Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

When the Invasion of Land Failed: The Legacy of the Devonian Extinctions
When the Invasion of Land Failed: The Legacy of the Devonian Extinctions
When the Invasion of Land Failed: The Legacy of the Devonian Extinctions
Ebook500 pages7 hours

When the Invasion of Land Failed: The Legacy of the Devonian Extinctions

Rating: 3.5 out of 5 stars

3.5/5

()

Read preview

About this ebook

The invasion of land by ocean-dwelling plants and animals was one of the most revolutionary events in the evolution of life on Earth, yet the animal invasion almost failed -- twice -- because of the twin mass extinctions of the Late Devonian Epoch. Some 359 to 375 million years ago, these catastrophic events dealt our ancestors a blow that almost drove them back into the sea. If those extinctions had been just a bit more severe, spiders and insects might have become the ecologically dominant forms of animal life on land. This book examines the profound evolutionary consequences of the Late Devonian extinctions, which shaped the composition of the modern terrestrial ecosystem. Only one group of four-limbed vertebrates now live on Earth while other tetrapod-like fishes are extinct. This gap is why the idea of "fish with feet" seems so peculiar yet these animals were once a vital part of our world.

LanguageEnglish
Release dateNov 12, 2013
ISBN9780231536363
When the Invasion of Land Failed: The Legacy of the Devonian Extinctions

Related to When the Invasion of Land Failed

Related ebooks

Biology For You

View More

Related articles

Related categories

Reviews for When the Invasion of Land Failed

Rating: 3.5 out of 5 stars
3.5/5

2 ratings1 review

What did you think?

Tap to rate

Review must be at least 10 words

  • Rating: 4 out of 5 stars
    4/5
    For me, the question with this type of book is whether it's going to be advanced enough that I feel as though I'm learning something but not so technical that it's simply impenetrable. On the whole McGhee maintains a nice balance in what feels like a bit of a textbook for a junior grad student, meaning that this is not the first book one should read on the subject. What McGhee does really well is to make clear the gaps in our understanding when it comes to fitting existing hard data (in relation to fossils and geological structures) into plausible explanatory systems. The one bottom line is that it took the evolution of amniotic reproductive systems before terrestrial vertebrates really became viable.

Book preview

When the Invasion of Land Failed - George R. McGhee, Jr.

Preface

This book is about the profound evolutionary consequences of the twin extinction events that occurred in the Late Devonian period of geologic time, some 375 and 359 million years ago, for the evolution of terrestrial animal life on Earth. Our ancestors, the first vertebrates on land, were dealt a severe blow in their infancy by those catastrophic events, one that almost drove them back into the sea. If those extinctions had been just a bit more severe, spiders and insects might well have become the ecologically dominant forms of animal life on land—we, the vertebrates, would have remained in the oceans.

We know that five catastrophic losses of biological diversity have occurred in the history of life on Earth since the evolution of animals more than 600 million years ago. Each of these biodiversity crises, commonly called the Big Five, changed the direction and pattern of the evolution of life on Earth and had consequences that extend to our present world.

The most recent of these biodiversity crises is the end-Cretaceous extinction, which occurred 65 million years ago. The environmental catastrophe that triggered the end-Cretaceous extinction destroyed the dinosaurian ecosystem, a highly successful terrestrial ecosystem that had persisted for 150 million years. We mammals owe our current ecological dominance to that event. It was only after the destruction of the dinosaurian ecosystem that mammals, in a classic example of ecological release, evolved explosively and rapidly established the mammalian terrestrial ecosystem that persists to the present day. If that event had not occurred, the dinosaurs might have continued to reign supreme, and we might have continued to be small nocturnal animals that scurried in the underbrush at night, seeking to avoid the dinosaur predators who hunted by day.

The next most recent biodiversity crisis is the end-Triassic extinction, which occurred approximately 200 million years ago. If the end-Cretaceous extinction is seen as beneficial to mammals, the end-Triassic was the opposite. Our ancestors, the synapsid amniotes, were in fact more diverse than the ancestors of the dinosaurs prior to that catastrophe, but that changed after this crisis. In essence, the end-Triassic extinction ushered in the dominance of the dinosaurs, and the end-Cretaceous extinction terminated that dominance. What if the end-Triassic event had never happened? Might we, the mammals, have become the ecologically dominant land animals and have kept the dinosaurs in check, rather than the other way around?

The next most recent biodiversity crisis in geologic time is the catastrophic end-Permian extinction, which nearly ended animal life on the planet Earth about 250 million years ago. All forms of life, both terrestrial and marine, suffered in that global catastrophe. If it had been only a little more severe, it would have erased the previous 350 million years of animal evolution, leaving only the simplest animals, such as jellyfish and sponges, as survivors. As it was, this extinction event triggered a restructuring of the ecosystems of the entire planet, both on the landmasses and in the seas, and is recorded in the fossil record as the largest loss of biodiversity ever seen in geologic time.

This brings us back in time to the two remaining biodiversity crises of the Big Five: the Late Devonian and the end-Ordovician. The end-Ordovician did not have a great impact on life on land simply because there was not much life on land to be affected at that time. For much the same reason, it is often thought that the Late Devonian crisis also did not have much of an effect on terrestrial life. That point of view is profoundly wrong, however, as will be demonstrated in this book.

Our world is the product of the Devonian extinctions. Only one group of four-limbed vertebrates exists on the Earth today: we, the tetrapods. All of our close relatives, the once diverse lineages of the other tetrapod-like fishes, are extinct. Our more distant relatives, the once numerous lobe-finned fishes, are rare—only three families of air-breathing, lobe-finned lungfishes exist today. Before the Devonian extinctions, the sight of an air-breathing fish crawling along a riverbank—out of water—was not unusual, but the Devonian extinctions eliminated all of our intermediate vertebrate relatives, resulting in a large phylogenetic gap between us, the limbed vertebrates, and our very distant living relatives, the familiar ray-finned fishes that fill our rivers and lakes today. This gap in living intermediate forms is why the very idea of a fish with limbs and feet seems so peculiar to us today; yet such fishes existed in the Late Devonian. If the Devonian extinctions had not happened, such fishes might still have been commonplace today. No one would have conceived of creating a Darwin bumper sticker for their car that shows a fish with feet simply because a fish with feet would not have been unusual—you could see one at any local fish market.

This book is a summary of more than three decades of my research into the Devonian extinctions. I thank Columbia University Press, publishers of my initial book The Late Devonian Mass Extinction (1996), for making this second volume possible. I thank Kalliopi Monoyios, scientific artist, for her original illustrations of our ancient vertebrate ancestors and kin. And last, I thank my wife, Marae, for her patient love.

CHAPTER 1

The Evolution of Life on Land

The Long Beginning

The planet Earth is 4,560 million years old. For over 3,000 million years, life on Earth was confined to the oceans, and the land areas of the planet looked much like the rocky landscapes of Mars today. Complex life on land, and the terrestrial ecosystem, is a relatively new phenomenon in the history of the Earth. The oldest simple terrestrial ecosystems of macroscopic organisms—plants, fungi, and animals—can be traced back only to the middle of the Ordovician Period of geologic time (for the geologic timescale, see table 1.1),¹ some 468 million years ago. That is, only in the past 10 percent of the total age of the Earth has complex life evolved to the point where it could emerge from the oceans and conquer the hostile environment of the terrestrial realm.

It is not easy living on land. Dehydration is a very serious problem for land-dwelling organisms. Many of the morphological and physiological changes that took place in the evolutionary sequence of fish to amphibians to amniotes were adaptations for preventing water loss in the dry environments of the land, where animals are surrounded only by the thin gaseous layer of the atmosphere. Life in the sea is surrounded at all times by water; dehydration is an unknown phenomenon. Land-dwelling organisms must deal with the crushing force of gravity, whereas, for neutrally buoyant swimming fish out in the oceans, gravity is an unknown phenomenon. Life in the oceans experiences very little variation in temperature; in places where temperature does vary, the change is very slow, gently rising and falling with the passing of the four seasons of the year. In contrast, land-dwelling organisms deal with large fluctuations in temperature that occur very rapidly: every 24 hours, temperatures quickly increase as the fireball that is the sun rises above the horizon, and they fall just as quickly as the sun sets into the cold darkness of the night. Even today, under the protective screen of the Earth’s ozone layer in the upper atmosphere, land-dwelling organisms are subject to radiation burns from that same fireball of the sun—radiation burns that can lead to fatal skin cancers. For much of the history of the Earth, that ozone layer did not exist at all, and the intense ultraviolet radiation flux present on the continental landmasses was deadly. In contrast, water is a very good radiation shield, and marine organisms located just a few centimeters below the surface of the oceans are protected from ultraviolet radiation from the sun.

TABLE 1.1  The geologic timescale.

Source : Modified from Gradstein et al. (2004), Walker and Geissman (2009), and Erwin et al. (2011).

Abbreviations : Ma BP = millions of years before the present.

Thus land-dwelling organisms have had to evolve complex adaptations to deal with four particularly serious problems that do not exist for organisms in the oceans: dehydration, gravity, temperature fluctuations, and radiation poisoning. The last problem was initially the most serious; it took 3,000 million years of environmental change for the radiation flux present on land to decrease to livable levels and evolutionary change for organisms to adapt to life on land.

Life on Earth Today

Three types of life exist on Earth today: the archaea, the bacteria, and the eukarya (table 1.2). The simplest forms of life are the unicellular; that is, they exist as a single individual cell. There are two types of unicellular life (fig. 1.1): the prokaryotes (simple cells) and the eukaryotes (complex cells). Prokaryotes are the simplest forms of unicellular life: their small cells have no nucleus and no organelles but consist only of a cell membrane within which the chromosomes of the organism are enclosed. The chromosomes contain the DNA of the organism, the coding mechanism of life on Earth. For many years it was thought that there was only one type of prokaryote life on Earth, the bacteria. Then, in 1977, microbiologist Carl Woese discovered that another type of prokaryote life existed—the archaea—a discovery for which he received the Crafoord Prize.² The archaea and bacteria look very much like one another when viewed under a microscope, but their genetics are quite different. The easiest way to tell the two types of life apart is by examining their cell membranes: the bacteria have ester-bonded lipid membranes, and the archaea have ether-bonded lipid membranes. Because the possession of a cell membrane is one of the fundamental characteristics of life (the other being the possession of DNA), the difference in cell-membrane construction between bacteria and archaea is of major evolutionary significance.³

TABLE 1.2  A phylogenetic classification of life.

Source : Modified from the phylogenetic classification of Lecointre and Le Guyader (2006).

Note : Multicellular eukaryote clades are marked in bold-faced type.

FIGURE 1.1  Top: a prokaryote cell, the bacterium Rhizobium leguminosarum. Bottom: a eukaryote cell, the excavobiontan Euglena spirogyra. Credit: Modified and redrawn from Lecointre and Le Guyader (2006).

In our day-to-day lives we are surrounded by a sea of bacteria, but we usually do not notice them because they are so small that they are invisible to the naked eye. Some of these bacteria are very beneficial to us: the Escherichia coli (E. coli) bacteria that live in our intestines help us to digest food (but are beneficial only if they remain in our intestines!). Others can be deadly: Clostridium botulinum will kill us with botulism, and its cousin Clostridium perfringens will give us gangrene. For the most part, however, the bacteria around us are benign and simply coexist with us.

The archaea usually live in extreme environments; thus, in our normal lives we do not encounter them as often as we do the bacteria.⁴ Some, like Methanopyrus kandleri, live at the bottom of the oceans in volcanic hydrothermal vents, under incredible pressure and in water as hot as 110°C (230°F). Others, like Thermoplasma acidophilum, flourish in the waste pits of coal mines, in sulfuric acid–rich waters with a pH value as low as one.

In contrast to the prokaryotes, the eukaryote unicells are much larger and more complex forms of life (fig. 1.1). Eukaryotes possess a nucleus (which contains the chromosomes) and specialized organelles such as chloroplasts (which can produce food for the cell by photosynthesis), mitochondria (which mediate the energetic enzyme activities of the cell), and Golgi apparati (for processing proteins within the cell). The difference in size between the simple prokaryote unicells and complex eukaryote unicells is striking: the prokaryote bacterium Rhizobium leguminosarum illustrated in figure 1.1 is 2 micrometers (0.08 thousandths of an inch) long, and the eukaryote Euglena spirogyra is 120 micrometers (4.7 thousandths of an inch) long. That is, the single eukaryote cell is 60 times larger than the prokaryote cell!

It has been proved that the organelles of the eukaryote cell were once separate, independent bacteria. For example, the chloroplasts of a eukaryote cell were once free-living cyanobacteria, and the mitochondria were once free-living purple bacteria.⁵ Even today, the mitochondria within the cells in our bodies still possess their own DNA, which is separate from the DNA in the nucleus of our cells that codes for the cell’s reproduction. The evolution of the eukaryote cell from the prokaryote cell occurred by the process of biological symbiosis: first as colonies of cooperating independent cells, and then later as a complex single cell of dependent organelles. This evolutionary transition had occurred by 2,100 million years ago, and the geologic evidence for the occurrence of this event will be considered later in this chapter.

The eukarya are the third form of life on Earth (table 1.2), and they all possess eukaryote cells (hence their name). The eukarya split into two separate lineages, the bikonta and the unikonta, early in their evolutionary history (table 1.2). The unicellular forms of these two groups are easy to tell apart: the bikonta have two flagella, and the unikonta have only one flagellum. We are unikonts, and the single flagellum of the human sperm cell is an ancient inherited trait from our unicellular ancestors.

The eukarya include both unicellular and multicellular forms of life (table 1.2). Multicellular organisms evolved in the same fashion as the eukaryote cell itself, i.e., by biological symbiosis, in which free-living eukaryote cells first formed colonies of cooperating independent cells and then multicellular tissues composed of dependent and specialized cells. However, the evolution of multicellularity in the eukarya is convergent—it occurred at least three times independently.⁶ In the lineage of the bikonta, multicellularity occurred in the evolution of the metabionta, the red and green algae (table 1.2). In the lineage of the unikonta, multicellularity occurred twice. First, the lineage of the opisthokonta produced the fungi and their sister group the choanozoa. The clade of the fungi is divided into the microsporidia, which remained unicellular, and their sister group, the eumycetes, which evolved multicellularity. Second, the clade of the choanozoa is divided into the choanoflagellata, which remained unicellular, and their sister group the metazoa, which independently evolved multicellularity (table 1.2). Thus the trees, mushrooms, and humans we encounter in our daily lives are all independent inventions of multicellularity by the ancient unicellular eukarya. The bikonta were the first to evolve multicellularity, producing the green algae some 1,200 million years ago, with the unikonta evolving multicellular animals only 780 million years ago, as will be discussed later in this chapter.

The Earliest Life on Earth

It is not known how life originated on Earth. We do know that the most ancient fossils of life yet discovered—simple, prokaryote cells like those found in living archaea and bacteria—are 3,450 million years old. Geochemical evidence suggests that life was present on Earth 3,830 million years ago, though actual fossil cells have yet to be found that are that old.

The first item of geochemical evidence for the presence of life 3,830 million years ago comes from the analysis of carbon isotopes present in the ancient strata. The carbon atom has two stable isotopes, carbon-12 and carbon-13. Carbon-12 possesses six neutrons and six protons in the nucleus of the atom, and the heavier carbon-13 possesses seven neutrons and six protons in its nucleus (carbon has another, heavier isotope, carbon-14, which is unstable and radioactively decays to nitrogen-14). Photoautotrophic organisms prefer the lighter isotope of carbon in photosynthesis; that is, photosynthetic bacteria preferentially extract carbon-12 from the atmosphere to synthesize hydrocarbons and leave the heavier isotope carbon-13 free in the environment. The naturally occurring ratio of carbon-12 to carbon-13 in the atmosphere in the absence of life is well known. Therefore, the isotopic ratio of carbon in ancient strata can be used as a tool to search for the presence of life: if the ratio of carbon-12 to carbon-13 is not the naturally occurring, inorganically-produced ratio in the atmosphere, then some other process is changing that ratio by preferentially removing carbon-12 from the atmosphere. Life is one such process. The carbon-isotope ratios in 3,830 million-year-old strata from Greenland are off; something skewed the ratio of carbon-12 to carbon-13, and this is taken as geochemical evidence that life was present.

The second item of geochemical evidence comes from the presence of banded-iron formations in Archaean strata 3,830 million years old. The early atmosphere of the Earth was composed mostly of carbon dioxide, much like the present-day atmospheres of Mars and Venus. The present-day atmosphere of the Earth, in contrast, contains about 21 percent molecular oxygen. That oxygen is the product of organic photosynthesis: plants and other aerobic photosynthetic organisms remove carbon dioxide from the atmosphere and combine it with hydrogen from water to produce sugars and other hydrocarbons. In this process, free oxygen is released as a waste product of aerobic photosynthesis. These organisms dump oxygen back into the environment.⁸ The earliest organisms to evolve photosynthesis, however, were not aerobic cyanobacteria but anaerobic purple bacteria.⁹ Phylogenetic analyses of the five groups of photosynthetic bacteria reveal that the purple bacteria are basal and that they first evolved anaerobic photosynthesis, followed by the evolution of the two types of green bacteria,¹⁰ the heliobacteria,¹¹ and finally the evolution of aerobic photosynthesis in the cyanobacteria.¹²

Banded-iron formations contain iron that has been oxidized. How could iron be oxidized on an early Earth that had an atmosphere that was anoxic, i.e., was mostly carbon dioxide? One potential way to oxidize iron is in the process of photosynthesis; that is, life. Some of the anaerobic-photosynthetic purple bacteria directly oxidize iron, whereas the aerobic-photosynthetic cyanobacteria produce free oxygen that then secondarily oxidizes iron. Thus the presence of oxidized iron in the banded-iron formations in strata that are 3,830 million years old is taken as geochemical evidence that life was present. Geochemical analyses of the types of iron isotopes present in these most ancient banded-iron formations indicate that anaerobic purple bacteria were responsible for the oxidation of the iron, not aerobic cyanobacteria.¹³ Our oldest fossil evidence for the aerobic-photosynthetic cyanobacteria indicates that they evolved later, nearly 400 million years later.¹⁴

Sometime and somewhere in the 730 million years between the formation of the Earth 4,560 million years ago and our oldest geochemical evidence for the presence of life at 3,830 million years ago, life originated on the planet. Or did it? Some of our modern theories about how life might have originated propose that life may in fact have originated on Mars first, and then was seeded on Earth by Martian meteorites—rocks that had been blasted off the surface of Mars by asteroid impacts, which then fell to the Earth’s surface, bringing life with them. If that scenario is ever proved, then we are all in fact Martians. We have found quite a few Martian meteorites here on Earth, and the most ancient of these—the famous 4,510 million-year-old ALH84001 meteorite—contains tiny cylindrical structures that look very much like fossil nanobacteria, embedded in carbonate deposits within the meteorite that are 3,900 million years old (fig. 1.2).¹⁵ These structures, if fossil bacteria, are smaller than any nanobacteria living on Earth today; for example, the one shown in figure 1.2 is only 480 nanometers (0.019 millionths of an inch) long. Alternatively, it has been suggested that these tiny structures may be fossil precursors of prokaryote cellular life, or protobacteria—actual preserved evidence of one of the steps in the evolution of cellular life.¹⁶ In that case, these structures could yield critical evidence about the origin of life itself, and their existence would be even more compelling evidence to go to Mars to find more of them.

FIGURE 1.2  One of numerous enigmatic cylindrical structures that look very similar to fossil prokaryote cells, found in the Martian meteorite ALH84001. Credit: Photography courtesy of NASA.

Scenarios for the origin of life, whether on Earth or on Mars, all consider life to have originated in water. Earlier scenarios considered it probable that life originated in shallow marine waters, and that the energy source necessary for assembling and rearranging molecules was provided by electric discharges in lightning or by the high ultraviolet radiation flux at the surface of the water, or both. More recent scenarios consider it more likely that life originated in deep marine waters around volcanic hydrothermal vents, and that the energy source was both thermal and chemical.¹⁷ Even today, most of the microbial forms of life found around deep-water hydrothermal vents are archaea, an ancient lineage of life that we considered in the previous section of the chapter. We may in fact have to go to Mars to hunt for the earliest evidence for the evolution of life. The ancient strata that we need to examine simply do not exist on Earth any longer. The surface of the Earth is a very dynamic environment: the gigantic plates of the Earth’s crust are constantly shifting around, spreading apart at the mid-oceanic ridges and grinding together to form mountain chains in other areas of the planet. The atmosphere is constantly convecting; water evaporates into the atmosphere in some regions of the Earth and condenses in torrential rainfall in other regions. Many of the Earth’s dynamic processes destroy sedimentary rocks: they are crushed and metamorphosed in uplifted mountain chains, they are subducted in the oceanic trenches and melted into lava, they are eroded by the rainfall and washed away in the rivers. Not surprisingly, the further we go back in geologic time, the fewer strata are preserved in the rock record. The oldest surviving sedimentary rocks on Earth that could possibly contain evidence of life are 3,830 to 3,850 million years old.¹⁸ We know older sedimentary rocks must have existed on Earth, but they are now long gone, destroyed by Earth’s dynamic surface processes. How can we answer the question of how life originated when no evidence remains to be examined on Earth in the critical time interval from 4,560 to 3,850 million years ago?

In contrast to the Earth, we know that 4,510 million-year-old crustal rocks still exist on Mars: we actually have a piece of them, the Martian meteorite ALH84001.¹⁹ The surface of Mars is much less dynamic than that of the Earth; the plates of the Martian crust stopped moving long ago, and most of the Martian volcanoes have ceased erupting molten rock. Mars is a smaller world than the Earth, and it lost its heat at a much faster rate than the Earth. The only major destroyer of sedimentary rocks on Mars today are the winds, which continue to etch away the ancient strata. Did life originate on Mars? If it did, ancient sedimentary rocks that could hold their fossils still exist on Mars.

It is important to realize that the origin of life and the evolution of life are two distinctly separate questions. If we ask the question, How did life originate?, the answer is: we do not know. We do have some good ideas for how life originated, but we honestly do not know for certain how life originated. However, if we ask the question, Does life evolve?, the answer is: yes. That is an empirical observation; it is a fact. There is nothing theoretical about the fact that life evolves. Theory only enters the equation when we ask the additional question, Why does life evolve? Charles Darwin’s theory of natural selection is one theory that has been proposed to explain why life evolves; another theory is that of neutral change and genetic drift, and so on. Continued debate about potential causes of evolution does not bring into question the fundamental empirical observation that life evolves.

Evolution in the Archaean and Proterozoic Eons

What are the facts, the empirical observations, about the early evolution of life on Earth? Our oldest (yet) discovered microfossils of prokaryote cells and macrofossil stromatolites are 3,450 million years old.²⁰ Stromatolites are conical, mound-like layered structures of sediment created by masses of filamentous bacteria that secrete mucilage slime (fig. 1.3). Modern-day cyanobacteria typically form stromatolites if they are allowed to grow undisturbed by grazing marine animals, such as snails, who consume the bacterial mats as food. Thus stromatolites today are found only in extreme marine environments, such as hypersaline lagoons where the salinity of the sea water is so high that it is lethal to complex animals. In the Archaean, however, there were no animals, and stromatolites formed in a variety of normal marine environments.

It is not known with certainty what type of prokaryotes produced the Archaean stromatolites that are found in more than 30 known sites around the Earth. It is generally assumed that at least some of these stromatolites, if not all of them, were formed by cyanobacteria because these bacteria continue to form stromatolites today.²¹ Cyanobacteria are aerobic photosynthesizers; that is, they produce free oxygen as a metabolic waste product when they form sugars by photosynthesis, and during the Archaean this process had a profound impact on future evolution of life on Earth.

FIGURE 1.3  A scene from the Archaean Earth, with mound-like stromatolites in the tidal zone. Stromatolites are produced by filamentous cyanobacteria that secrete mucilage slime; they are rare today but were abundant and the only visible form of life on the Earth in the Archaean. Credit: From Evolution: What the Fossils Say and Why It Matters, by D. R. Prothero, copyright © 2007 Columbia University Press; figure by Carl Buell. Reprinted with permission.

Molecular evidence from Australia has been used to argue that the cyanobacteria were present in the Earth’s oceans 2,700 million years ago and—more controversial—that the eukarya were present as well.²² This evidence is the presence of distinctive cyanobacterial biomarker chemicals in ancient sediments that are argued to have undergone very little metamorphism for 2,700 million years (metamorphism will destroy organic molecules); these biomarkers have been called molecular fossils because actual cellular fossils are absent. These fragile biomarker chemicals are surprising in that they have survived for 2,700 million years, as the previously discussed 3,450 million-year-old stromatolite fossils are generally considered to have been produced by cyanobacteria (table 1.3). What is more surprising is the argued ancient presence of other distinctive biomarkers that are only found in the cell membranes of the eukarya. Our oldest eukaryote cellular fossils are 2,100 million years old—thus, the presence of these molecular fossils would mean that the symbiotic evolution of eukaryote cells actually occurred some 600 million years earlier (table 1.3). If the eukarya had evolved by 2,700 million years ago, then their presence would have important implications for the oxygen content of the Earth’s atmosphere. Modern eukaryote cells need an ambient oxygen concentration of 0.042 percent to 0.21 percent in the atmosphere²³ to synthesize these biomarker molecules, which are components of their cell membranes. After 750 million years, since the evolution of the cyanobacteria some 3,450 million years ago, oxygen had finally reached high enough concentrations in sea water that it began to diffuse into the atmosphere.

The atmosphere of the Earth has been radically transformed by the presence of life. Both anaerobic and aerobic photosynthesizers actively remove carbon dioxide from the atmosphere and use the carbon to form complex hydrocarbons for food. The original atmosphere of the Earth was probably composed mostly of carbon dioxide, very similar to that of its rocky volcanic neighbors, Mars and Venus. The atmosphere of Mars today is 95 percent carbon dioxide, and the atmosphere of Venus is 97 percent carbon dioxide, whereas the atmosphere of the preindustrial-age Earth was only 0.03 percent carbon dioxide.²⁴ On Earth, life has been removing carbon dioxide from the atmosphere for the last 3,830 million years and has contributed to the transformation of the Earth’s atmosphere to its present carbon dioxide–depleted state.

TABLE 1.3  Major biotic and geologic events that occurred in the Archaean and Proterozoic Eons.

Abbreviations : Ma BP = millions of years before the present.

About 2,500 million years ago, the oxygen-producing activity of the cyanobacteria had its first major impact on the atmosphere of the Earth: the GOE, or great oxygenation event (table 1.3), marked the end of the Archaean Eon and the beginning of the Proterozoic (table 1.1). For the future evolution of complex—and large—life forms with aerobic metabolism, the GOE was good news indeed. For the ancient anaerobic life forms—the original inhabitants of Earth—the GOE was a disaster. The first mass extinction in the history of life on Earth probably occurred 2,500 million years ago, when vast unknown numbers of species of anaerobic bacteria and archaea perished by oxygen poisoning.

Diffusion of oxygen into the atmosphere was most probably the trigger of the first Snowball-Earth glaciation seen in the Proterozoic, which occurred about 200 million years after the GOE.²⁵ Snowball-Earth glaciations were much more extensive than any seen in the Phanerozoic, in that ice extended from the poles of the planet all the way down to the equator. From space, the entire planet may have looked like one giant snowball, hence the name Snowball-Earth. The steady downdraw of carbon dioxide, due to the photosynthetic activity of life, was already reducing the greenhouse effect of this gas in the atmosphere. The presence of free oxygen in the atmosphere may then have begun to remove an even more powerful greenhouse gas—methane—via oxidation.²⁶ The resulting sharp drop in the greenhouse capacity of the atmosphere of the Earth may have triggered the first great ice age in Earth’s history.

Some 300 million years later, oxygen concentrations in the atmosphere reached high enough levels to begin oxidizing iron on land, and the first redbed strata appeared in the terrestrial rock record (table 1.3). Redbeds are layered beds of red sandstones and shales in which the iron has been oxidized, giving the strata a characteristic rusty-red color. By this time, the aerobic photosynthetic activity of cyanobacteria had raised the amount of oxygen in the atmosphere to something between 1 percent and 4 percent of the atmosphere.²⁷

The Earliest Invaders of Land

Enjoying the preview?
Page 1 of 1