Comprehensive Physiology Wiley Online Library

Evolutionary Aspects of Atmospheric Oxygen and Organisms

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 The Origin of Oxygen in the Universe
2 Oxygen Compounds in the Universe
3 Early History of the Earth
4 Production of the First Living Organisms
5 Energetics of the First Living Organisms
6 When did the Earth'S Biosphere Originate?
7 The Transition Stage Between a Reducing and an Oxidizing Atmosphere
8 Development of the Oxidizing Atmosphere
9 Chronology of Events in the Oxygen Atmosphere
10 Oxygen as a Biological Energy Store
11 Oxygen Compounds in the Atmosphere
12 Stability of the Oxygen Cycle
13 Dual Effects of Oxygen on the Biosphere
14 The Future Atmosphere on Earth
15 Summary and Conclusion
Figure 1. Figure 1.

Nuclear binding energies. These values were calculated from the following equation 314:where EBA = binding energy per nucleon; Z = number of protons; N = number of neutrons; Me = mass excess in MeV; H = mass excess in MeV for the proton; and n = mass excess in MeV for the neutron. The mass excess equals the atomic mass minus the integer atomic mass unit. The atomic mass of 12C is defined to equal 12 integer atomic mass units. H = 8.0714 MeV and n = 7.2890 MeV. For 16O, the atomic mass equals 15.99491466 and the atomic mass number equals 16. Thus the mass excess equals ‐0.00508534. The Einstein equation (energy equals mass times the square of the speed of light) enables one to calculate the equivalent energy for this mass excess. One MeV is the equivalent energy of 0.001073535 atomic mass units. Dividing the ‐0.00508534 mass units by 0.001073535 mass units/MeV gives ‐4.737 MeV for the mass excess of 16O 202. For 16O, Z = N = 8, so by substitution into this equation, the binding energy per nucleon equals 7.9762 MeV.

Values for the mass excesses were taken from McGervey 183. Only one isotope for each element was chosen. These isotopes were generally either the most common or the longest‐lived ones for the element. Note that about 99% of naturally occurring hydrogen, helium, carbon, nitrogen, and oxygen is in the form of the isotopes 1H, 4He, 12C, 14N, and 16O, respectively. For almost all the elements, it was not significant what isotope was used for representing the element. Note that the atomic number equals the number of protons in the nucleus and that the number of nucleons equals the number of protons plus neutrons. The dotted lines represent peaks in nuclear stability. The first peak on the left is the 4He peak, which points out that 4He is much more stable than 1H. This fact accounts for the energy released in reaction 1 (in the text). Similarly, note that 12C and 16O are also peaks, which accounts for the energy released in reactions 2 and 3. Note that 56Fe is the most stable nucleus of all the atoms; this is the reason why the big stars end up with an iron core. Fusion of atoms lighter than iron can result in a release of energy, whereas fission of atoms heavier than iron can also result in a release of energy. According to the shell theory of the nucleus, Mayer 182 has shown that nuclei that contain 2, 8, 20, 28, 50, 82, or 126 protons or neutrons are unusually stable. These nuclear shells are analogous to the atomic valence shells of electrons. Mayer's chart indicates that the only nuclei to contain double magic numbers are 4He, 16O, 40Ca, 48Ca, and 208Pb. Of these five nuclei, only 4He, 16O, and 40Ca have the same number of protons and neutrons. This figure illustrates the stability of 4He and 16O.

Figure 2. Figure 2.

Evolution of mantle. In Greenberg's model 111 of the evolution of a typical particle in a dust cloud, there is a change in the chemical composition over the comparatively short period of time lasting for about 1 · 106 years, or 1 Myr. Note that initial conditions of water, methane, and ammonia are reducing in nature. As the evolution of the mantle progresses, there is a change toward an oxidized mantle, consisting of carbon dioxide, carbon monoxide, and molecular nitrogen. Thus these dust particles become a pinpoint of oxidation in a reducing interstellar dust cloud.

Figure 3. Figure 3.

Carbon isotopic values. (See Table 7 for explanation.) The reference used for measuring δ13C was the Peedee belemnite standard (PBS). The δ13C is a measure of the distribution of heavy carbon (13C) as compared to light carbon (12C) in a sample. Note that inorganic carbonate carbon has relatively more heavy carbon than organic carbon. This is due to the kinetic fixation of carbon by the biosphere. Most of the carbon fixation proceeds by the ribulose 1,5‐bisphosphate carboxylase (RuBP, RUBISCO) pathway to phosphoglycerate (PGA), a three‐carbon compound as in C3 plants. Of lesser importance is the phosphoenolpyruvate (PEP) pathway to produce oxaloacetate (OAA), a four‐carbon compound, as in C4 plants. Carbon fixed by the RUBISCO pathway has δ13C values between ‐20‰ and ‐30‰, whereas carbon fixed by the PEP pathway has δ13C values between ‐2‰ and ‐3‰ 248. Note that the geological samples are obtained at different depths (stratigraphic position), corresponding to the different ages when the samples were deposited. The ages are shown on top and they correspond to the Late Proterozoic, which lasted from about 550 Myr to 900 Myr. Individual data points are not shown, but they clearly showed dips in the δ13C for both carbonate and organic carbons. Dips are due to an increase of biological activity. If more carbon dioxide is involved in photosynthesis, then the organic carbon would be composed of more light carbon. If photosynthesis is increased, then the heavy carbon would be diluted by biological respiration in both the atmosphere and hydrosphere. This would explain the dips in both records. One consequence of this activity would be an increase in the oxygen concentrations in both the atmosphere and biosphere. Now, as a result of the increased oxygen, there would be an increase of oxygen toxicity. The increased oxygen would produce an evolutionary pressure to combat this poison by producing more natural antioxidant defenses. This would lead to the cyclic characteristics seen in both records. The figure was adapted from Holland 124.

Figure 4. Figure 4.

Stability of some carbon compounds in the presence of hydrogen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = 2a(n ‐ w).

The more thermodynamically stable a chemical is in the presence of hydrogen, the lower it is on this diagram. The free energies for Figures 4,5,7, and 8 were obtained from Latimer 164. The standard free energy for 1/6C6H12O6 is 5.8 kJ/mol, so the free energy for the reaction of 1/6C6H12O6 plus H2O yielding CO2 plus 2H2 is ‐5.8 kJ/mol. Similarly, the standard free energy for CH4 is ‐129.8 kJ/mol, so the free energy for the reaction of CO2 plus 4H2 yielding CH4 plus 2H2O is ‐129.8 kJ/mol. Summing these two reactions gives a value of ‐135.6 kJ/mol for the net reaction of 1/6C6H12O6 plus 2H2 yielding CH4 plus H2O.

Figure 5. Figure 5.

Stability of some nitrogen compounds in the presence of hydrogen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = 2(w ‐ n)/a.

For an explanation of the type of diagram, see Figure 4.

Figure 6. Figure 6.

Life‐energy profile in different atmospheres. A living cell is characterized by possessing more energy than its external environment. What keeps this energy difference? The cell membrane represented by A provides a barrier to the flow of energy out of the cell by means of active transport mechanisms. Gases, such as hydrogen and oxygen, represented by X2, can cross the membrane barrier by diffusion. X2 is taken up by the cell to produce biological energy, represented in this diagram by the X2 SINK. Some of the X2 in the X2 SINK can be used for anabolic synthesis as well. However, some of the X2 can destroy the cellular constituents. As a consequence, anti‐X2 defenses, represented by B, had to be developed by the cell to survive. Thus the first primitive cells had to deal with hydrogen as an energy source and hydrogen toxicity, whereas modern aerobic cells have to deal with oxygen as an energy source and oxygen toxicity.

Figure 7. Figure 7.

Stability of some carbon compounds in the presence of oxygen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = a(w + 2 ‐ 2n).

For an explanation of the type of diagram, see Figure 4.

Figure 8. Figure 8.

Stability of some nitrogen compounds in the presence of oxygen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = (2n + w)/a.

For an explanation of the type of diagram, see Figure 4.

Figure 9. Figure 9.

O2 exchanges in the atmosphere. Most of the values are derived from Gilbert 102. The value for the CO2 reservoir was calculated from Hall 117. For simplicity, we have rounded off the rates determined by Keeling and Shertz for the anthropogenic factors as well as the rate of CO2 entering the ocean 147.

The three most abundant oxygen species residing in the atmosphere are molecular oxygen or dioxygen, water vapor, and carbon dioxide. Other important oxygen species are ozone, carbon monoxide, hydrogen peroxide, nitrous oxide (N2O), sulfur oxides, the free radicals nitric oxide (NO) and nitrogen dioxide (NO2), the hydroperoxyl radical (HO2), and the hydroxyl radical (OH). Note that the bulk of the oxygen in the atmosphere is in the form of molecular oxygen or dioxygen. Photosynthetic production of oxygen is the predominant fast process, and the corresponding respiratory uptake rate of oxygen occurs at almost the same rate. Plants that produce oxygen also use oxygen. The plant respiration rate is 8,500 Emol/Myr, so the net rate of photosynthesis is 15,000 Emol/Myr minus 8,500 Emol/Myr, or 6,500 Emol/Myr. Thus the 37 Emol of dioxygen in the atmosphere is replaced by photosynthesis every 5,700 years.

Weathering and erosion processes expose reduced chemicals to the atmosphere and allow these chemicals to be oxidized by atmospheric oxygen. The release of reduced chemicals from volcanoes and midocean ridges due to tectonic activity also has the same effect as weathering on the atmospheric oxygen reservoir; therefore, this activity has been considered as part of weathering 17,300. However, decay and burial of organic chemicals removes carbon dioxide and increases oxygen 17. Mountain formation can bury organic carbon and lead to a release of oxygen 65. Planting trees also has the same effect as burial processes, while deforestation has the opposite effect. It has been suggested by many authors that planting trees can decrease the buildup of carbon dioxide produced by the burning of fossil fuels 69,132,201,256. Deforestation is occurring today at such a high rate that tree planting barely offsets its effect. Although these processes are of considerable importance to the small carbon dioxide reservoir, they are of little significance to the much larger oxygen reservoir.

Other gases, such as methane and carbon monoxide, also are oxidized by dioxygen. When carbohydrates are burned, one mole of oxygen is consumed for each mole of carbon dioxide released (see stage 6 in Table 9. Physiologists have used the respiratory exchange ratio or respiratory quotient (R) as the symbol for the ratio of carbon dioxide released to oxygen consumed by organisms 276,308. Similarly, we can use the atmospheric exchange of these two gases as also being represented by the symbol R. Thus when carbohydrates are burned, the value of the atmospheric exchange ratio, or atmospheric quotient R, is one. However, when hydrocarbons are burned, the atmospheric R is less than one, as illustrated in the following reaction:

In this reaction, the atmospheric R is 0.5. This results in more water being formed. The burning of fossil fuels reduces the atmospheric dioxygen at the rate of 670 Emol/Myr; 470 Emol/Myr of carbon dioxide are produced from this burning and 200 Emol/Myr of water are also produced. The atmospheric R due to anthropogenic sources is equal to 470 Emol/Myr divided by 670 Emol/Myr, or 0.7. About 250 Emol/Myr of carbon dioxide enter the ocean so that the net rate of increase in the atmospheric carbon dioxide reservoir is about 220 Emol/Myr.

What will be the final outcome of these anthropogenic factors on these atmospheric reservoirs? The carbon in the Earth's fossil‐fuel reservoir is about 0.9 Emol 102. If we assume that the burning of carbon from this reservoir is the principal anthropogenic factor leading to the increase of atmospheric carbon dixoide and that the rate of this increase is kept constant, then we can answer this question. The fossil‐fuel reservoir will be depleted in roughly 2,000 years (0.9 Emol divided by 0.47 Emol/thousand years). The increase of the atmospheric carbon dioxide reservoir will then equal 2,000 years times 0.22 Emol/thousand years, or 0.44 Emol; this reservoir will then equal 0.0619 Emol plus 0.44 Emol, or roughly 0.5 Emol. Dividing the 0.5 Emol by 0.0619 Emol gives a value of 8.1; this means that the present carbon dioxide reservoir will have increased by 810%! The corresponding decrease of the atmospheric oxygen will then equal 2,000 years times 0.67 Emol/thousand years, or 1.34 Emol; this reservoir will then equal 37 Emol minus 1.34 Emol, or roughly 35.7 Emol. Dividing the 35.7 Emol by 37 Emol gives a value of 0.965; this means that the present oxygen reservoir will have decreased by only 3.5%. Hence, the atmospheric carbon dioxide is easier to change than the relatively stable oxygen reservoir.

Walker 298 assumed that atmospheric oxygen was in a steady state, but he emphasized that this was just an assumption. Kump 161 has pointed out that due to the lack of precise knowledge about cycling rates it is not possible to state if the atmospheric dioxygen reservoir is decreasing, increasing, or remaining the same.

Figure 10. Figure 10.

Slow processes in the oxygen cycle. Photosynthesis is the predominant biological process for producing oxygen; this process is reversed by respiration (see Fig. 9. The mantle releases gases in volcanoes and solid materials in midocean rifts caused by plate tectonics. Iron is released in the ferrous state in these rifts. X represents reduced chemicals such as iron 300, sulfur 17,299,301, and other gases, such as methane and carbon monoxide 300. Reduction of the oxidized X (XO2) is accomplished by biological activity. Hydrogen is released from volcanoes and is removed from the Earth by photodissociation of water in the upper atmosphere at about equal rates 300.



Figure 1.

Nuclear binding energies. These values were calculated from the following equation 314:where EBA = binding energy per nucleon; Z = number of protons; N = number of neutrons; Me = mass excess in MeV; H = mass excess in MeV for the proton; and n = mass excess in MeV for the neutron. The mass excess equals the atomic mass minus the integer atomic mass unit. The atomic mass of 12C is defined to equal 12 integer atomic mass units. H = 8.0714 MeV and n = 7.2890 MeV. For 16O, the atomic mass equals 15.99491466 and the atomic mass number equals 16. Thus the mass excess equals ‐0.00508534. The Einstein equation (energy equals mass times the square of the speed of light) enables one to calculate the equivalent energy for this mass excess. One MeV is the equivalent energy of 0.001073535 atomic mass units. Dividing the ‐0.00508534 mass units by 0.001073535 mass units/MeV gives ‐4.737 MeV for the mass excess of 16O 202. For 16O, Z = N = 8, so by substitution into this equation, the binding energy per nucleon equals 7.9762 MeV.

Values for the mass excesses were taken from McGervey 183. Only one isotope for each element was chosen. These isotopes were generally either the most common or the longest‐lived ones for the element. Note that about 99% of naturally occurring hydrogen, helium, carbon, nitrogen, and oxygen is in the form of the isotopes 1H, 4He, 12C, 14N, and 16O, respectively. For almost all the elements, it was not significant what isotope was used for representing the element. Note that the atomic number equals the number of protons in the nucleus and that the number of nucleons equals the number of protons plus neutrons. The dotted lines represent peaks in nuclear stability. The first peak on the left is the 4He peak, which points out that 4He is much more stable than 1H. This fact accounts for the energy released in reaction 1 (in the text). Similarly, note that 12C and 16O are also peaks, which accounts for the energy released in reactions 2 and 3. Note that 56Fe is the most stable nucleus of all the atoms; this is the reason why the big stars end up with an iron core. Fusion of atoms lighter than iron can result in a release of energy, whereas fission of atoms heavier than iron can also result in a release of energy. According to the shell theory of the nucleus, Mayer 182 has shown that nuclei that contain 2, 8, 20, 28, 50, 82, or 126 protons or neutrons are unusually stable. These nuclear shells are analogous to the atomic valence shells of electrons. Mayer's chart indicates that the only nuclei to contain double magic numbers are 4He, 16O, 40Ca, 48Ca, and 208Pb. Of these five nuclei, only 4He, 16O, and 40Ca have the same number of protons and neutrons. This figure illustrates the stability of 4He and 16O.



Figure 2.

Evolution of mantle. In Greenberg's model 111 of the evolution of a typical particle in a dust cloud, there is a change in the chemical composition over the comparatively short period of time lasting for about 1 · 106 years, or 1 Myr. Note that initial conditions of water, methane, and ammonia are reducing in nature. As the evolution of the mantle progresses, there is a change toward an oxidized mantle, consisting of carbon dioxide, carbon monoxide, and molecular nitrogen. Thus these dust particles become a pinpoint of oxidation in a reducing interstellar dust cloud.



Figure 3.

Carbon isotopic values. (See Table 7 for explanation.) The reference used for measuring δ13C was the Peedee belemnite standard (PBS). The δ13C is a measure of the distribution of heavy carbon (13C) as compared to light carbon (12C) in a sample. Note that inorganic carbonate carbon has relatively more heavy carbon than organic carbon. This is due to the kinetic fixation of carbon by the biosphere. Most of the carbon fixation proceeds by the ribulose 1,5‐bisphosphate carboxylase (RuBP, RUBISCO) pathway to phosphoglycerate (PGA), a three‐carbon compound as in C3 plants. Of lesser importance is the phosphoenolpyruvate (PEP) pathway to produce oxaloacetate (OAA), a four‐carbon compound, as in C4 plants. Carbon fixed by the RUBISCO pathway has δ13C values between ‐20‰ and ‐30‰, whereas carbon fixed by the PEP pathway has δ13C values between ‐2‰ and ‐3‰ 248. Note that the geological samples are obtained at different depths (stratigraphic position), corresponding to the different ages when the samples were deposited. The ages are shown on top and they correspond to the Late Proterozoic, which lasted from about 550 Myr to 900 Myr. Individual data points are not shown, but they clearly showed dips in the δ13C for both carbonate and organic carbons. Dips are due to an increase of biological activity. If more carbon dioxide is involved in photosynthesis, then the organic carbon would be composed of more light carbon. If photosynthesis is increased, then the heavy carbon would be diluted by biological respiration in both the atmosphere and hydrosphere. This would explain the dips in both records. One consequence of this activity would be an increase in the oxygen concentrations in both the atmosphere and biosphere. Now, as a result of the increased oxygen, there would be an increase of oxygen toxicity. The increased oxygen would produce an evolutionary pressure to combat this poison by producing more natural antioxidant defenses. This would lead to the cyclic characteristics seen in both records. The figure was adapted from Holland 124.



Figure 4.

Stability of some carbon compounds in the presence of hydrogen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = 2a(n ‐ w).

The more thermodynamically stable a chemical is in the presence of hydrogen, the lower it is on this diagram. The free energies for Figures 4,5,7, and 8 were obtained from Latimer 164. The standard free energy for 1/6C6H12O6 is 5.8 kJ/mol, so the free energy for the reaction of 1/6C6H12O6 plus H2O yielding CO2 plus 2H2 is ‐5.8 kJ/mol. Similarly, the standard free energy for CH4 is ‐129.8 kJ/mol, so the free energy for the reaction of CO2 plus 4H2 yielding CH4 plus 2H2O is ‐129.8 kJ/mol. Summing these two reactions gives a value of ‐135.6 kJ/mol for the net reaction of 1/6C6H12O6 plus 2H2 yielding CH4 plus H2O.



Figure 5.

Stability of some nitrogen compounds in the presence of hydrogen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = 2(w ‐ n)/a.

For an explanation of the type of diagram, see Figure 4.



Figure 6.

Life‐energy profile in different atmospheres. A living cell is characterized by possessing more energy than its external environment. What keeps this energy difference? The cell membrane represented by A provides a barrier to the flow of energy out of the cell by means of active transport mechanisms. Gases, such as hydrogen and oxygen, represented by X2, can cross the membrane barrier by diffusion. X2 is taken up by the cell to produce biological energy, represented in this diagram by the X2 SINK. Some of the X2 in the X2 SINK can be used for anabolic synthesis as well. However, some of the X2 can destroy the cellular constituents. As a consequence, anti‐X2 defenses, represented by B, had to be developed by the cell to survive. Thus the first primitive cells had to deal with hydrogen as an energy source and hydrogen toxicity, whereas modern aerobic cells have to deal with oxygen as an energy source and oxygen toxicity.



Figure 7.

Stability of some carbon compounds in the presence of oxygen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = a(w + 2 ‐ 2n).

For an explanation of the type of diagram, see Figure 4.



Figure 8.

Stability of some nitrogen compounds in the presence of oxygen. The ordinate represents the decrease in free energy for the various compounds, according to the following equation 96:where b = (2n + w)/a.

For an explanation of the type of diagram, see Figure 4.



Figure 9.

O2 exchanges in the atmosphere. Most of the values are derived from Gilbert 102. The value for the CO2 reservoir was calculated from Hall 117. For simplicity, we have rounded off the rates determined by Keeling and Shertz for the anthropogenic factors as well as the rate of CO2 entering the ocean 147.

The three most abundant oxygen species residing in the atmosphere are molecular oxygen or dioxygen, water vapor, and carbon dioxide. Other important oxygen species are ozone, carbon monoxide, hydrogen peroxide, nitrous oxide (N2O), sulfur oxides, the free radicals nitric oxide (NO) and nitrogen dioxide (NO2), the hydroperoxyl radical (HO2), and the hydroxyl radical (OH). Note that the bulk of the oxygen in the atmosphere is in the form of molecular oxygen or dioxygen. Photosynthetic production of oxygen is the predominant fast process, and the corresponding respiratory uptake rate of oxygen occurs at almost the same rate. Plants that produce oxygen also use oxygen. The plant respiration rate is 8,500 Emol/Myr, so the net rate of photosynthesis is 15,000 Emol/Myr minus 8,500 Emol/Myr, or 6,500 Emol/Myr. Thus the 37 Emol of dioxygen in the atmosphere is replaced by photosynthesis every 5,700 years.

Weathering and erosion processes expose reduced chemicals to the atmosphere and allow these chemicals to be oxidized by atmospheric oxygen. The release of reduced chemicals from volcanoes and midocean ridges due to tectonic activity also has the same effect as weathering on the atmospheric oxygen reservoir; therefore, this activity has been considered as part of weathering 17,300. However, decay and burial of organic chemicals removes carbon dioxide and increases oxygen 17. Mountain formation can bury organic carbon and lead to a release of oxygen 65. Planting trees also has the same effect as burial processes, while deforestation has the opposite effect. It has been suggested by many authors that planting trees can decrease the buildup of carbon dioxide produced by the burning of fossil fuels 69,132,201,256. Deforestation is occurring today at such a high rate that tree planting barely offsets its effect. Although these processes are of considerable importance to the small carbon dioxide reservoir, they are of little significance to the much larger oxygen reservoir.

Other gases, such as methane and carbon monoxide, also are oxidized by dioxygen. When carbohydrates are burned, one mole of oxygen is consumed for each mole of carbon dioxide released (see stage 6 in Table 9. Physiologists have used the respiratory exchange ratio or respiratory quotient (R) as the symbol for the ratio of carbon dioxide released to oxygen consumed by organisms 276,308. Similarly, we can use the atmospheric exchange of these two gases as also being represented by the symbol R. Thus when carbohydrates are burned, the value of the atmospheric exchange ratio, or atmospheric quotient R, is one. However, when hydrocarbons are burned, the atmospheric R is less than one, as illustrated in the following reaction:

In this reaction, the atmospheric R is 0.5. This results in more water being formed. The burning of fossil fuels reduces the atmospheric dioxygen at the rate of 670 Emol/Myr; 470 Emol/Myr of carbon dioxide are produced from this burning and 200 Emol/Myr of water are also produced. The atmospheric R due to anthropogenic sources is equal to 470 Emol/Myr divided by 670 Emol/Myr, or 0.7. About 250 Emol/Myr of carbon dioxide enter the ocean so that the net rate of increase in the atmospheric carbon dioxide reservoir is about 220 Emol/Myr.

What will be the final outcome of these anthropogenic factors on these atmospheric reservoirs? The carbon in the Earth's fossil‐fuel reservoir is about 0.9 Emol 102. If we assume that the burning of carbon from this reservoir is the principal anthropogenic factor leading to the increase of atmospheric carbon dixoide and that the rate of this increase is kept constant, then we can answer this question. The fossil‐fuel reservoir will be depleted in roughly 2,000 years (0.9 Emol divided by 0.47 Emol/thousand years). The increase of the atmospheric carbon dioxide reservoir will then equal 2,000 years times 0.22 Emol/thousand years, or 0.44 Emol; this reservoir will then equal 0.0619 Emol plus 0.44 Emol, or roughly 0.5 Emol. Dividing the 0.5 Emol by 0.0619 Emol gives a value of 8.1; this means that the present carbon dioxide reservoir will have increased by 810%! The corresponding decrease of the atmospheric oxygen will then equal 2,000 years times 0.67 Emol/thousand years, or 1.34 Emol; this reservoir will then equal 37 Emol minus 1.34 Emol, or roughly 35.7 Emol. Dividing the 35.7 Emol by 37 Emol gives a value of 0.965; this means that the present oxygen reservoir will have decreased by only 3.5%. Hence, the atmospheric carbon dioxide is easier to change than the relatively stable oxygen reservoir.

Walker 298 assumed that atmospheric oxygen was in a steady state, but he emphasized that this was just an assumption. Kump 161 has pointed out that due to the lack of precise knowledge about cycling rates it is not possible to state if the atmospheric dioxygen reservoir is decreasing, increasing, or remaining the same.



Figure 10.

Slow processes in the oxygen cycle. Photosynthesis is the predominant biological process for producing oxygen; this process is reversed by respiration (see Fig. 9. The mantle releases gases in volcanoes and solid materials in midocean rifts caused by plate tectonics. Iron is released in the ferrous state in these rifts. X represents reduced chemicals such as iron 300, sulfur 17,299,301, and other gases, such as methane and carbon monoxide 300. Reduction of the oxidized X (XO2) is accomplished by biological activity. Hydrogen is released from volcanoes and is removed from the Earth by photodissociation of water in the upper atmosphere at about equal rates 300.

References
 1. Ahrens, T. J. Impact erosion of terrestrial planetary atmospheres. Annu. Rev. Earth Planet. Sci. 21: 525–555, 1993.
 2. Ahrens, T. J., and A. W. Harris. Deflection and fragmentation of near‐Earth asteroids. Nature 360: 429–433, 1992.
 3. Aihara, J. Why aromatic compounds are stable. Sci. Am. 266 (No. 3): 62–68, 1992.
 4. Ames, B. N., M. K. Shigenaga, and T. M. Hagen. Oxidants, antioxidants, and the degenerative diseases of aging. Proc. Natl. Acad. Sci. USA 90: 7915–7922, 1993.
 5. Anders, E., and M. Ebihara. Solar‐system abundances of the elements. Geochim. Cosmochim. Acta 46: 2363–2380, 1982.
 6. Anderson, M. J., and E. J. Stanbridge. Tumor suppressor genes studied by cell hybridization chromosome transfer. FASEB J. 7: 826–833, 1993.
 7. Aneshansley, D. J., T. Eisner, J. M. Widom, and B. Widom. Biochemistry at 100°C: explosive secretory discharge of bombardier beetles (Brachinus). Science 165: 61–63, 1969.
 8. Aono, S., F. O. Bryant, and M. W. Adams. A novel and remarkably thermostable ferredoxin from the hyperthermophilic archaebacterium Pyrococcus furiosus. J. Bacteriol. 171: 3433–3439, 1989.
 9. Archibald, F. S. Lactobacillus plantarum, an organism not requiring iron. FEMS Microbiol. Lett. 19: 29–32, 1983.
 10. Babbs, C. F. Free radicals and the etiology of colon cancer. Free Radic. Biol. Med. 8: 191–200, 1990.
 11. Babior, B. M., R. S. Kipnes, and J. T. Curnette. Biological defense mechanisms. The production by leukocytes of superoxide, a potential bactericidal agent. J. Clin. Invest. 52: 741–744, 1973.
 12. Bartlett, J. G., A. Blanchard, J. Silk, and M. S. Turner. The case for a Hubble constant of 30 km·s‐1·Mpc‐1. Science 267: 980–983, 1995.
 13. Beaman, L., and B. L. Beaman. The role of oxygen and its derivatives in microbial pathogenesis and host defense. Annu. Rev. Microbiol. 38: 27–48, 1984.
 14. Beaman, L., and B. L. Beaman. Monoclonal antibodies demonstrate that superoxide dismutase contributes to protection of Nocardia asteroides within the intact host. Infect. Immun. 58: 3122–3128, 1990.
 15. Becker, B. F. Towards the physiological function of uric acid. Free Radic. Biol. Med. 14: 615–631, 1993.
 16. Berkner, L. V., and L. C. Marshall. The history of growth of oxygen in the earth's atmosphere. In: The Origin and Evolution of Atmospheres and Oceans, edited by P. J. Brancazio and A. G. W. Cameron. New York: Wiley, 1964 p. 102–126.
 17. Bernal, J. D. The physical basis of life. Proc. Phys. Soc. 62 (A): 537–558, 1949.
 18. Berner, R. A., and D. E. Canfield. A new model for atmospheric oxygen over Phanerozoic time. Am. J. Sci. 289: 333–361, 1989.
 19. Berry, J. A. Biosphere, atmosphere, ocean interactions: a plant physiologist's perspective. In: Primary Productivity and Biogeochemical Cycles in the Sea. Brookhaven Symposium in Biology No. 37, edited by P. G. Falkowski and A. B. Woodhead. New York: Plenum, 1992, p. 441–454.
 20. Beyer, W. F., Jr., and I. Fridovich. Pseudocatalase from Lactobacillus plantarum: evidence for a homopentameric structure containing two atoms of manganese per subunit. Biochemistry 24: 6460–6467, 1985.
 21. Bice, D. M., C. R. Newton, S. McCauley, P. W. Reiners, and C. A. McRoberts. Shocked quartz at the Triassic‐Jurassic boundary in Italy. Science 255: 443–446, 1992.
 22. Bielski, B. H. J., D. E. Cabelli, R. L. Arudi, and A. B. Ross. Reactivity of HO2/O2‐ radicals in aqueous solution. J. Phys. Chem. Ref. Data 14: 1041–1100, 1985.
 23. Bishop, N. L., and L. W. Jones. Alternate fates of the photochemical reducing power generated in photosynthesis: hydrogen production and nitrogen fixation. Curr. Topics Bioenerg. 8: 3–31, 1978.
 24. Biswell, H. H. Prescribed Burning in California Wildlands Vegetation Management. Berkeley: Univ. California Press, 1989.
 25. Blaut, M., V. Müller, and G. Gottschalk. Energetics of methanogens. In: The Bacteria. Volume 12. Bacterial Energetics, edited by T. A. Krulwich. New York: Academic, 1990, p. 505–537.
 26. Blum, J. D., C. P. Chamberlain, M. P. Hingston, C. Koeberl, L. E. Marin, B. C. Schuraytz, and V. L. Sharpton. Isotopic comparison of K/T boundary impact glass with melt rock from the Chicxulub and Manson impact structures. Nature 364: 325–327, 1993.
 27. Bonner, W. A. Terrestrial and extraterrestrial sources of molecular homochirality. Orig. Life Evol. Biosph. 21: 407–420, 1992.
 28. Bonner, W. A. Chirality and life. Orig. Life Evol. Biosph. 25: 175–190, 1995.
 29. Bors, W., W. Heller, C. Michel, and M. Saran. Flavonoids as antioxidants: determination of radical‐scavenging efficiencies. In: Methods in Enzymology. Oxygen Radicals in Biological Systems. Oxygen Radicals and Antioxidants, edited by L. Packer, and A. N. Glazer. San Diego, CA: Academic, 1990, vol. 186, pt. B, p. 343–355.
 30. Boveris, A., and E. Cadenas. Production of superoxide radicals and hydrogen peroxide in mitochondria. In: Superoxide Dismutase, edited by L. W. Oberley. Boca Raton, FL: CRC, 1982, vol. II, p. 15–30.
 31. Bowring, S. A., J. P. Grotzinger, C. E. Isachsen, A. H. Knoll, S. M. Pelechaty, and P. Kolosov. Calibrating rates of early Cambrian evolution. Science 261: 1293–1299, 1993.
 32. Braun, C., and W. G. Zumft. Marker exchange of the structural genes for nitric oxide reductase blocks the denitrification pathway of Pseudomonas stutzeri at nitric oxide. J. Biol. Chem. 266: 22785–22788, 1991.
 33. Bryant, F. O., and M. W. Adams. Characterization of hydrogenase from the hyperthermophic archaebacterium, Pyrococcus furiosus. J. Biol. Chem. 264: 5070–5079, 1989.
 34. Budyko, M. I., A. B. Ronov, and A. L. Yanshin, History of the Earth's Atmosphere, translated by S. F. Lemehko and V. G. Yanuta. New York: Springer‐Verlag, 1987.
 35. Burton, G. W., and K. U. Ingold. β‐Carotene: an unusual type of lipid antioxidant. Science 224: 569–573, 1984.
 36. Byers, V. S., and R. W. Baldwin (Eds). Immunology of Malignant Diseases. Boston: MTP, 1987.
 37. Cairns‐Smith, A. G., and H. Hartman (Eds). Clay Minerals and the Origin of Life. New York: Cambridge Univ. Press, 1986.
 38. Caldeira, M., and J. F. Kasting. The life span of the biosphere revisited. Nature 360: 721–723, 1992.
 39. Calloway, D. H. (Ed). Human Ecology in Space Flight. New York: New York Academy of Sciences, 1966.
 40. Calvin, C. Chemical Evolution. Molecular Evolution Towards the Origin of Living Systems on the Earth and Elsewhere. New York: Oxford Univ. Press, 1969.
 41. Cameron, A. G. W., and W. Benz. The origin of the moon and the single impact hypothesis IV. Icarus 92: 204–216, 1991.
 42. Campbell, I. M. Energy and the Atmosphere. A Physical‐Chemical Approach (2nd ed). New York: Wiley, 1986.
 43. Canfield, L. M., J. W. Forage, and J. G. Valenzuela. Carotenoids as cellular antioxidants. Proc. Soc. Exp. Biol. Med. 200: 260–265, 1992.
 44. Cannon, W. B. The Way of an Investigator. A Scientist's Experience in Medical Research (1968 ed.). New York: Hafner, 1945, p. 108.
 45. Carroll, D. Role of clay minerals in the transportation of iron. Geochim. Cosmochim. Acta 14: 1–28, 1958.
 46. Chalmers, A. H. Simple, sensitive measurement of carbon monoxide in plasma. Clin. Chem. 37: 1442–1445, 1991.
 47. Chapman, D. J., and J. W. Schopf. Biological and biochemical effects of the development of an aerobic environment. In: Earth's Earliest Biosphere. Its Origin and Evolution, edited by J. W. Schopf. Princeton, NJ: Princeton Univ. Press, 1983, p. 302–320.
 48. Chevalier, R. A. Supernova 1987A at five years of age. Nature 355: 691–696, 1992.
 49. Chiueh, C. C., D. L. Gilbert, and C. A. Colton (Eds). The Neurobiology of NO and OH. Annals NY Acad. Sci. New York: New York Academy of Sciences, 1994, vol. 738.
 50. Chyba, C. A. Extraterrestrial amino acids and terrestrial life. Nature 348: 113–114, 1990.
 51. Chyba, C. F. Explosions of small Spacewatch objects in the Earth's atmosphere. Nature 363: 701–703, 1993.
 52. Chyba, C., and C. Sagan. Endogenous production, exogenous delivery and impact‐shock synthesis of organic molecules: an inventory for the origins of life. Nature 355: 125–132, 1992.
 53. Chyba, C. F., P. J. Thomas, and K. J. Zahnle. The 1908 Tunguska explosion: atmospheric disruption of a stony asteroid. Nature 361: 40–44, 1993.
 54. Clark, R. T., Jr., J. N. Stannard, and W. O. Fenn. The burning of CO to CO2 by isolated tissues as shown by the use of radioactive carbon. Am. J. Physiol. 161: 40–46, 1950.
 55. Cloud, P. Beginnings of a biospheric evolution and their biogeochemical consequences. Paleobiology 2: 351–357, 1976.
 56. Colton, C. A., and D. L. Gilbert. Microglia, an in vivo source of reactive oxygen species in the brain. In: Advances in Neurology. Neural Regeneration, edited by F. Seil. New York: Raven, 1993, vol. 59, p. 321–326.
 57. Crick, F. Forward. In: The RNA World. The Nature of Modern RNA Suggests a Prebiotic RNA World, edited by R. F. Gesteland and J. F. Atkins. Plainview, NY: Cold Spring Harbor Lab., 1993, p. xi–xiv.
 58. Crowley, T. J. Past CO2 changes and tropical sea surface temperatures. Paleoceanography 6: 387–394, 1991.
 59. Crowley, T. J., and G. R. North. Paleoclimatology. New York: Oxford Univ. Press, 1991.
 60. Cutler, R. G. Antioxidants and aging. Am. J. Clin. Nutr. 53 (suppl. 1): 373S–379S, 1991.
 61. Dalgarno, A. Interstellar chemistry. In: Chemistry in Space, edited by J. M. Greenberg and V. Pirronello. Boston: Kluwer Acad. Pub., 1991, p. 71–87.
 62. Damsté, J. S. S., S. G. Wakeham, M. E. L. Kohnen, J. M. Hayes, and J.W. De Leeuw. A 6,000‐year sedimentary molecular record of chemocline excursions in the Black Sea. Nature 362: 827–829, 1993.
 63. Davis, D. H., M. Doudoroff, R. Y. Stanier, and M. Mandel. Proposal to reject the genus Hydrogenomonas: taxonomic implications. Int. J. Syst. Bacteriol. 19: 375–390, 1969.
 64. Davis, J. C., and T. K. Hunt (Eds). Hyperbaric Oxygen Therapy. Bethesda, MD: Undersea Medical Society, 1977.
 65. Dean, J., D. J. Aneshansley, H. E. Edgerton, and T. Eisner. Defensive spray of the bombardier beetle: a biological pulse jet. Science 248: 1219–1221, 1990.
 66. De Baar, H. J. W., J. T. M. De Jong, D. C. E. Bakker, B. M. Löscher, C. Veth, U. Bathmann, and V. Semtacek. Importance of iron for planton blooms and carbon dioxide drawdown in the Southern ocean. Nature 373: 412–415, 1995.
 67. Demple, B. Regulation of bacterial oxidative stress genes. Annu. Rev. Genet. 25: 315–337, 1991.
 68. Des Marais, D. J., H. Strauss, R. E. Summons, and J. M. Hayes. Carbon isotope evidence for the stepwise oxidation of the Proterozoic environment. Nature 359: 605–609, 1992.
 69. Dickens, R. J., B. F. W. Croke, R. D. Cannon, and R. A. Bell. Evidence from stellar abundances for a large age difference between two globular clusters. Nature 351: 212–214, 1991.
 70. DuBois, A. B., T. Turaids, R. E. Mammen, and F. T. Nobrega. Pulmonary atelectasis in subjects breathing oxygen at sea level or at simulated altitude. J. Appl. Physiol. 21: 828–836, 1966.
 71. Dykens, J. A., and J. M. Shick. Oxygen production by endosymbiotic algae controls superoxide dismutase activity in their animal host. Nature 297: 579–580, 1982.
 72. Dyson, F. J. Can we control the carbon dioxide in the atmosphere? Energy 2: 287–291, 1977.
 73. Edmundson, W. T. Phosphorus, nitrogen, and algae in Lake Washington after diversion of sewage. Science 169: 690–691, 1970.
 74. Engel, M. H., S. A. Macko, and J. A. Silfer. Carbon isotope composition of individual amino acids in the Murchison meteorite. Nature 348: 47–49, 1990.
 75. Ensign, S. A., and P. W. Ludden. Characterization of the CO oxidation/H2 evolution system of Rhodospirillum rubrum. Role of a 22‐kDa iron‐sulfur protein in mediating electron transfer between carbon monoxide dehydrogenase and hydrogenase. J. Biol. Chem. 266: 18395–18403, 1991.
 76. Ernst, W. G. The early Earth and the Archean rock record. In: Earth's Earliest Biosphere. Its Origin and Evolution, edited by J. W. Schopf. Princeton, NJ: Princeton Univ. Press, 1983, p. 41–52.
 77. Fandrey, J., and H. F. Bunn. In vivo and in vitro regulation of erythropoietin mRNA: measurement by competitive polymerase chain reaction. Blood 81: 617–623, 1993.
 78. Faure, G. Principles of Isotope Geology (2nd ed.). New York: Wiley, 1986.
 79. Faust, B. C., C. Anastasio, J. M. Allen, and T. Arakaki. Aqueous‐phase photochemical formation of peroxides in authentic cloud and fog waters. Science 260: 73–75, 1993.
 80. Fay, P. Oxygen relations of nitrogen fixation in cyanobacteria. Microbiol. Rev. 56: 340–373, 1992.
 81. Feldman, P. D. The volatile composition of comets deduced from ultraviolet spectroscopy. In: Chemistry in Space, edited by J. M. Greenberg and V. Pirronello. Boston: Kluwer, 1991, p. 339–361.
 82. Feng, Q., T. K. Park, and J. Rebek. Crossover reactions between synthetic replicators yield active and inactive recombinants. Science 256: 1179–1180, 1992.
 83. Fenn, W. O., and D. M. Cobb. The burning of carbon monoxide by heart and skeletal muscle. Am. J. Physiol. 102: 393–401, 1932.
 84. Finlayson‐Pitts, B. J., and J. N. Pitts, Jr. Atmospheric Chemistry: Fundamentals and Experimental Techniques. New York: Wiley, 1986.
 85. Fletcher, M. The physiological activity of bacteria attached to solid surfaces. Adv. Microb. Physiol. 32: 53–85, 1991.
 86. Fornace, A. J., Jr. Mammalian genes induced by radiation; activation of genes associated with growth control. Annu. Rev Genet. 26: 507–526, 1992.
 87. Fox, S. W. Life from an orderly cosmos. Naturwissenschaften 67: 576–581, 1980.
 88. Fox, S. W., and K. Harada. Synthesis of uracil under conditions of a thermal model of prebiological chemistry. Science 133: 1923–1924, 1961.
 89. Frank, L. Oxygen toxicity in eukaryotes. In: Superoxide Dismutase. Pathological States, edited by L. W. Oberley. Boca Raton, FL: CRC, 1985, vol. III, p. 1–43.
 90. Fridovich, I. Oxygen toxicity in prokaryotes: the importance of superoxide dismutase. In: Superoxide Dismutase. edited by L. W. Oberley. Boca Raton, FL: CRC, 1982, vol. I, p. 79–88.
 91. Fridovich, I. Superoxide dismutases. An adaptation to a paramagnetic gas. J. Biol. Chem. 264: 7761–7764, 1989.
 92. Gajdusek, D. C. Transmissible and non‐transmissible amyloidoses: autocatalytic post‐translational conversion of host precursor proteins to β‐pleated configurations. J. Neuroimmunol. 20: 95–110, 1988.
 93. Gates, D. M. Climate Change and Its Biological Consequences. Sunderland, MA: Sinauer, 1993.
 94. Gerschman, R. Oxygen effects in biological systems. In: Symp. Spec. Lect. XXI Int. Cong. Physiol. Sci. Buenos Aires, 1959, p. 222–226.
 95. Gerschman, R., D. L. Gilbert, and D. Caccamise. Effect of various substances on survival times of mice exposed to different high oxygen tensions. Am. J. Physiol. 192: 563–571, 1958.
 96. Gerschman, R., D. L. Gilbert, and J. N. Frost. Sensitivity of Paramecium caudatum to high oxygen tensions and its modification by cobalt and manganese ions. Am. J. Physiol. 192: 572–576, 1958.
 97. Gerschman, R., D. L. Gilbert, S. W. Nye, P. Dwyer, and W. O. Fenn. Oxygen poisoning and x‐irradiation: a mechanism in common. Science 119: 623–626, 1954.
 98. Gilbert, D. L. Relationship between ion distribution and membrane potential during a steady state. Bull. Math. Biophys. 22: 323–349, 1960.
 99. Gilbert, D. L. Speculation on the relationship between organic and atmospheric evolution. Perspect. Biol. Med. 4: 58–71, 1960.
 100. Gilbert, D. L. The role of pro‐oxidants and antioxidants in oxygen toxicity. Radiat. Res. (suppl. 3): 44–53, 1963.
 101. Gilbert, D. L. Cosmic and geophysical aspects of the respiratory gases. In: Handbook of Physiology: Respiration, edited by W. Fenn and H. Rahn. Washington, DC: Am. Physiol. Soc., 1964, sect. 3, vol. I, p. 153–176.
 102. Gilbert, D. L. Atmosphere and oxygen. Physiologist 8: 9–34, 1965.
 103. Gilbert, D. L. Introduction: Oxygen and life. Anesthesiology 37: 100–111, 1972.
 104. Gilbert, D. L. Discussion: What controls atmospheric oxygen? BioSystems 12: 123–124, 1980.
 105. Gilbert, D. L. Significance of oxygen on earth. In: Oxygen and Living Processes: An Interdisciplinary Approach, edited by D. L. Gilbert. New York: Springer‐Verlag, 1981, p. 73–101.
 106. Gilbert, D. L. Oxygen: an overall biological view. In: Oxygen and Living Processes: An Interdisciplinary Approach, edited by D. L. Gilbert. New York: Springer‐Verlag, 1981, p. 376–392.
 107. Gilbert, D. L. From the breath of life to reactive oxygen species (ROS). In: Free Radicals—Mechanisms in Neurology and Psychiatry, edited by J. L. Cadet and J. E. Johnson. New York: Wiley, 1995 (in press).
 108. Gilbert, D. L., and G. Ehrenstein. Membrane surface charge. Curr. Top. Membr. Trans. 22: 407–421, 1984.
 109. Gilbert, D. L., R. Gerschman, J. Cohen, and W. Sherwood. The influence of high oxygen pressures on the viscosity of solutions of sodium desoxyribonucleic acid and of sodium alginate. J. Am. Chem. Soc. 79: 5677–5680, 1957.
 110. Gilliland, R. L. Solar evolution. Palaeogeogr. Palaeoclimatol. Palaeoecol. (Global Planet. Change Sect.) 75: 35–55, 1989.
 111. Gleason, J. F., P. K. Bhartia, J. R. Herman, R. McPeters, P. Newman, R. S. Stolarski, L. Flynn, G. Labow, D. Larko, C. Seftor, C. Wellemeyer, W. D. Komhyr, A. J. Miller, and W. Planet. Record low global ozone in 1992. Science 260: 523–526, 1993.
 112. Goldanskii, V. I., and V. V. Kuzmin. Chirality and cold origin of life. Nature 352: 114, 1991.
 113. Gorst, C. M., and S. W. Ragsdale. Characterization of the NiFeCO complex of carbon monoxide dehydrogenase as a cat‐alytically competent intermediate in the pathway of acetylcoenzyme A synthesis. J. Biol. Chem. 266: 20687–20693, 1991.
 114. Gould, S. J. Bully for Brontosaurus. Reflections in Natural History. New York: Norton, 1992, p. 339.
 115. Greenberg, J. M. Physical, chemical and optical interactions with interstellar dust. In: Chemistry in Space, edited by J. M. Greenberg and V. Pirronello. Boston: Kluwer, 1991, p. 227–261.
 116. Gregory, D. P. The hydrogen economy. Sci. Am. 228: 13–21, 1973.
 117. Grieve, R. A. F. When will enough be enough? Nature 363: 670–671, 1993.
 118. Gutteridge, J. M. C. Ageing and free radicals. Med. Lab. Sci. 49: 313–318, 1992.
 119. Guy, R. D., J. A. Berry, M. L. Fogel, and T. C. Hoering. Differential fractionation of oxygen isotopes by cyanide‐resistant and cyanide‐sensitive respiration in plants. Planta Med. 177: 483–491, 1989.
 120. Haldane, J. B. S. The origins of life. In: Adventures in Earth History. 1970, edited by P. Cloud. San Francisco: Freeman, 1954, p. 377–384.
 121. Hall, D. O. Carbon flows in the biosphere: present and future. J. Geol. Soc. Lond. 146: 175–181, 1989.
 122. Hall, D. T., D. F. Strobel, P. D. Feldman, M. A. McGrath, and H. A. Weaver. Detection of an oxygen atmosphere on Jupiter's moon Europa. Nature 373: 677–679, 1995.
 123. Halliwell, B., and J. M. C. Gutteridge. Free Radicals in Biology and Medicine (2nd ed.). New York: Oxford Univ. Press, 1989.
 124. Halliwell, J. J. Quantum cosmology and the creation of the universe. Sci. Am. 265: 76–79, 82–85, 1991.
 125. Harman, D. Aging: a theory based on free radical and radiation chemistry. J. Gerontol. 11: 298–299, 1956.
 126. Hayes, J. M. Geochemical evidence bearing on the origin of aerobiosis, a speculative hypothesis. In: Earth's Earliest Biosphere. Its Origin and Evolution, edited by J. W. Schopf. Princeton, NJ: Princeton Univ. Press, 1983, p. 291–301.
 127. Henderson, L. J. The Fitness of the Environment. Boston: Beacon Press, 1958.
 128. Hill, H. A. O. Iron: an element well‐fitted for its task? In: The Biological Chemistry of Iron, edited by H. B. Dunford, D. Dolphin, K. N. Raymond, and L. Sieker. Boston: Reidel, 1982, p. 3–12.
 129. Holland, H. Chemistry and evolution of the Proterozoic ocean. In: The Proterozoic Biosphere. A Multidisciplinary Study, edited by J. W. Schopf and C. Klein. New York: Cambridge Univ. Press, 1992, p. 169–172.
 130. Holland, H. Early Proterozoic atmospheric change. In: Early Life on Earth: Nobel Symposium 84, edited by S. Bengtson. New York: Columbia Univ. Press, 1994, p. 237–244.
 131. Holm, N. G. Why are hydrothermal systems proposed as plausible environments for the orign of life? Orig. Life Evol. Biosph. 22: 5–14, 1992.
 132. Holm, N. G., and R. J. Hennet. Hydrothermal systems: their varieties, dynamics, and suitability for prebiotic chemistry. Orig. Life Evol. Biosph. 22: 15–31, 1992.
 133. Hoban, S. Serendipitous images of methanol in comet Levy (1900 XX). Icarus 104: 149–151, 1993.
 134. Huebner, W. F., and D. C. Boice. Comets as a possible source of prebiotic molecules. Orig. Life Evol. Biosph. 21: 299–315, 1992.
 135. Hughes, M. N., and R. K. Poole. Metals and Micro‐organisms. New York: Chapman and Hall, 1989.
 136. Hunten, D. M. Atmospheric evolution of the terrestrial planets. Science 259: 915–920, 1993.
 137. Hunten, D. M. Europa's oxygen atmosphere. Nature 373: 654, 1995.
 138. Idso, S. B. The aerial fertilization effect of CO2 and its implications for global carbon cycling and maximum greenhouse warming. Bull. Am. Meteorol. Soc. 72: 962–965, 1991.
 139. Isaksen, I. S. A., T. Berntsen, and S. Solberg. Estimates of past and future tropospheric ozone changes from changes in human released source gases. In: Ozone in the Atmosphere, edited by R. D. Bojkov and P. Fabian. Hampton, VA: A. Deepak, 1989, p. 576–579.
 140. Ishii, T., H. Iwahashi, R. Sugata, and R. Kido. Superoxide dismutase enhances the toxicity of 3‐hydroxyanthranilic acid to bacteria. Free Radic. Res. Commun. 14: 187–194, 1991.
 141. Isidorov, V. A. Organic Chemistry of the Earth's Atmosphere, translated by E. A. Koroleva. New York: Springer‐Verlag, 1990.
 142. Jablonski, D. Extinctions: a paleontological perspective. Science 253: 754–757, 1991.
 143. Jamieson, D., B. Chance, E. Cadenas, and A. Boveris. The relation of free radical production to hyperoxia. Annu. Rev. Physiol. 48: 703–719, 1986.
 144. Jannasch, H. W., and C. D. Taylor. Deep‐sea microbiology. Annu. Rev. Microbiol. 38: 487–514, 1984.
 145. Jetten, M. S. M., A. J. Pierik, and W. R. Hagen. EPR characterization of a high‐spin system to carbon monoxide dehydrogenase from Methanothrix soehngenii. Eur. J. Biochem. 202: 1291–1297, 1991.
 146. Jones, T. P., and W. G. Chaloner. Fossil charcoal, its recognition and palaeoatmospheric significance. Palaeogeogr. Palaeoclimatol. Palaeoecol. (Global Planet. Change Sect.) 97: 39–50, 1991.
 147. Jones, W. J., D. P. Nagle, Jr., and W. B. Whitman. Methanogens and the diversity of archaebacteria. Microbiol. Rev. 51: 135–177, 1987.
 148. Joyce, G. F. RNA evolution and the origins of life. Nature 338: 217–224, 1989.
 149. Kasting, J. F. Bolide impacts and the oxidation state of carbon in the earth's early atmosphere. Orig. Life Evol. Biosph. 20: 199–231, 1990.
 150. Kasting, J. F. Models relating to Proterozoic atmospheric and ocean chemistry. In: The Proterozoic Biosphere. A Multidisciplinary Study, edited by J. W. Schopf and C. Klein. New York: Cambridge Univ. Press, 1992, p. 1185–1187.
 151. Kasting, J. F. Earth's early atmosphere. Science 259: 920–926, 1993.
 152. Kasting, J. F., H. D. Holland, and L. R. Kump. Atmospheric evolution: the rise of oxygen. In: The Proterozoic Biosphere. A Multidisciplinary Study, edited by J. W. Schopf and C. Klein. New York: Cambridge Univ. Press, 1992, p. 159–163.
 153. Keeling, R. F., and S. R. Shertz. Seasonal and interannual variations in atmospheric oxygen and implications for the global carbon cycle. Nature 358: 723–727, 1992.
 154. Keosian, J. Life's beginnings—origin or evolution? Orig. Life 5: 285–293, 1974.
 155. Kirchhoff, V. W. J. H., E. V. A. Marinho, P. L. S. Dias, E. B. Pereira, R. Calheiros, R. André, and C. Volpe. Enhancements of CO and O3 from burnings in sugar cane fields. J. Atmos. Chem. 12: 87–102, 1991.
 156. Klausner, R. D., and T. A. Rouault. A double life: cytosolic aconitase as a regulatory RNA binding protein. Mol. Biol. Cell 4: 1–5, 1993.
 157. Klausner, R. D., T. A. Rouault, and J. B. Harford. Regulating the fate of mRNA: the control of cellular iron metabolism. Cell 72: 19–28, 1993.
 158. Knoll, A. H. The early evolution of eukaryotes: a geological perspective. Science 256: 622–627, 1992.
 159. Kono, Y., and I. Fridovich. Functional significance of manganese catalase in Lactobacillus plantarum. J. Bacteriol. 155: 742–746, 1983.
 160. Kooyman, G. L. Pressure and the diver. Can. J. Zool. 66: 84–88, 1988.
 161. Korona, R., and B. R. Levin. Phage‐mediated selection and the evolution and maintenance of restriction‐modification. Evolution 47: 556–575, 1993.
 162. Koury, M. J., and M. C. Bondurant. Erythropoietin retards DNA breakdown and prevents programmed death in erythroid progenitor cells. Science 248: 378–381, 1990.
 163. Koury, M. J., and M. C. Bondurant. Control of red cell production: the roles of programmed cell death (apoptosis) and erythropoietin. Transfusion 30: 673–674, 1990.
 164. Krieg, N. R., and P. S. Hoffman. Microaerophily and oxygen toxicity. Annu. Rev. Microbiol. 40: 107–130, 1986.
 165. Krinsky, N. I. Mechanism of action of biological antioxidants. Proc. Soc. Exp. Biol. Med. 200: 248–254, 1992.
 166. Kuhn, W. R., and S. K. Atreya. Ammonia photolysis and the greenhouse effect in the primordial atmosphere of the Earth. Icarus 37: 207–213, 1979.
 167. Kump, L. R. Chemical stability of the atmosphere and ocean. Palaeogeogr. Palaeoclimatol. Palaeoecol. (Global Planet. Change Sect.) 75: 123–136, 1989.
 168. Kusunose, E., K. Ichihara, Y. Noda, and M. Kusunose. Superoxide dismutase from Mycobacterium tuberculosis. J. Biochem. 80: 1343–1352, 1976.
 169. Kutter, G. S. The Universe and Life. Origins and Evolution. Boston: Jones and Bartlett, 1987.
 170. Latimer, W. M. The Oxidation States of the Elements and Their Elements in Aqueous Solutions (2nd ed.). New York: Prentice‐Hall, 1952.
 171. Léger, A., L. d'Hendecourt, L. Verstraete, and P. Ehrenfreund. Small grains and large aromatic molecules. In: Chemistry in Space, edited by J. M. Greenberg and V. Pirronello. Boston: Kluwer, 1991, p. 211–225.
 172. Lelieveld, J., and P. J. Crutzen. The role of clouds in tropospheric photochemistry. J. Atmos. Chem. 12: 229–267, 1991.
 173. Levin, G. V., and L. R. Zehner. L‐sugars: Lev‐O‐Cal. In: Alternative Sweeteners (2nd ed.), edited by L. O. Nabors and R. C. Gelardi. New York: Dekker, 1991, p. 117–125.
 174. Lide, D. R. (Ed.). CRC Handbook of Chemistry and Physics. A Ready‐Reference Book of Chemical and Physical Data. 1992–1993 (73rd ed.). Boca Raton, FL: CRC, 1992.
 175. Little, G. H., and A. Flores. Inhibition of programmed cell death by catalase and phenylalanine methyl ester. Comp. Biochem. Physiol. 105A: 79–83, 1993.
 176. Lovelock, J. The Ages of Gaia. A Biography of Our Living Earth, New York: Bantam, 1990.
 177. Lowe, D. R. Major events in the geological development of the Precambrian Earth. In: The Proterozoic Biosphere. A Multidisciplinary Study, edited by J. W. Schopf and C. Klein. New York: Cambridge Univ. Press, 1992, p. 67–75.
 178. Lu, Y., and M. A. K. Khalil. Methane and carbon monoxide in OH chemistry: the effects of feedbacks and reservoirs generated by the reactive products. Chemosphere 26: 641–655, 1993.
 179. Luria, S. E., S. J. Gould, and S. Singer. A View of Life. Menlo Park, CA: Benjamin/Cummings, 1981.
 180. Ma, K., C. Zirngibl, D. Linder, K. O. Stetter, and R. K. Thauer. N5,N10‐Methylenetetrahydromethanopterin dehydrogenase (H2‐forming) from the extreme thermophile Methanopyrus kandleri. Arch. Microbiol. 156: 43–48, 1991.
 181. Macdermott, A. J. Electroweak enantioselection and the origin of life. Orig. Life Evol. Biosph. 25: 191–199, 1995.
 182. Maloney, P. C., and T. H. Wilson. The evolution of ion pumps. BioScience 35: 43–48, 1985.
 183. Margulis, L., and K. V. Schwartz. Five Kingdoms. An Illustrated Guide to the Phyla of Life on Earth (2nd ed.). New York: Freeman, 1988.
 184. Margulis, L., J. C. G. Walker, and M. Rambler. Reassessment of roles of oxygen and ultraviolet light in Precambrian evolution. Nature 264: 620–624, 1976.
 185. Martin, J. H., R. M. Gordon, S. E. Fitzwater. Iron in Antarctic waters. Nature 345: 156–158, 1990.
 186. Maskos, Z., and W. H. Koppemol. Oxyradicals and multivitamin tablets. Free Radic. Biol. Med. 11: 609–610, 1991.
 187. Matheson, R. The utilization of aquatic plants as aids in mosquito control. Am. Nat. 54: 56–86, 1930.
 188. Mauzerall, D. The photochemical origins of life and photore‐action of ferrous ion in the Archean oceans. Orig. Life Evol. Biosph. 20: 293–302, 1990.
 189. Mayer, M. G. The structure of the nucleus. Sci. Am. 184 (3): 22–26, 1951.
 190. McCormick, M. P., L. W. Thomason, and C. R. Trepte. Atmospheric effects of the Mt Pinatubo eruption. Nature 373: 399–404, 1995.
 191. McGervey, J. D. Introduction to Modern Physics. New York: Academic, 1971.
 192. McLaren, D. J., and W. D. Goodfellow. Geological and biological consequences of giant impacts. Annu. Rev. Earth Planet. Sci. 18: 123–171, 1990.
 193. Mekhail‐Ishak, K., N. Hudson, M. Tsao, and G. Batist. Implications for therapy of drug‐metabolizing enzymes in human colon cancer. Cancer Res. 49: 4866–4869, 1989.
 194. Melosh, H. J. Giant impacts and the thermal state of the early earth. In: Origin of the Earth, edited by H. E. Newsom and J. H. Jones. New York: Oxford Univ. Press, 1990, p. 69–83.
 195. Menzel, D. B. Ozone: an overview of its toxicity in man and animals. J. Toxic Environ. Health 13: 183–204, 1984.
 196. Meyer, D. M., M. Jura, I. Hawkins, and J. A. Cardelli. The abundance of interstellar oxygen toward Orion: evidence for recent infall. Astrophys. J. 437: L59–L61, 1994.
 197. Miller, S. L. A production of amino acids under possible primitive Earth conditions. Science 117: 528–529, 1953.
 198. Milne, L., P. Nicotera, S. Orrenius, and M. J. Burkitt. Effects of glutathione and chelating agents on copper‐mediated DNA oxidation: pro‐oxidant and antioxidant properties of glutathione. Arch. Biochem. Biophys. 304: 102–109, 1993.
 199. Milton, R. C. d., S. C. F. Milton, and S. B. H. Kent. Total chemical synthesis of a d‐enzyme: the enantiomers of HIV‐1 protease show demonstration of reciprocal chiral substrate specificity. Science 256: 1445–1448, 1992.
 200. Miyake, C., F. Michihata, and K. Asada. Scavenging of hydrogen peroxide in prokaryotic and eukaryotic algae: acquisition of ascorbate peroxidase during the evolution of cyanobacteria. Plant Cell. Physiol. 32: 33–43, 1991.
 201. Molina, M. J. Chemistry of stratospheric ozone depletion. In: Atmospheric Chemistry. Models and Predictions for Climate and Air Quality, edited by C. S. Sloane and T. W. Tesche. Chelsea, MI: Lewis, 1991, p. 1–8.
 202. Molina, M. J., and F. S. Rowland. Stratospheric sink for chlorofluoromethanes: chlorine atom‐catalysed destruction of ozone. Nature 249: 810–812, 1974.
 203. Moncada, S., and E. A. Higgs. Endogenous nitric oxide: physiology, pathology and clinical relevance. Eur. J. Clin. Invest. 21: 361–374, 1991.
 204. Moncada, S., R. M. J. Palmer, and E. A. Higgs. Nitric oxide: physiology, pathophysiology, and pharmacology. Pharmacol. Rev. 43: 109–142, 1991.
 205. Mooney, H. A. Photosynthesis. In: Plant Ecology, edited by M. J. Crawley. Boston: Blackwell Scientific, 1986, p. 345–373.
 206. Mukhin, L. M., Y. P. Dikov, E. N. Evlanov, M. N. Fomenkova, A. D. Grechinskiy, M. A. Nazarov, O. F. Prilutskiy, T. V. Ruzmaikina, R. Z. Sagdeev, and B. V. Zubkov. Chemical composition of Halley's dust component from the PUMA‐2 data. In: Chemistry in Space, edited by J. M. Greenberg and V. Pirronello. Boston: Kluwer, 1991, p. 399–414.
 207. National Oceanic and Atmospheric Administration. U.S. Standard Atmosphere, 1976. Washington, DC: National Oceanic and Atmospheric Administration, 1976.
 208. Neilands, J. B. Microbial iron transport compounds (siderophores) as chelating agents. In: Development of Iron Chelators for Clinical Use, edited by A. E. Martell, W. F. Anderson, and D. G. Badman. New York: Elsevier, 1981, p. 13–31.
 209. Neta. P., R. E. Huie, and A. B. Ross. Rate constants for reactions of peroxyl radicals in fluid solutions. J. Phys. Chem. Ref. Data 19: 413–513, 1990.
 210. Norby, R. J., C. A. Gunderson, S. D. Wullschleger, E. G. O'Neill, and M. K. McCracken. Productivity and compensatory responses of yellow‐polar trees in elevated CO2. Nature 357: 322–324, 1992.
 211. Ohanian, H. C. Modern Physics. Englewood Cliffs, NJ: Prentice‐Hall, 1987.
 212. Olinski, R., T. Zastawny, J. Budzbon, J. Skokowski, W. Zegarsky, and M. Dizdaroglu. DNA base modifications in chromatin of human cancerous tissues. FEBS Lett. 309: 193–198, 1992.
 213. Omont, A. Circumstellar chemistry. In: Chemistry in Space, edited by J. M. Greenberg and V. Pirronello. Boston: Kluwer, 1991, p. 171–197.
 214. Oparin, A. I. The Origin of Life, translated by S. Morgulis. New York: Dover, 1953.
 215. Orgel, L. E. Molecular replication. Nature 358: 203–209, 1992.
 216. Oró, J. Comets and the formation of biochemical compounds on the primitive earth. Nature 190: 389–390, 1961.
 217. Oró, J., T. Mills, and A. Lazcano. Comets and the formation of biochemical compounds on the primitive earth—a review. Orig. Life Evol. Biosph. 21: 267–277, 1992.
 218. Pagel, E. J. Beryllium and the Big Bang. Nature 354: 267–268, 1991.
 219. Parchment, R. E. The implications of a unified theory of programmed cell death, polyamines, oxyradicals and histogenesis in the embryo. Int. J. Dev. Biol. 37: 75–83, 1993.
 220. Peebles, P. J. E., D. N. Schramm, E. L. Turner, and R. G. Kron. The case for the relativistic hot Big Bang cosmology. Nature 352: 769–776, 1991.
 221. Peñuelas, J. High oxygen tension inhibits vascular aquatic plant growth in deep waters. Photosynthetica 21: 494–502, 1987.
 222. Pierce, M. J., D. L. Welch, R. D. McClure, S. Van den Bergh, R. Racine, and P. B. Stetson. The Hubble constant and Virgo cluster distance from observations of Cepheid variables. Nature 371: 385–389, 1994.
 223. Pihl, T. D., R. N. Schicho, R. M. Kelly, and R. J. Maier. Characterization of hydrogen‐uptake activity in the hyperthermophile Pyrodictium brockii. Proc. Natl. Acad. Sci. USA 86: 138–141, 1989.
 224. Podsiadlowski, P., J. E. Pringle, and M. J. Rees. The origin of the planet orbiting PSR1829–10. Nature 352: 783–784, 1991.
 225. Ponnamperuma, C., Y. Honda, and R. Navarro‐Conzález. Chemical studies on the existence of extraterrestrial life. J. Br. Interplanet. Soc. 45: 241–249, 1992.
 226. Ponnamperuma, C., A. Shimoyama, M. Yamada, T. Hobo, and R. Pal. Possible surface reaction on Mars: implications for Viking biology results. Science 197: 455–457, 1977.
 227. Pramanik, A. M., S. Bingsmark, M. Lindahl, H. Baltscheffsky, M. Baltscheffsky, and B. Andersson. Inorganic‐pyrophosphate‐dependent phosphorylation of spinach thylakoid proteins. Eur. J. Biochem. 198: 183–186, 1991.
 228. Prinn, G. G., and B. Fegley, Jr. The atmospheres of Venus, Earth, and Mars: a critical comparison. Annu. Rev. Earth Planet. Sci. 15: 171–212, 1987.
 229. Prosser, J. I. Autotrophic nitrification in bacteria. Adv. Microb. Physiol. 30: 125–181, 1989.
 230. Prusiner, S. B. Transgenic investigations of prion diseases of humans and animals. Phil. Trans. R. Soc. Lond. B 339: 239–254, 1993.
 231. Rabinowitz, D. L., T. Gehrels, J. V. Scotti, R. S. McMillan, M. L. Perry, W. Wisniewski, S. M. Larson, E. S. Howell, and E. A. Mueleer. Evidence for a near‐Earth asteroid belt. Nature 363: 704–706, 1993.
 232. Ragsdale, S. Enzymology of the acetyl‐CoA pathway of CO2 fixation. Crit. Rev. Biochem. Mol. Biol. 26: 261–300, 1991.
 233. Rahn, H., and L. E. Farhi. Gaseous environment and atelectasis. Federation Proc. 23: 1035–1041, 1963.
 234. Rast, W., and A. O. Ryding. Introduction. In: The Control of Eutrophication of Lakes and Reservoirs, edited by S. O. Ryding and W. Rast. Park Ridge, NJ: Parthenon, 1989, p. 1–4.
 235. Raven, J. A. Plant responses to high O2 concentrations: relevance to previous high O2 episodes. Palaeogeogr. Palaeoclimatol. Palaeoecol. (Global Planet. Change Sect.) 97: 19–38, 1991.
 236. Richardson, J. B. Origin and evolution of the earliest land plants. In: Major Events in the History of Life, edited by J. W. Schopf. Boston: Jones and Bartlett, 1992, p. 95–118.
 237. Richter, C. Pro‐oxidants and mitochondrial Ca2+: their relationship to apoptosis and oncogenesis. FEBS Lett. 325: 104–107, 1993.
 238. Robin, E., L. Froget, C. Jéhanno, and R. Rocchia. Evidence for a K/T impact event in the Pacific Ocean. Nature 363: 615–617, 1993.
 239. Robinson, J. M. Phanerozoic O2 variation, fire, and terrestrial ecology. Palaeogeogr. Palaeoclimatol. Palaeoecol. (Global Planet. Change Sect.) 75: 223–240, 1989.
 240. Robinson, J. M. Lignin, land plants, and fungi: biological evolution affecting Phanerozoic oxygen balance. Geology 15: 607–610, 1990.
 241. Robinson, J. M. Phanerozoic atmospheric reconstructions: a terrestrial perspective. Palaeogeogr. Palaeoclimatol. Palaeoecol. (Global Planet. Change Sec.) 97: 51–62, 1991.
 242. Rodriguez, J. M. Probing stratospheric ozone. Science 261: 1128–1129, 1993.
 243. Romankevich, Y. A. Biogeochemical aspects of the earth's living matter, translated from Geokhim. 2: 292–306. Geochemistry Int. 25: 123–136, 1988.
 244. Rospert, S., J. Breitung, K. Ma, B. Schwörer, C. Zirngibl, R. K. Thauer, D. Linder, R. Huber, and K. O. Stetter. Methylcoenzyme M reductase and other enzymes involved in methanogenesis from CO2 and H2 in the extreme thermophile Methanopyrus kandleri. Arch. Microbiol. 156: 49–55, 1991.
 245. Rowland, F. S. Ozone depletion theory. Science 261: 1102–1103, 1993.
 246. Rubbo, H., R. Radi, M. Trujillo, R. Telleri, B. Kalyanaraman, S. Barnes, M. Kirk, and B. A. Freeman. Nitric oxide regulation of superoxide and peroxynitrite‐dependent lipid peroxidation. Formation of novel nitrogen‐containing oxidized lipid derivatives. J. Biol. Chem. 269: 26066–26075, 1994.
 247. Rubey, W. W. Geologic history of sea water. An attempt to state the problem. Bull. Geol. Soc. Am. 62: 1111–1148, 1951.
 248. Runnegar, B. Precambrian oxygen levels estimated from the biochemistry and physiology of early eukaryotes. Palaeogeogr. Palaeoclimatol. Palaeoecol. (Global Planet. Change Sect.) 97: 97–111, 1991.
 249. Runnegar, B. Evolution of the earliest animals. In: Major Events in the History of Life, edited by J. W. Schopf. Boston: Jones and Bartlett, 1992, p. 65–93.
 250. Rusting, R. L. Why do we age? Sci. Am. 267: 130–135, 138–141, 1992.
 251. Sagan, C. Radiation and the origin of the gene. Evolution 11: 40–55, 1957.
 252. Sagan, C. On the origin and planetary distribution of life. Radiat. Res. 15: 174–192, 1961.
 253. Sagan, C. Ultraviolet selection pressure on the earliest organisms. J. Theoret. Biol. 39: 195–200, 1973.
 254. Sagan, C., and G. Mullen. Earth and Mars: evolution of atmospheres and surface temperatures. Science 177 (6): 52–56, 1972.
 255. Sahu, S. C., and G. C. Gray. Interactions of flavonoids, trace metals, and oxygen: nuclear DNA damage and lipid peroxidation induced by myricetin. Cancer Lett. 70: 73–79, 1993.
 256. Sanders, R. W., S. Solomon, M. A. Carroll, and A. L. Schmeltekopf. Ground‐based measurements of O3, NO2, OClO, and BrO during the 1987 Antarctic ozone depletion event. In: Ozone in the Atmosphere, edited by R. D. Bojkov and P. Fabian. Hampton, VA: A. Deepak, 1989, p. 65–69.
 257. Sasaki, S. The primary solar‐type atmosphere surrounding the accreting earth: H2O‐induced high surface temperature. In: Origin of the Earth, edited by H. E. Newsom and J. H. Jones. New York: Oxford Univ. Press, 1990, p. 195–209.
 258. Scheid, P., B. Pelster, and H. Kobayashi. Gas exchange in the fish swimbladder. In: Oxygen Transport to Tissue XII, edited by J. Piiper, T. K. Goldstick, and M. Meyer. New York: Plenum, 1990, p. 735–742.
 259. Schidlowski, M. A 3,800‐million‐year isotopic record of life from carbon in sedimentary rocks. Nature 333: 313–318, 1988.
 260. Schidlowski, M., J. M. Hayes, and I. R. Kaplan. Isotopic inferences of ancient biochemistries: carbon, sulfur, hydrogen, and nitrogen. In: Earth's Earliest Biosphere. Its Origin and Evolution, edited by J. W. Schopf. Princeton, NJ: Princeton Univ. Press, 1983, p. 149–186.
 261. Schlegel, H. G. Physiology and biochemistry of Knallgasbacteria. Adv. Comp. Physiol. Biochem. 2: 185–236, 1966.
 262. Schopf, J. W. (Ed). Major Events in the History of Life. Boston: Jones and Bartlett, 1992.
 263. Schopf, J. W. The oldest fossils and what they mean. In: Major Events in the History of Life, edited by J. W. Schopf. Boston: Jones and Bartlett, 1992, p. 29–63.
 264. Schopf, J. W. The oldest known records of life: Archean stromatolites, microfossils, and organic matter. In: Early Life on Earth: Nobel Symposium 84, edited by S. Bengtson. New York: Columbia Univ. Press, 1994, p. 192–206.
 265. Schopf, J. W. Microfossils of the Early Archean Apex Chert: new evidence of the antiquity of life. Science 260: 640–646, 1993.
 266. Schrödinger, E. What is Life? And Other Scientific Essays. Garden City, NY: Doubleday, 1956.
 267. Schroeder, P., and L. Ladd. Slowing the increase of atmospheric carbon dioxide: a biological approach. Climatic Change 19: 283–290, 1991.
 268. Schueler, F. W. Chemobiodynamics and Drug Design. New York: McGraw‐Hill, 1960.
 269. Schwartz, J. L., D. Z. Antoniades, and S. Zhao. Molecular and biochemical reprogramming of oncogenesis through the activity of prooxidants and antioxidants. Ann. NY Acad. Sci 686: 262–279, 1993.
 270. Schwartzman, R. A., and J. A. Cidlowski. Apoptosis: the biochemistry and molecular biology of programmed cell death. Endoc. Rev. 14: 133–151, 1993.
 271. Scott, D. H., J. W. Rice, Jr., and J. M. Dohm. Martian paleo‐lakes and waterways: exobiological implications. Orig. Life Evol. Biosph. 21: 189–198, 1991.
 272. Scotti, J. V., D. L. Rabinowitz, and B. G. Marsden. Near miss of the earth by a small asteroid. Nature 354: 287–289, 1991.
 273. Seiler, W., H. Giehl, and G. Bunse. The influence of plants on atmospheric carbon monoxide and dinitrogen oxide. Pure Appl. Geophys. (Pageoph) 116: 439–451, 1978.
 274. Seliger, H. H. Environmental photobiology. In: The Science of Photobiology, edited by K. C. Smith. New York: Plenum, 1977, p. 143–173.
 275. Shaffer, G. A model of biogeochemical cycling of phosphorus, nitrogen, oxygen, and sulphur in the ocean: one step toward a global climate model. J. Geophys. Res. 94: 1979–2004, 1989.
 276. Shapiro, J. Effect of yellow organic acids on iron and other metals in water. J. Am. Water Works Assoc. 56: 1062–1082, 1964.
 277. Shapiro, B. M. The control of oxidant stress at fertilization. Science 252: 533–536, 1991.
 278. Sharma, V. S., D. Bandyopadhyay, M. Berjis, J. Rifkind, and G. Boss. Double‐mixing kinetic studies of the reactions of monoliganded species of hemoglobin: α2(CO)1β2 and α2β2(CO)1. J. Biol. Chem. 266: 24492–24497, 1991.
 279. Sharpe, A. G. Inorganic Chemistry (3rd ed). New York and Essex, UK: Wiley and Longman, 1992.
 280. Sharpton, V. L., K. Burke, A. Carnago‐Zanoguera, S. A. Hall, D. S. Lee, L. E. Marín, G. Suárez‐Reynoso, J. M. Queaada‐Muñeton, P. D. Spudis, and J. Urrutia‐Fucugauchi. Chicxulub multiring impact basin: size and other characteristics derived from gravity analysis. Science 261: 1564–1567, 1993.
 281. Sharpton, V. L., G. B. Dalrymple, L. E. Marín, G. Ryder, B. C. Schuraytz, and J. Urrutia‐Fucugauchi. New links between the Chicxulub impact structure and the Cretaceous/Tertiary boundary. Nature 359: 819–821, 1992.
 282. Silk, J. The Big Bang. The Creation and Evolution of the Universe. San Francisco: Freeman, 1980.
 283. Silk, J. Slow‐motion galactic birth. Nature 351: 191, 1991.
 284. Sillén, L. G., and A. E. Martell. Stability Constants of Metal‐Ion Complexes. Special Publication 17. London: Burlington House, 1964.
 285. Simic, M. G. Urinary biomarkers and the rate of DNA damage in carcinogenesis and anticarcinogenesis. Mutat. Res. 267: 277–290, 1992.
 286. Sleep, N. H., K. J. Zahnle, J. F. Kasting, and H. J. Morowitz. Annihilation of ecosystems by large asteroid impacts on the early Earth. Nature 342: 139–142, 1989.
 287. Slonim, N. B., and L. H. Hamilton. Respiratory Physiology (4th ed.). St. Louis: Mosby, 1981.
 288. Smith, B. N. Evolution of C4 photosynthesis in response to changes in carbon and oxygen concentrations in the atmosphere through time. BioSystems 8: 24–32, 1976.
 289. Smith, C., B. Halliwell, and O. I. Aruoma. Protection by albumin against the pro‐oxidant actions of phenolic dietary components. Food Chem. Toxicol. 30: 483–489, 1992.
 290. Smith, R. C., B. B. Prézelin, K. S. Baker, R. R. Bidigare, N. P. Boucher, T. Coley, D. Karentz, S. MacIntyre, H. A. Matlick, D. Menzies, M. Ondrusek, Z. Wan, and K. J. Waters. Ozone depletion: ultraviolet radiation and phytoplankton biology in Antarctic waters. Science 255: 952–959, 1992.
 291. Spinrad, H., A. Dey, and J. R. Graham. Keck observations of the most distant galaxy: 8C 1435+63 AT z = 4.25. Astrophys. J. 438: L51–L54, 1995.
 292. Stadtman, E. R. Protein oxidation and aging. Science 257: 1220–1224, 1992.
 293. Stephenson, J. A. E., and M. W. J. Scourfield. Importance of energetic solar protons in ozone depletion. Nature 352: 137–139, 1991.
 294. Stolarski, R., R. Boijkov, L. Bishop, C. Zerefos, J. Staehelin, and J. Zawodny. Measure trends in stratospheric ozone. Science 256: 342–349, 1992.
 295. Szatriwski, T. P., and C. F. Nathan. Production of large amounts of hydrogen peroxide by human tumor cells. Cancer Res. 51: 794–798, 1991.
 296. Tabazadeh, A., and R. P. Turco. Stratospheric chlorine injection by volcanic eruptions: HCl scavenging and implications for ozone. Science 260: 1082–1085, 1993.
 297. Takao, M., A. Yasui, and A. Oikawa. Unique characteristics of a superoxide dismutase of a strictly anaerobic archaebacterium Methanobacterium thermoautotrophicum. J. Biol. Chem. 266: 14151–14152, 1991.
 298. Tappan, H. Extinction or survival: selectivity and causes of Phanerozoic crises. In: Geological Implications of Impacts of Large Asteroids and Comets on the Earth. Geological Special Paper 190, edited by L. T. Silver and P. Schultz. Boulder, CO: Geological Society of America, 1982, p. 265–276.
 299. Thompson, A. M. The oxidizing capacity of the earth's atmosphere: probable past and future. Science 256: 1157–1165, 1992.
 300. Tingle, T. N., J. A. Tyburczy, T. J. Ahrens, and C. H. Becker. The fate of organic matter during planetary accretion; preliminary studies of the organic chemistry of experimentally shocked Murchison meteorite. Orig. Life Evol. Biosph. 21: 385–397, 1992.
 301. Toumi, R., R. L. Jones, and J. A. Pyle. Stratospheric ozone depletion by ClONO2 photolysis. Nature 365: 37–39, 1993.
 302. Tyagi, S., and C. Ponnamperuma. Nonrandomness in prebiotic peptide synthesis. J. Mol. Evol. 30: 391–399, 1990.
 303. Urey, H. C. On the early chemical history of the Earth and the origin of life. Proc. Natl. Acad. Sci. USA 38: 351–363, 1952.
 304. Van den Bergh, S. The age and size of the universe. Science 258: 421–424, 1992.
 305. Verma, A., D. J. Hirsch, C. E. Glatt, G. V. Ronnett, and S. H. Snyder. Carbon monoxide: a putative neural messenger. Science 259: 381–384, 1993.
 306. von Gunter, U., and W. Schnedier. Primary products of the oxygenation of iron (II) at an oxic‐anoxic boundary: nucleation, aggregation, and aging. J. Coll. Interface Sci. 145: 127–139, 1991.
 307. Voßwinkel, R., I. Neidt, and H. Bothe. The production and utilization of nitric oxide by a new, denitrifying strain of Pseudomonas aeruginosa. Arch. Microbiol. 156: 62–69, 1991.
 308. Waldo, G. S., J. Ling, J. Sanders‐Loehr, and E. C. Theil. Formation of an Fe(III)‐tyrosinate complex during biomineralization of H‐subunit ferritin. Science 259: 796–798, 1993.
 309. Walker, J. C. G. Evolution of the Atmosphere. New York: Macmillan, 1977.
 310. Walker, J. C. G. How life affects the atmosphere. BioScience 34: 486–491, 1984.
 311. Walker, J. C. G. Iron and sulfur in the pre‐biologic ocean. Precambrian Res. 28: 205–222, 1985.
 312. Walker, J. C. G. Earth History. The Several Ages of the Earth. Boston: Jones and Bartlett, 1986.
 313. Walker, J. C. G. Global geochemical cycles of carbon, sulfur and oxygen. Marine Geology 70: 159–174, 1986.
 314. Walker, J. C. G., C. Klein, M. Schidlowski, J. W. Schopf, D. J. Stevenson, and M. R. Walter. Environmental evolution of the Archean‐Early Proterozoic earth: In: Earth's Earliest Biosphere: Its Origin and Evolution, edited by J. W. Schopf. Princeton, NJ: Princeton Univ. Press, 1983, p. 260–291.
 315. Warneck, P. Chemistry of the Natural Atmosphere. New York: Academic, 1988.
 316. Watson, J.D. Prologue. Early speculations and facts about RNA templates. In: The RNA World. The Nature of Modern RNA Suggests a Prebiotic RNA World, edited by R. F. Gesteland and J. F. Atkins. Plainview, NY: Cold Spring Harbor Lab., 1993, p. xv–xxiii.
 317. Watson, A., J. E. Lovelock, and L. Margulis. Methanogenesis, fires and the regulation of atmospheric oxygen. BioSystems 10: 293–298, 1978.
 318. Wege, K. Extremely low temperatures in the stratosphere and very low total ozone amount above Northern and Central Europe during winter 1989. J. Atmos. Chem. 12: 381–390, 1991.
 319. Weinberg, S. The First Three Minutes. A Modern View of the Origin of the Universe. New York: Basic Books, 1977.
 320. West, J. B. Respiratory Physiology—The Essentials (4th ed.). Baltimore: Williams & Wilkins, 1990, p. 53–54.
 321. Whitman, W. B. Methanogenic bacteria. In: The Bacteria. A Treatise on Structure and Function. Archaebacteria, edited by C. R. Woese and R. S. Wolfe. New York: Academic, 1985, vol. 8, p. 3–84.
 322. Woese, C. R. Bacterial evolution. Microbiol. Rev. 51: 221–271, 1987.
 323. Woese, C. R., O. Kandler, and M. L. Wheelis. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proc. Natl. Acad. Sci. USA 87: 4576–4579, 1990.
 324. Woese, C. R., L. J. Magrum, and G. E. Fox. Archaebacter. J. Mol. Evol. 11: 245–252, 1978.
 325. Wolbach, W. S., I. Gilmour, and E. Anders. Major wildfires at the Cretaceous/Tertiary boundary. In: Global Catastrophes in Earth History; An Interdisciplinary Conference on Impacts, Volcanism, and Mass Mortality. Geological Society of America Special Paper 247, edited by V. L. Sharpton and P. D. Ward. Geological Society of America: Boulder, CO, 1990, p. 391–400.
 326. Wong, S. S. M. Introductory Nuclear Physics. Englewood Cliffs, NJ: Prentice Hall, 1990.
 327. Wood, H. G. Life with CO or CO2 and H2 as a source of carbon and energy. FASEB J. 5: 156–163, 1991.
 328. Wyss, O., and P. W. Wilson. Mechanism of biological nitrogen fixation. VI. Inhibition of Azotobacter by hydrogen. Proc. Natl. Acad. Sci. USA 27: 162–168, 1941.
 329. Yarmolinsky, M. B. Programmed cell death in bacterial populations. Science 267: 836–837, 1995.
 330. Zelitch, I. Selection and characterization of tobacco plants with novel O2‐resistant photosynthesis. Plant Physiol. 90: 1457–1464, 1989.
 331. Zelitch, I. Physiological investigations of a tobacco mutant with O2‐resistant photosynthesis and enhanced catalase activity. Plant Physiol. 93: 1521–1524, 1990.
 332. Zhong, L. T., T. Sarafian, D. J. Kane, A. C. Charles, S. P. Mah, R. H. Edwards, and D. E. Bredesen. bcl‐2 inhibits death of central neural cells induced by multiple agents. Proc. Natl. Acad. Sci. USA 90: 4533–4537, 1993.
 333. Zhuang, G., and R. A. Duce. The adsorption of dissolved iron on marine aerosol particles in surface waters of the open ocean. Deep‐Sea Res. Part I 40: 1413–1429, 1993.
 334. Zuo, Y., and J. Hoigné. Evidence for photochemical formation of H2O2 and oxidation of SO2 in authentic fog water. Science 260: 71–73, 1993.
 335. Zurer, P. Ozone hits low levels over Antarctica, U. S. Chem. Eng. News Oct. 4: 5, 1993.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Daniel L. Gilbert. Evolutionary Aspects of Atmospheric Oxygen and Organisms. Compr Physiol 2011, Supplement 14: Handbook of Physiology, Environmental Physiology: 1059-1094. First published in print 1996. doi: 10.1002/cphy.cp040246