydia. 2016. Seven Skeletons: The Evolution of the World’s Most Famous Human Fossils. New York: Viking Books.
Repcheck, Jack. 2009. The Man Who Found Time: James Hutton and the Discovery of the Earth’s Antiquity. New York: Basic Books.
Taylor, Paul D., Aaron O’Dea. 2014. A History of Life in 100 Fossils. Washington, DC: Smithsonian Books.
Ward, David. 2002. Smithsonian Handbooks: Fossils. Washington, DC: Smithsonian Books.
Winchester, Simon. 2009. The Map That Changed the World: William Smith and the Birth of Modern Geology. New York: Harper Perennial.
Websites
East Tennessee State University Center of Excellence in Paleontology
Granger Historical Picture Archive
Indigenous Archaeology Collective
Natural History Museum (London), on Mary Anning
Petrified Forest National Park (NE Arizona)
Poozeum: The No. 2 Wonder of the World
Smithsonian National Museum of Natural History, Department of Paleobiology
Smithsonian National Museum of Natural History, on “What Does It Mean to be Human”
Society for American Archaeology, on “Ethics in Professional Archaeology”
Society for American Archaeology, “Principles of Archaeological Ethics”
References
Antoine, Pierre-Oliver, Maeva J. Orliac, Gokhan Atici, Inan Ulusoy, Erdal Sen, H. Evren Çubukçu, Ebru lbayrak, Neşe Oyal, Erkan Aydar, and Sevket Sen. 2012. “A Rhinocerotid Skull Cooked to Death in a 9.2 Mya-Old Ignimbrite Flow of Turkey.” PLoS ONE 7 (11): e49997.
Aufderheide, Arthur C. 2003. The Scientific Study of Mummies. Cambridge, UK: Cambridge University Press.
Bar-Yosef, O., and M. Belmaker. 2011. “Early and Middle Pleistocene Faunal and Hominins Dispersals through Southwestern Asia.” Quaternary Science Reviews 30 (11–12): 1318–1337.
Barras, C. 2022. “Lost Footprints of Our Ancestors.” New Scientist 254 (3381): 40–44.
Blong, John C., Martin E. Adams, Gabriel Sanchez, Dennis L. Jenkins, Ian D. Bull, and Lisa-Marie Shillito. 2020. “Younger Dryas and Early Holocene Subsistence in the Northern Great Basin: Multiproxy Analysis of Coprolites from the Paisley Caves, Oregon, USA.” Archaeological and Anthropological Sciences 12 (9): 1–29.
Boaz, Noel T., Russel L. Ciochon, Qinqi Xu, and Jinyi Liu. 2004. “Mapping and Taphonomic Analysis of the Homo erectus Loci at Locality 1 Zhoukoudian, China.” Journal of Human Evolution 46 (5): 519–549.
Booth, Thomas J., Andrew T. Chamberlain, and Mike Parker Pearson. 2015. “Mummification in Bronze Age Britain.” Antiquity 89 (347): 1,155–1,173.
Bradley, Raymond S. 2015. “Chapter 3: Dating Methods I.” In Paleoclimatology, edited by Raymond S. Bradley, 55–101. Cambridge, MA: Academic Press.
Brown, Theodore L., H. Eugene LeMay Jr., Bruce E. Burston, Catherine J. Murphy, Patrick M. Woodward, and Matthew Stoltzfus. 2018. Chemistry: The Central Science. New York: Pearson.
Campbell, Neil A., and Jane B. Reece. 2005. Biology 7th ed. New York: Pearson.
Carvajal, Eduar, Luis Montes, and Ovidio A. Almanza. 2011. “Quaternary Dating by Electron Spin Resonance (ESR) Applied to Human Tooth Enamel.” Earth Sciences Research Journal 15 (2): 115–120.
Chatters, James C., Joaquin Arroyo-Cabrales, and Pilar Luna-Erreguerena. 2022. “The Pre-Ceramic Skeletal Record of Mexico and Central America.” In The Routledge Handbook of Mesoamerican Bioarchaeology, edited by V. Tieslar, 49–74. New York: Routledge.
Chatters, James C., Douglas J. Kennett, Yemane Asmerom, Brian M. Kemp, Victor Polyak, Alberto Nava Blank, Patricia A. Beddows, et al. 2014. “Late Pleistocene Human Skeleton and mtDNA Link Paleoamericans and Modern Native Americans.” Science 344 (6185): 750–754.
Clough, Sharon. 2020. "Ethics in Human Osteology." The Archaeologist 109 (2020): 3–5.
Cochrane, Grant W. G., Trudy Doelman, and Lyn Wadley. 2013. “Another Dating Revolution for Prehistoric Archaeology?” Journal of Archaeological Method and Theory 20 (1): 42–60.
Collins, S. V., E. G. Reinhardt, D. Rissolo, J. C. Chatters, A. Nava-Blank, and P. Luna-Erreguerena. 2015. “Reconstructing Water Level in Hoyo Negro, Quintana Roo, Mexico: Implications for Early Paleoamerican and Faunal Access.” Quaternary Science Reviews 124: 68–83.
Cook, Harold J. 1928. “Glacial Age Man in New Mexico.” Scientific American 139 (1): 38–40.
Cook, S. F., and H. C. Ezra-Cohn. 1959. “An Evaluation of the Fluorine Dating Method.” Southwestern Journal of Anthropology 15 (3): 276–290.
Cooper, Arnie. 2010. “Sticky Situation at the Tar Pits.” LA Weekly, May 27, 2010. https://www.laweekly.com/sticky-situation-at-the-tar-pits/.
Crompton, Robin H., Todd C. Pataky, Russell Savage, Kristiaan D’Août, Matthew R. Bennett, Michael H. Day, Karl Bates, Sarita Morse, and William I. Sellers. 2012. “Human-like External Function of the Foot, and Fully Upright Gait, Confirmed in the 3.66 Million Year Old Laetoli Hominin Footprints by Topographic Statistics, Experimental Footprint-Formation and Computer Simulation.” Journal of the Royal Society Interface 9 (69): 707–719. doi: 10.1098/rsif.2011.0258
Crutzen, Paul J., and Eugene F. Stoermer. 2000. “The ‘Anthropocene.’” Global Change Newsletter 41: 17–18.
Darwin, Charles. 1859. On the Origin of Species. London, UK: John Murray.
Dean, Jeffery S. 2009. “One Hundred Years of Dendroarchaeology: Dating, Human Behavior, and Past Climate.” In Tree-rings, Kings, and Old World Archaeology and Environment: Papers Presented in Honor of Peter Ian Kuniholm, edited by S. Manning and M. J. Bruce, 25–32. Oxford, UK: Oxbow Books.
Dolnick, Edward. 2011. The Clockwork Universe: Isaac Newton, the Royal Society, and the Birth of the Modern World. New York: HarperCollins.
Duller, G.A.T. 2008. Luminescence Dating: Guidelines on Using Luminescence Dating in Archaeology. Swindon, UK: English Heritage.
Eisenbeiss, Sabine. 2016. “Preserved in Peat: Decoding Bodies from Lower Saxony, Germany.” Expedition 58 (2): 18–21.
Emling, Shelley. 2009. The Fossil Hunter: Dinosaurs, Evolution, and the Woman Whose Discoveries Changed the World. New York: St. Martin’s Griffin.
Faux, Jennifer L. 2012. “Hail the Conquering Gods: Ritual Sacrifice of Children in Inca Society.” Journal of Contemporary Anthropology 3 (1): 1–15.
Feder, Kenneth L. 2017. The Past in Perspective: An Introduction to Human Prehistory. 7th ed. New York: Oxford University Press.
Fletcher, Michael-Shawn, Tegan Hall, and Andreas Nicholas Alexandra. 2021. “The Loss of an Indigenous Constructed Landscape Following British Invasion of Australia: An Insight into the Deep Human Imprint on the Australian Landscape.” Ambio 50 (1): 138–149.
Flowers, Paul, Klaus Theopold, Richard Langley, and William R. Robinson. 2018. Chemistry. Houston, TX: Openstax, Rice University.
Fumagalli, Matteo, Ida Moltke, Niels Grarup, Fernando Racimo, Peter Bjerregaard, Marit E. Jørgensen, Thorfinn S. Korneliussen, Pascale Gerbault, Line Skotte, Allan Linneberg, Cramer Christensen, Ivan Brandslund, Torben Jørgensen, Emilia Huerta-Sánchez, Erik B. Schmidt, Oluf Pedersen, Torben Hansen, Anders Albrechtsen, and Rasmus Nielsen. 2015. “Greenlandic Inuit show genetic signatures of diet and climate adaptation.” Science 349 (6254): 1343-1347.
Funkhouser, J. G., I. L. Barnes, and J. J. Naughton. 1966. ”Problems in the Dating of Volcanic Rocks by the Potassium-Argon Method.” Bull Volcanol 29: 709–717.
Giles, Melanie. 2020. Bog Bodies: Face to Face with the Past. Manchester, UK: Manchester University Press.
Goodrum, Matthew R., and Cora Olson. 2009. “The Quest for an Absolute Chronology in Human Prehistory: Anthropologists, Chemists, and the Fluorine Dating Method in Palaeoanthropology.” British Journal for the History of Science 42 (1): 95–114.
Granger, Darryl E., Ryan J. Gibbon, Kathleen Kuman, Ronald J. Clarke, Laurent Bruxelles, and Marc W. Caffee. 2015. “New Cosmogenic Burial Ages for Sterkfontein Member 2 Australopithecus and Member 5 Oldowan.” Nature 522 (7554): 85–88.
Granger Historical Picture Archive. 2018. Cro-Magnon Footprint. Image no. 0167868. Accessed March, 02, 2023. https://www.granger.com/results.asp?image=0167868&itemw=4&itemf=0001&itemstep=1&itemx=11&screenwidth=1085.
Gruhn, R., 2020. “Evidence Grows That Peopling of the Americas Began More Than 20,000 Years Ago.” Nature 584: 47–48.
Haddy, A., and A. Hanson. 1982. “Research Notes and Application Reports Nitrogen and Fluorine Dating of Moundville Skeletal Samples. Archaeometry 24 (1): 37–44.
Hajdas, I., P. Ascough, M. H. Garnett, S. J. Fallon, C. L. Pearson, G. Quarta, K. L. Spalding, H. Yamaguchi, and M. Yoneda. 2021. “Radiocarbon Dating.” Nature Reviews Methods Primers 1 (62): 1–26.
Hatala, Kevin G., Brigitte Demes, and Brian G. Richmond. 2016. “Laetoli Footprints Reveal Bipedal Gait Biomechanics Different from Those of Modern Humans and Chimpanzees.” Proceedings of the Royal Society B 283: 20160235. https://dx.doi.org/10.1098/rspb.2016.0235.
Hester, Thomas R., Harry J. Shafer, and Kenneth L. Feder. 1997. Field Methods in Archaeology, 7th ed. Mountain View, CA: Mayfield Publishing.
Hillam, J., C. M. Groves, D. M. Brown, M. G. L. Baillie, J. M. Coles, and B. J. Coles. 1990. “Dendrochronology of the English Neolithic.” Antiquity 64 (243): 210–220.
Holmes, Richard. 2010. The Age of Wonder: How the Romantic Generation Discovered the Beauty and Terror of Science. New York: Vintage.
Ingber, Sasha. 2012. “Volcano Eruption Baked Rare Rhino Fossil.” National Geographic News, November 30, 2012. Accessed July 25, 2018. https://www.nationalgeographic.com/science/article/121130-rare-rhino-fossil-created-by-volcanic-explosion.
Jay, Mandy, B. T. Fuller, Michael P. Richards, Christopher J. Knüsel, and Sarah S. King. 2008. “Iron Age Breastfeeding Practices in Britain: Isotopic Evidence from Wetwang Slack, East Yorkshire.” American Journal of Physical Anthropology 136 (3): 327–337.
Jones, Nicola. 2020. “Carbon Dating, the Archaeological Workhorse, Is Getting a Major Reboot.” Nature News, May 19, 2020. doi: https://doi.org/10.1038/d41586-020-01499-y
Klein, Richard G. 1999. The Human Career: Human Biological and Cultural Origins, 2nd ed. Chicago: University of Chicago Press.
Komar, Debra A., and Jane E. Buikstra. 2008. Forensic Anthropology: Contemporary Theory and Practice. New York: Oxford University Press.
Kuniholm, Peter Ian, and Cecil L. Striker. 1987. “Dendrochronological Investigations in the Aegean and Neighboring Regions, 1983–1986.” Journal of Field Archaeology 14 (4): 385–398.
Laurenzi, Marinella A., Maria Laura Balestrieri, Giulio Bigazzi, Julio C. Hadler Neto, Pedro J. Iunes, Pio Norelli, Massimo Oddone, Ana Maria Osorio Araya, and José G. Viramonte. 2007. “New Constraints on Ages of Glasses Proposed as Reference Materials for Fission-Track Dating.” Geostandards & Geoanalytical Research 31 (2): 105–124.
Lebatard, Anne-Elisabeth, Didier L. Bourlès, Philippe Duringer, Marc Jolivet, Régis Braucher, Julien Carcaillet, Mathieu Schuster, et al. 2008. “Cosmogenic Nuclide Dating of Sahelanthropus tchadensis and Australopithecus bahrelghazali: Mio-Pliocene Hominids from Chad.” Proceedings of the National Academy of Sciences 105 (9): 3226–3231.
Lyell, Charles. 1830–1833. Principles of Geology, Being an Attempt to Explain the Former Changes of the Earth’s Surface, by Reference to Causes Now in Operation. 3 vols. London: John Murray.
Maixner, Frank, Dmitrij Turaev, Amaury Cazenave-Gassiot, Marek Janko, Ben Krause-Kyora, Michael R. Hoopmann, Ulrike Kusebauch, et al. 2018. “The Iceman’s Last Meal Consisted of Fat, Wild Meat, and Cereals.” Current Biology Report 28 (14): 2348–2355.
Masao, Fidelis T., Elgidius B. Ichumbaki, Marco Cherin, Angelo Barili, Giovanni Boschian, David A. Lurino, Sofia Menconero, Jacopo Moggi-Cecchi, and Giorgio Manzi. 2016. “New Footprints from Laetoli (Tanzania) Provide Evidence for Marked Body Size Variation in Early Hominins.” eLife 5: e19568. https://elifesciences.org/articles/19568.
McDonough, Katelyn, Taryn Johnson, Ted Goebel, and Karl Reinhard. 2022. “Disease, Diet, and Thorny Headed Worm Infection in the Great Basin.” Paper presented at 50th Annual Meeting of the Nevada Archaeological Association, Tonopah, Nevada, April 22nd, 2022.
Michels, Joseph W. 1972. “Dating Methods.” Annual Review of Anthropology 1: 113–126.
Monastersky, Richard. 2015. “Anthropocene: The Human Age.” Nature 519 (7542): 144–147.
Montgomery, Janet, Jane A. Evans, Dominic Powlesland, and Charlotte A. Roberts. 2005. “Continuity or Colonization in Anglo-Saxon England? Isotope Evidence for Mobility, Subsistence Practice, and Status at West Heslerton.” American Journal of Physical Anthropology 126 (2): 123–138.
National Park Service. 2022. “Fossilized Footprints.” National Park Service website, February 1. Accessed May 31, 2022. https://www.nps.gov/whsa/learn/nature/fossilized-footprints.htm.
Pastoors, Andreas, Tilman Lenssen-Erz, Bernd Breuckmann, Tsamkxao Ciqae, Ui Kxunta, Dirk Rieke-Zapp, and Thui Thao. 2017. “Experience Based Reading of Pleistocene Human Footprints in Pech-Merle.” Quaternary International 430 (A): 155–162.
Potts, Richard. 2012. “Evolution and Environmental Change in Early Human Prehistory.” Annual Review of Anthropology 41: 151–167.
Raichlen, David A., and Adam D. Gordon. 2017. “Interpretation of Footprints From Site S Confirms Human-like Bipedal Biomechanics in Laetoli Hominins.” Journal of Human Evolution 107: 134–138.
Ravn, Morten. 2010. “Bronze and Early Iron Age Bog Bodies from Denmark.” Acta Archaeologica 81 (1): 112–113.
Reimer, Paula J., Edouard Bard, Alex Bayliss, J. Warren Beck, Paul G. Blackwell, Christopher Bronk Ramsey, Caitlin E. Buck, et al. 2013. “INTCAL13 and Marine13 Radiocarbon Age Calibration Curves 0–50,000 Years cal BP.” Radiocarbon 55 (4): 1869–1887.
Reinhard, Johan. 2006. Ice Maiden: Inca Mummies, Mountain Gods, and Sacred Sites in the Andes. Washington, DC.: National Geographic.
Reitz, Elizabeth J., and Elizabeth S. Wing. 1999. Zooarchaeology. Cambridge, UK: Cambridge University Press.
Renfrew, Colin, and Paul Bahn. 2016. Archaeology: Theories, Methods, and Practice, 7th ed. New York: Thames and Hudson.
Ruddiman, William F., Zhengtang Guo, Xin Zhou, Hanbin Wu, and Yanyan Yu. 2008. “Early Rice Farming and Anomalous Methane Trends.” Quaternary Science Reviews 27 (13–14): 1291–1295.
Schmidt, Gavin. 1999. “Science Briefs: Cold Climates, Warm Climates: How Can We Tell Past Temperatures?” National Aeronautics and Space Administration Goddard Institute for Space Studies. https://www.giss.nasa.gov/research/briefs/1999_schmidt_01/
Shillito, Lisa-Marie, John C. Blong, Eleanor J. Green, and Eline N. van Asperen. 2020a. “The What, How and Why of Archaeological Coprolite Analysis.” Earth-Science Reviews 207. https://doi.org/10.1016/j.earscirev.2020.103196.
Shillito, Lisa-Marie, Helen L. Whelton, John C. Blong, Dennis L. Jenkins, Thomas J. Connolly, and Ian D. Bull. 2020b. “Pre-Clovis occupation of the Americas Identified by Human Fecal Biomarkers in Coprolites from Paisley Caves, Oregon.” Science Advances 6 (29). https://www.science.org/doi/10.1126/sciadv.aba6404.
Sims, Douglas., and W. Spaulding. 2017. “Shallow Subsurface Evidence for Postglacial Holocene Lakes at Ivanpah Dry Lake: An Alternative Energy Development Site in the Central Mojave Desert, California, USA.” Journal of Geography and Geology 9 (1): 1–24. DOI: 10.5539/jgg.v9n1p1.
Slon, Viviane, Fabrizio Mafessoni, Benjamin Vernot, Cesare De Filippo, Steffi Grote, Bence Viola, Mateja Hajdinjak, Stéphane Peyrégne, Sarah Nagel, Samantha Brown, et al. 2018. “The Genome of the Offspring of a Neanderthal Mother and a Denisovan Father.” Nature 561 (7721): 113–116.
Smithsonian National Museum of Natural History. 2018. Laetoli Footprint Trails. Smithsonian National Museum of History, October 23. Accessed February 14, 2023. https://humanorigins.si.edu/evidence/behavior/footprints/laetoli-footprint-trails.
Snoeck, Christophe, John Pouncett, Philippe Claeys, Steven Goderis, Nadine Mattielli, Mike Parker Pearson, Christie Willis, Antoine Zazzo, Julia A. Lee-Thorp, and Rick J. Schulting. 2018. “Strontium Isotope Analysis on Cremated Human Remains from Stonehenge Support[sic] Links With West Wales.” Scientific Reports 8. https://www.nature.com/articles/s41598-018-28969-8.
Spaulding, W. G., and D. B. Sims. 2018. “A Glance to the East: Lake Ivanpah—An Isolated Southern Great Basin Paleolake. The 2018 Desert Symposium Field Guide and Proceedings, 121-131. https://www.researchgate.net/publication/340493333_A_glance_to_the_east_Lake_Ivanpah-_an_isolated_southern_Great_Basin_paleolake
Spray, Aaron. “La Brea Woman: Only (& Controversial) Human from La Brea Tar Pits.” The Travel, April 3, 2022. Accessed July 31, 2022. https://www.thetravel.com/what-to-know-of-the-la-brea-woman/.
Steffen, Will, Johan Rockström, Katherine Richardson, Timothy M. Lenton, Carl Folke, Diana Liverman, Colin P. Summerhayes, et al. 2018. “Trajectories of the Earth System in the Anthropocene.” PNAS 115 (33): 8252–8259.
Stodder, Ann L. W. 2008. “Taphonomy and the Nature of Archaeological Assemblages.” In Biological Anthropology of the Human Skeleton, edited by M. Anne Katzenberg and Shelley R. Saunders, 71–115. Hoboken, NJ: Wiley Blackwell.
Straus, Lawrence Guy. 1989. “Grave Reservations: More on Paleolithic Burial Evidence.” Current Anthropology 30 (5): 633–634.
Sullivan, Terry, and Harry Gifford. 1908. She Sells Sea-Shells. London: Francis, Day, and Hunter.
Taylor, Anthony, Jarod M. Hutson, Vaughn M. Bryant, and Dennis L. Jenkins. 2020. “Dietary Items in Early to Late Holocene Human Coprolites from Paisley Caves, Oregon, USA.” Palynology 44 (1): 12–23.
Taylor, Paul D., Aaron O’Dea. 2014. A History of Life in 100 Fossils. Washington, D.C.: Smithsonian Books.
Törnqvist, T. E., B. E. Rosenheim, P. Hu, and A. B. Fernandez. 2015. “Radiocarbon Dating and Calibration.” In Handbook of Sea-Level Research, edited by Ian Shennan, Antony J. Long, and Benjamin P. Horton, 347-360. Oxford, UK: John Wiley & Sons. https://doi.org/10.1002/9781118452547.ch23
University of Arizona. n.d. “Uranium-Thorium Dating: The Uranium 238 Decay Series.” Accessed November 21, 2022. https://www.geo.arizona.edu/Antevs/ecol438/uthdating.html
University of the Witwatersrand. 2017. “Little Foot Takes a Bow: South Africa’s Oldest and the World’s Most Complete Australopithecus Skeleton Ever Found, Introduced to the World.” ScienceDaily, December 6. Accessed February 14, 2023. https://www.sciencedaily.com/releases/2017/12/171206100104.htm.
van Calsteren, Peter, and Louise Thomas. 2006. “Uranium-Series Dating Applications in Natural Environmental Science.” Earth-Science Reviews 75 (1–4): 155–175.
Vanzetti, A., M. Vidale, M. Gallinaro, D. W. Frayer, and L. Bondioli. 2010. “The Iceman as a Burial.” Antiquity 84 (325): 681–692. https://doi.org/10.1017/S0003598X0010016X.
Verghese, Namrata. 2021. “What Is Necropolitics? The Political Calculation of Life and Death.” Teen Vogue. March 10, 2021. Accessed February 14, 2023. https://www.teenvogue.com/story/what-is-necropolitics.
Vidale, M., L. Bondioli, D. W. Frayer, M. Gallinaro, and A. Vanzetti. 2016. “Ötzi the Iceman.” Expedition 58 (2): 13–17. Accessed February 14, 2023. https://www.penn.museum/sites/expedition/otzi-the-iceman/.
Wade, Lizzie. 2021. "Footprints Support Claim of Early Arrival in the Americas." Science 373 (6562): 1426. Accessed February 14, 2023. https://www.sciencemagazinedigital.org/sciencemagazine/24_september_2021/MobilePagedArticle.action?articleId=1727132#articleId1727132.
Waters, Colin N., Jan Zalasiewicz, Anthony D. Barnosky, Alejandro Cearreta, Agieszka Galuszka, Juliana A. Ivar Do Sul, Catherine Jeandel, et al. 2016 “Is the Anthropocene Distinct from the Holocene?” Science 351 (6269): aad2622-1-10. DOI:10.1126/science.aad2622
Watson, Traci. 2017. “Ancient Bones Reveal Girl’s Tough Life in Early Americas.” Nature 544 (7648): 15–16.
Wendt, Kathleen, A., Xianglei Li,, and R. Lawrence Edwards. 2021. “Uranium-Thorium Dating of Speleothems.” Elements 17 (2): 87–92.
White, Tim D. 1986. “Cut Marks on the Bodo Cranium: A Case of Prehistoric Defleshing.” American Journal of Physical Anthropology 69 (4): 503–509.
Williams, Linda D. 2004. Earth Science Demystified. New York: McGraw-Hill Professional.
Wilson, Andrew S., Timothy Taylor, Maria Constanza Ceruti, Jose Antonio Chavez, Johan Reinhard, Vaughan Grimes, Wolfram Meier-Augenstein, et al. 2007. “Stable Isotope and DNA Evidence for Ritual Sequences in Inca Child Sacrifice.” PNAS 104 (42): 16456–16461.
Acknowledgments
We are grateful to Lee Anne Zajicek, who coauthored the first edition. Her original contributions continue to be an integral part of this chapter. We thank the staff of the Maturango Museum, Ridgecrest, California. Specifically, for their generous help with photography and fossil images, we acknowledge Debbie Benson, executive director; Alexander K. Rogers, former archaeology curator; Sherry Brubaker, natural history curator; and Elaine Wiley, history curator. We thank Sharlene Paxton, a librarian at Cerro Coso Community College, Ridgecrest, California, for her guidance and expertise with OER and open-source images, and John Stenger-Smith and Claudia Sellers from Cerro Coso Community College, Ridgecrest, California, for their feedback on the chemistry and plant biology content. Finally, we thank William Zajicek and Lauren Zajicek, our community college students, for providing their impressions and extensive feedback on early drafts of the chapter.