The Gray Goo Problem

March 20, 2001 by Robert A. Freitas Jr.

In Eric Drexler’s classic “grey goo” scenario, out-of-control nanotech replicators wipe out all life on Earth. This paper by Robert A. Freitas Jr. was the first quantitative technical analysis of this catastrophic scenario, also offering possible solutions. It was written in part as an answer to Bill Joy’s recent concerns.

Research Scientist, Zyvex

Originally published April 2000 as “Some Limits to Global Ecophagy by Biovorous Nanoreplicators, with Public Policy Recommendations.” Excerpted version published on KurzweilAI.net March 20, 2001.

Abstract

The maximum rate of global ecophagy by biovorous self-replicating nanorobots is fundamentally restricted by the replicative strategy employed; by the maximum dispersal velocity of mobile replicators; by operational energy and chemical element requirements; by the homeostatic resistance of biological ecologies to ecophagy; by ecophagic thermal pollution limits (ETPL); and most importantly by our determination and readiness to stop them.

Assuming current and foreseeable energy-dissipative designs requiring ~100 MJ/kg for chemical transformations (most likely for biovorous systems), ecophagy that proceeds slowly enough to add ~4°C to global warming (near the current threshold for immediate climatological detection) will require ~20 months to run to completion; faster ecophagic devices run hotter, allowing quicker detection by policing authorities. All ecophagic scenarios examined appear to permit early detection by vigilant monitoring, thus enabling rapid deployment of effective defensive instrumentalities.

Introduction

Recent discussions [1] of the possible dangers posed by future technologies such as artificial intelligence, genetic engineering and molecular nanotechnology have made it clear that an intensive theoretical analysis of the major classes of environmental risks of molecular nanotechnology (MNT) is warranted. No systematic assessment of the risks and limitations of MNT-based technologies has yet been attempted. This paper represents a first effort to begin this analytical process in a quantitative fashion.

Perhaps the earliest-recognized and best-known danger of molecular nanotechnology is the risk that self-replicating nanorobots capable of functioning autonomously in the natural environment could quickly convert that natural environment (e.g., “biomass”) into replicas of themselves (e.g., “nanomass”) on a global basis, a scenario usually referred to as the “gray goo problem” but perhaps more properly termed “global ecophagy.”

As Drexler first warned in Engines of Creation [2]:

“Plants” with “leaves” no more efficient than today’s solar cells could out-compete real plants, crowding the biosphere with an inedible foliage. Tough omnivorous “bacteria” could out-compete real bacteria: They could spread like blowing pollen, replicate swiftly, and reduce the biosphere to dust in a matter of days. Dangerous replicators could easily be too tough, small, and rapidly spreading to stop–at least if we make no preparation. We have trouble enough controlling viruses and fruit flies.

Among the cognoscenti of nanotechnology, this threat has become known as the “gray goo problem.” Though masses of uncontrolled replicators need not be gray or gooey, the term “gray goo” emphasizes that replicators able to obliterate life might be less inspiring than a single species of crabgrass. They might be superior in an evolutionary sense, but this need not make them valuable.

The gray goo threat makes one thing perfectly clear: We cannot afford certain kinds of accidents with replicating assemblers.

Gray goo would surely be a depressing ending to our human adventure on Earth, far worse than mere fire or ice, and one that could stem from a simple laboratory accident.

Lederberg [3] notes that the microbial world is evolving at a fast pace, and suggests that our survival may depend upon embracing a “more microbial point of view.” The emergence of new infectious agents such as HIV and Ebola demonstrates that we have as yet little knowledge of how natural or technological disruptions to the environment might trigger mutations in known organisms or unknown extant organisms [81], producing a limited form of “green goo” [92].

However, biovorous nanorobots capable of comprehensive ecophagy will not be easy to build and their design will require exquisite attention to numerous complex specifications and operational challenges. Such biovores can emerge only after a lengthy period of purposeful focused effort, or as a result of deliberate experiments aimed at creating general-purpose artificial life, perhaps by employing genetic algorithms, and are highly unlikely to arise solely by accident.

The Ecophagic Threat

Classical molecular nanotechnology [2], 4] envisions nanomachines predominantly composed of carbon-rich diamondoid materials. Other useful nanochemistries might employ aluminum-rich sapphire (Al2O3) materials, boron-rich (BN) or titanium-rich (TiC) materials, and the like. TiC has one the highest possible operating temperatures allowed for commonplace materials (melting point ~3410°K [5]), and while diamond can scratch TiC, TiC can be used to melt diamond.

However, atoms of Al, Ti and B are far more abundant in the Earth’s crust (81,300 ppm, 4400 ppm and 3 ppm, respectively [5]) than in biomass, e.g., the human body (0.1 ppm, 0 ppm, and 0.03 ppm [6]), reducing the direct threat of ecophagy by such systems. On the other hand, carbon is a thousand times less abundant in crustal rocks (320 ppm, mostly carbonates) than in the biosphere (~230,000 ppm).

Furthermore, conversion of the lithosphere into nanomachinery is not a primary concern because ordinary rocks typically contain relatively scarce sources of energy. For instance, natural radioactive isotopes present in crustal rocks vary greatly as a function of the geological composition and history of a region, but generally range from 0.15-1.40 mGy/yr [7], giving a raw power density of 0.28-2.6 ×10-7 W/m3 assuming crustal rocks of approximately mean terrestrial density (5522 kg/m3 [5]).

This is quite insufficient to power nanorobots capable of significant activities; current nanomachine designs typically require power densities on the order of 105-109 W/m3 to achieve effective results [6]. (Biological systems typically operate at 102-106 W/m3 [6].) Solar power is not readily available below the surface, and the mean geothermal heat flow is only 0.05 W/m2 at the surface [6], just a tiny fraction of solar insolation.

Hypothesized crustal abiotic highly-reduced petroleum reserves [16] probably could not energize significant replicator nanomass growth due to the anoxic environment deep underground, although potentially large geobacterial populations have been described [1016] and in principle some unusual though highly limited bacterial energy sources could also be tapped by nanorobots.

For example, some anaerobic bacteria use metals (instead of oxygen) as electron-acceptors [13], with iron present in minerals such as pyroxene or olivine being converted to iron in a more oxidized state in magnetic minerals such as magnetite and maghemite, and using geochemically produced hydrogen to reduce CO2 to methane [11]. Underground bacteria in the Antrim Shale deposit produce 1.2 ×107 m3/day of natural gas (methane) by consuming the 370 MY-old remains of ancient algae [17].

Bioremediation experiments have also been done by Envirogen and others in which pollution-eating bacteria are purposely injected into the ground to metabolize organic toxins; in field tests it has proven difficult to get the bacteria to move through underground aquifers, because the negatively-charged cells tend to adhere to positively charged iron oxides in the soil [18].

However, the primary ecophagic concern is that runaway nanorobotic replicators or “replibots” will convert the entire surface biosphere (the ecology of all living things on the surface of the Earth) into alternative or artificial materials of some type–especially, materials like themselves, e.g., more self-replicating nanorobots.

Since advanced nanorobots might be constructed predominantly of carbon-rich diamondoid materials [4], and since ~12% of all atoms in the human body (representative of biology generally) are carbon atoms [6], or ~23% by weight, the global biological carbon inventory may support the self-manufacture of a final mass of replicating diamondoid nanorobots on the order of ~0.23 Mbio, where Mbio is the total global biomass.

Unlike almost any other natural material, biomass can serve both as a source of carbon and as a source of power for nanomachine replication. Ecophagic nanorobots would regard living things as environmental carbon accumulators, and biomass as a valuable ore to be mined for carbon and energy. Of course, biosystems from which all carbon has been extracted can no longer be alive but would instead become lifeless chemical sludge.

Additional Scenarios

Four related scenarios which may lead indirectly to global ecophagy have been identified and are described below. In all cases, early detection appears feasible with advance preparation, and adequate defenses are readily conceived using molecular nanotechnologies of comparable sophistication.

Gray Plankton

The existence of 1-2 ×1016 kg [24] of global undersea carbon storage on continental margins as CH4 clathrates and a like amount (3.8 ×1016 kg) of seawater-dissolved carbon as CO2 represent a carbon inventory more than an order of magnitude larger than in the global biomass. Methane and CO2 can in principle be combined to form free carbon and water, plus 0.5 MJ/kg C of free energy. (Some researchers are studying the possibility of reducing greenhouse gas accumulations by storing liquid [44] or solid [45] CO2 on the ocean floor, which could potentially enable seabed replibots to more easily metabolize methane sources.)

Oxygen could also be imported from the surface in pressurized microtanks via buoyancy transport, with the conversion of carbon clathrates to nanomass taking place on the seabed below. The subsequent colonization of the land-based carbon-rich ecology by a large and hungry seabed-grown replicator population is the “gray plankton” scenario. (Phytoplankton, 1-200 microns in size, are the particles most responsible for the variable optical properties of oceanic water because of the strong absorption of these cells in the blue and red portions of the optical spectrum [37].)

If not largely confined to the sea floor during most of their replication cycle, the natural cell/device ratio could increase by many orders of magnitude, requiring a more diligent census effort. Census-taking nanorobots can alternatively be used to identify, disable, knapsack or destroy the gray plankton devices.

Gray Dust (Aerovores)

Traditional diamondoid nanomachinery designs [4] have employed 8 primary chemical elements, along with the associated atmospheric abundances [46] of each element. (Silicon is present in air as particulate dust which may be taken as ~28% Si for crustal rock [5], with a global average dust concentration of ~0.0025 mg/m3). The requirement for elements that are relatively rare in the atmosphere greatly constrains the potential nanomass and growth rate of airborne replicators.

However, note that at least one of the classical designs exceeds 91% CHON by weight. Although it would be very difficult, it is at least theoretically possible that replicators could be constructed almost solely of CHON, in which case such devices could replicate relatively rapidly using only atmospheric resources, powered by sunlight. A worldwide blanket of airborne replicating dust or “aerovores” that blots out all sunlight has been called the “gray dust” scenario [47]. (There have already been numerous experimental aerial releases of recombinant bacteria [48].)

The most efficient cleanup strategy appears to be the use of air-dropped non-self-replicating nanorobots equipped with prehensile microdragnets.

Alternative airborne or ground-based atmospheric filtration configurations that could permit more rapid filtering are readily envisioned. For example, since drag power varies as the square of the velocity, then by increasing mesh volume 10,000-fold while decreasing airflow velocity 100-fold, total drag power remains unchanged but whole-atmosphere turnover proceeds 100-fold faster, e.g., ~15 minutes.

Gray Lichens

Colonies of symbiotic algae and fungi known as lichens (which some have called a form of sub-aerial biofilm) are among the first plants to grow on bare stone, helping in soil formation by slowly etching the rock [55]. Lithobiontic microbial communities such as crustose saxicolous lichens penetrate mineral surfaces up to depths of 1 cm using a complex dissolution, selective transport, and recrystallization process sometimes termed “biological weathering” [56].

Colonies of epilithic (living on rock surfaces) microscopic bacteria produce a 10 micron thick patina on desert rocks (called “desert varnish” [57]) consisting of trace amounts of Mn and Fe oxides that help to provide protection from heat and UV radiation [5759].

In theory, replicating nanorobots could be made almost entirely of nondiamondoid materials including noncarbon chemical elements found in great abundance in rock such as silicon, aluminum, iron, titanium and oxygen. The subsequent ecophagic destruction of land-based biology by a maliciously programmed noncarbon epilithic replicator population that has grown into a significant nanomass is the “gray lichen” scenario.

Continuous direct census sampling of the Earth’s land surfaces will almost certainly allow early detection, since mineralogical nanorobots should be easily distinguishable from inert rock particles and from organic microbes in the top 3-8 cm of soil.

Malicious Ecophagy

More difficult scenarios involve ecophagic attacks that are launched not to convert biomass to nanomass, but rather primarily to destroy biomass. The optimal malicious ecophagic attack strategy appears to involve a two-phase process.

In the first phase, initial seed replibots are widely distributed in the vicinity of the target biomass, replicating with maximum stealth up to some critical population size by consuming local environmental substrate to build nanomass. In the second phase, the now-large replibot population ceases replication and exclusively undertakes its primary destructive purpose. More generally, this strategy may be described as Build/Destroy.

During the Build phase of the malicious “badbots,” and assuming technological equivalence, defensive “goodbots” enjoy at least three important tactical advantages over their adversaries:

1. Preparation–defensive agencies can manufacture and position in advance overwhelming quantities of (ideally, non-self-replicating) defensive instrumentalities, e.g., goodbots, which can immediately be deployed at the first sign of trouble, with minimal additional risk to the environment;

2. Efficiency–while badbots must simultaneously replicate and defend themselves against attack (either actively or by maintaining stealth), goodbots may concentrate exclusively on attacking badbots (e.g., because of their large numerical superiority in an early deployment) and thus enjoy lower operational overhead and higher efficiency in achieving their purpose, all else equal; and

3. Leverage–in terms of materials, energy, time and sophistication, fewer resources are generally required to confine, disable, or destroy a complex machine than are required to build or replicate the same complex machine from scratch (e.g., one small bomb can destroy a large bomb-making factory; one small missile can sink a large ship).

It is most advantageous to engage a malicious ecophagic threat while it is still in its Build phase. This requires foresight and a commitment to extensive surveillance by the defensive authorities.

Conclusions and Public Policy Recommendations

The smallest plausible biovorous nanoreplicator has a molecular weight of ~1 gigadalton and a minimum replication time of perhaps ~100 seconds, in theory permitting global ecophagy to be completed in as few as ~104 seconds. However, such rapid replication creates an immediately detectable thermal signature enabling effective defensive policing instrumentalities to be promptly deployed before significant damage to the ecology can occur.

Such defensive instrumentalities will generate their own thermal pollution during defensive operations. This should not significantly limit the defense strategy because knapsacking, disabling or destroying a working nanoreplicator should consume far less energy than is consumed by a nanoreplicator during a single replication cycle, hence such defensive operations are effectively endothermic.

Ecophagy that proceeds near the current threshold for immediate climatological detection, adding perhaps ~4°C to global warming, may require ~20 months to run to completion, which is plenty of advance warning to mount an effective defense.

Ecophagy that progresses slowly enough to evade easy detection by thermal monitoring alone would require many years to run to completion, could still be detected by direct in situ surveillance, and may be at least partially offset by increased biomass growth rates due to natural homeostatic compensation mechanisms inherent in the terrestrial ecology.

Ecophagy accomplished indirectly by a replibot population pre-grown on nonbiological substrate may be avoided by diligent thermal monitoring and direct census sampling of relevant terrestrial niches to search for growing, possibly dangerous, pre-ecophagous nanorobot populations.

Specific public policy recommendations suggested by the results of the present analysis include:

1. An immediate international moratorium on all artificial life experiments implemented as nonbiological hardware. In this context, “artificial life” is defined as autonomous foraging replicators, excluding purely biological implementations (already covered by NIH guidelines [65] tacitly accepted worldwide) and also excluding software simulations which are essential preparatory work and should continue. Alternative “inherently safe” replication strategies such as the broadcast architecture [66] are already well-known.

2. Continuous comprehensive infrared surveillance of Earth’s surface by geostationary satellites, both to monitor the current biomass inventory and to detect (and then investigate) any rapidly-developing artificial hotspots. This could be an extension of current or proposed Earth-monitoring systems (e.g., NASA’s Earth Observing System [67]and disease remote-sensing programs [93]) originally intended to understand and predict global warming, changes in land use, and so forth–initially using non-nanoscale technologies. Other methods of detection are feasible and further research is required to identify and properly evaluate the full range of alternatives.

3. Initiating a long-term research program designed to acquire the knowledge and capability needed to counteract ecophagic replicators, including scenario-building and threat analysis with numerical simulations, measure/countermeasure analysis, theory and design of global monitoring systems capable of fast detection and response, IFF (Identification Friend or Foe) discrimination protocols, and eventually the design of relevant nanorobotic systemic defensive capabilities and infrastructure.

A related long-term recommendation is to initiate a global system of comprehensive in situ ecosphere surveillance, potentially including possible nanorobot activity signatures (e.g. changes in greenhouse gas concentrations), multispectral surface imaging to detect disguised signatures, and direct local nanorobot census sampling on land, sea, and air, as warranted by the pace of development of new MNT capabilities.

Acknowledgments

The author thanks Robert J. Bradbury, J. Storrs Hall, James Logajan, Markus Krummenacker, Thomas McKendree, Ralph C. Merkle, Christopher J. Phoenix, Tihamer Toth-Fejel, James R. Von Ehr II, and Eliezer S. Yudkowsky for helpful comments on earlier versions of this manuscript; J. S. Hall for the word “aerovore”; and R. J. Bradbury for preparing the hypertext version of this document.

References

1. Bill Joy, “Why the future doesn’t need us,” Wired (April 2000); response by Ralph Merkle, “Text of prepared comments by Ralph C. Merkle at the April 1, 2000 Stanford Symposium organized by Douglas Hofstadter“.

2. K. Eric Drexler, “Engines of Creation: The Coming Era of Nanotechnology,” Anchor Press/Doubleday, New York, 1986. See:.

3. Joshua Lederberg, “Infectious History,” Science288(14 April 2000):287-293.

4. K. Eric Drexler, Nanosystems: Molecular Machinery, Manufacturing, and Computation, John Wiley & Sons, NY, 1992.

5. Robert C. Weast, Handbook of Chemistry and Physics, 49th Edition, CRC, Cleveland OH, 1968.

6. Robert A. Freitas Jr., Nanomedicine, Volume I, Landes Bioscience, Georgetown, TX, 1999. See at: http://www. nanomedicine.com.

7. Edward L. Alpen, Radiation Biophysics, Second Edition, Academic Press, New York, 1998.

8. Walter M. Elsasser, “Earth,” Encyclopedia Britannica 7 (1963):845-852.

9. G. Buntebarth, A. Gliko, “Heat Flow in the Earth’s Crust and Mantle,” in A.S. Marfunin, ed., Advanced Mineralogy, Volume 1: Composition, Structure, and Properties of Mineral Matter: Concepts, Results, and Problems, Springer-Verlag, New York, 1994, pp. 430-435.

10. Karsten Pedersen, “The deep subterranean biosphere,” Earth Sci. Rev. 34(1993):243-260.

11. Todd O. Stevens, James P. McKinley, “Lithoautotrophic Microbial Ecosystems in Deep Basalt Aquifers,” Science270(20 October 1995):450-454; see also: G. Jeffrey Taylor, “Life Underground,” PSR Discoveries, 21 December 1996.

12. Stephen Jay Gould, Life’s Grandeur: The Spread of Excellence from Plato to Darwin, Jonathan Cape, 1996.

13. Bill Cabage, “Digging Deeply,” September 1996.

14. James K. Fredrickson, Tullis C. Onstott, “Microbes Deep Inside the Earth,” Sci. Am. 275(October 1996):68-73.

15. Richard Monastersky, “Deep Dwellers: Microbes Thrive Far Below Ground,” Science News151(29 March 1997):192-193.

16. Thomas Gold, The Deep Hot Biosphere, Copernicus Books, 1999; “The deep, hot biosphere,” Proc. Natl. Acad. Sci. 89(1992):6045-6049. See also: P.N. Kropotkin, “Degassing of the Earth and the Origin of Hydrocarbons,” Intl. Geol. Rev. 27(1985):1261-1275.

17. Karl Leif Bates, “Michigan’s natural gas fields: Blame it on underground bacteria,” The Detroit News, 12 September 1996.

18. JoAnn Gutin, “Making Bacteria Move,” Princeton Weekly Bulletin, 17 November 1997.

19. Robert A. Freitas Jr., William P. Gilbreath, eds., Advanced Automation for Space Missions, Proceedings of the 1980 NASA/ASEE Summer Study held at the University of Santa Clara, Santa Clara, CA, June 23-August 29, 1980; NASA Conference Publication CP-2255, November 1982.

20. R.K. Dixon, S. Brown, R.A. Houghton, A.M. Solomon, M.C. Trexler, J. Wisniewski, “Carbon Pools and Flux of Global Forest Ecosystems,” Science263(14 January 1994):185-190.

21. Christopher B. Field, Michael J. Behrenfeld, James T. Randerson, Paul Falkowski, “Primary Production of the Biosphere: Integrating Terrestrial and Oceanic Components,” Science 281(10 July 1998):237-240.

22. Peter M. Vitousek, Harold A. Mooney, Jane Lubchenco, Jerry M. Melillo, “Human Domination of Earth’s Ecosystems,” Science277(25 July 1997):494-499.

23. Colin J. Campbell, Jean H. Laherrere, “The End of Cheap Oil,” Scientific American 278(March 1998):78-83; Robert G. Riley Enterprises, “World Petroleum Reserves,” 1999; L.F. Ivanhoe, “Future world oil supplies: There is a finite limit,” World Oil, October 1995.

24. James P. Kennett, Kevin G. Cannariato, Ingrid L. Hendy, Richard J. Behl, “Carbon Isotopic Evidence for Methane Hydrate Instability During Quaternary Interstadials,” Science 288(7 April 2000):128-133.

25. World Coal Institute, “Coal–Power for Progress,” Third Edition, January 1999, Statistics Canada, “World Coal Reserves,” 1996; “U.S. Coal Reserves: 1997 Update,” February 1999, Energy Information Administration, Washington, DC.

26. F.J. Millero, “Thermodynamics of the carbon dioxide system in the oceans,” Geochim. Cosmochim. Acta59(1995):661-677; see also F.J. Millero, “Carbon Dioxide in the South Pacific“.

27. Michael T. Madigan, John M. Martinko, Jack Parker, eds., Brock’s Biology of Microorganisms, 9th Edition, Prentice-Hall, NJ, 1999; Kenneth J. Ryan, ed., Sherris Medical Microbiology: An Introduction to Infectious Diseases, 3rd Edition, McGraw-Hill, New York, 1994.

28. ORNL, “Major World Ecosystem Complexes Ranked by Carbon in Live Vegetation,” April 1997.

29. J.H. Martin, The IronEx Group, “Testing the iron hypothesis in the ecosystems of the equatorial Pacific Ocean,” Nature 371(1994):123-129; Sallie W. Chisholm, “The iron hypothesis: Basic research meets environmental policy,” Rev. Geophys. 33(1995):Supplement. See also: “Extra iron makes blue deserts bloom,” New Scientist 152(12 October 1996).

30. Richard W. Hughes, Ruby & Sapphire, RWH Publishing, Boulder CO, 1997.

31. F. Albert Cotton, Geoffrey Wilkinson, Advanced Inorganic Chemistry: A Comprehensive Text, Second Edition, John Wiley & Sons, New York, 1966.

32. Ralph C. Merkle, personal communication, 22 March 2000.

33. P.G. Jarvis, Tree Physiol.2(1986):347-.

34. Oliver L. Phillips et al, “Changes in the Carbon Balance of Tropical Forests: Evidence from Long-Term Plots,” Science282(16 October 1998):439-442.

35. S. Fan, M. Gloor, J. Mahlman, S. Pacala, J. Sarmiento, T. Takahashi, P. Tans, “A Large Terrestrial Carbon Sink in North America Implied by Atmospheric and Oceanic Carbon Dioxide Data and Models,” Science 282(16 October 1998):442-446.

36. D. Stramski, D.A. Kiefer, “Light Scattering by Microorganisms in the Open Ocean,” Prog. Oceanogr.28(1991):343.

37. Curtis D. Mobley, “Chapter 43. The Optical Properties of Water,” in Michael Bass, ed., Handbook of Optics, Volume I, McGraw-Hill, Inc., New York, 1995, pp. 43.3-43.56.

38. Neil A. Campbell, Jane B. Reece, Lawrence G. Mitchell, Biology–Interactive Study Guide, Benjamin/Cummings Science, San Francisco, CA, 1999. See also: Paul Broady, “BIOL 113–Diversity of Life,” lecture notes.

39. William B. Whitman, David C. Coleman, “Prokaryotes: the unseen majority,” Proc. Natl. Acad. Sci. (USA) 94(June 1998):6578-6583.

40. B.R. Strain, J.D. Cure, eds., Direct Effects of Increasing Carbon Dioxide on Vegetation, Publ. ER-0238, U.S. Department of Energy, Washington, DC, 1985; R.J. Luxmoore, R.J. Norby, E.G. O’Neill, in Forest Plants and Forest Protection, 18th Intl. Union of Forestry Research Organizations (IUFRO), World Congress, Div. 2, 1987, IUFRO Secretariate, Vienna, 1987, Vol. 1, pp. 178-183; P.S. Curtis, B.G. Drake, P.W. Leadley, W.J. Arp, D.F. Whigham, Oecologia 78(1989):20; D. Eamus, P.G. Jarvis, Adv. Ecol. Res. 19(1989):1; P.G. Jarvis, Philos. Trans. R. Soc. London B 324(1989):369; R.J. Norby, E.G. O’Neill, New Phytol.117(1991):515.

41. Christian Korner, John A. Arnone III, “Responses to Elevated Carbon Dioxide in Artificial Tropical Ecosystems,” Science257(18 September 1992):1672-1675.

42. Eric T. Sundquist, “The Global Carbon Dioxide Budget,” Science 259(12 February 1993):934-941.

43. Hubertus Fischer, Martin Wahlen, Jesse Smith, Derek Mastroianni, Druce Deck, “Ice Core Records of Atmospheric CO2 Around the Last Three Glacial Terminations,” Science 283(12 March 1999):1712-1714.

44. Peter G. Brewer, Gernot Friederich, Edward T. Peltzer, Franklin M. Orr Jr., “Direct Experiments on the Ocean Disposal of Fossil Fuel CO2,” Science 284(7 May 1999):943-945; “Ocean studied for carbon dioxide storage,” 10 May 1999.

45. C.N. Murray, L. Visintini, G. Bidoglio, B. Henry, “Permanent Storage of Carbon Dioxide in the Marine Environment: The Solid CO2 Penetrator,” Energy Convers. Mgmt.37(1996):1067-1072.

46. Dennis K. Killinger, James H. Churnside, Laurence S. Rothman, “Chapter 14. Atmospheric Optics,” in Michael Bass, Eric W. Van Stryland, David R. Williams, William L. Wolfe, eds., Handbook of Optics, Volume I: Fundamentals, Techniques, and Design, Second Edition, McGraw-Hill, Inc., New York, 1995, pp. 44.1-44.50.

47. Ralph C. Merkle, personal communication, 6 April 2000.

48. Guy R. Knudsen, Louise-Marie C. Dandurand, “Model for Dispersal and Epiphytic Survival of Bacteria Applied to Crop Foliage,” paper presented at the 7th Symposium on Environmental Releases of Biotechnology Products: Risk Assessment Methods and Research Progress, 6-8 June 1995, Pensacola, FL.

49. Jake Page, “Making the Chips that Run the World,” Smithsonian 30(January 2000):36-46.

50. A. Borghesi, G. Guizzetti, “Graphite (C),” in Edward D. Palik, ed., Handbook of Optical Constants of Solids II, Academic Press, New York, 1991, pp. 449-460.

51. B. Ranby, J.F. Rabek, Photodegradation, Photo-oxidation and Photostabilization of Polymers, John Wiley & Sons, New York, 1975.

52. William S. Spector, ed., Handbook of Biological Data, W.B. Saunders Company, Philadelphia PA, 1956.

53. W.J. Kowalski, William Bahnfleth, “Airborne Respiratory Diseases and Mechanical Systems for Control of Microbes,” HPAC (July 1998).

54. M. Edmund Speare, Wayne Anthony McCurdy, Allan Grierson, “Coal and Coal Mining,” Encyclopedia Britannica5(1963):961-975; Helmut E. Landsberg, “Dust,” Encyclopedia Britannica7(1963):787-791; and Gerrit Willem Hendrik Schepers, “Pneumonoconiosis,” Encyclopedia Britannica 18(1963):99-100.

55. T.H. Nash, Lichen Biology, Cambridge University Press, Cambridge, 1996.

56. W.W. Barker, J.F. Banfield, “Biologically- versus inorganically-mediated weathering: relationships between minerals and extracellular polysaccharides in lithobiontic communities,” Chemical Geology132(1996):55-69; J.F. Banfield, W.W. Barker, S.A. Welch, A. Taunton, “Biological impact on mineral dissolution: Application of the lichen model to understanding mineral weathering in the rhizosphere,” Proc. Nat. Acad. Sci. (USA) 96(1999):3404-3411. See also: W.W. Barker, “Interactions between silicate minerals and lithobiontic microbial communities (lichens),”.

57. Ronald L. Dorn, Theodore M. Oberlander, “Microbial Origin of Desert Varnish,” Science 213(1981):1245-1247; R.L. Dorn, “Rock varnish,” Amer. Sci. 79(1991):542-553.

58. W.W. Barker, S.A. Welch, S. Chu, J.F. Banfield, “Experimental observations of the effects of bacteria on aluminosilicate weathering,” Amer. Mineral.83(1998):1551-1563.

59. S.A. Welch, W.W. Barker, J.F. Banfield, “Microbial extracellular polysaccharides and plagioclase dissolution,” Geochim. Cosmochim. Acta 63(1999):1405-1419.

60. K.L. Temple, A.R. Colmer, “The autotrophic oxidation of iron by a new bacterium, Thiobacillus ferrooxidans,” J. Bacteriol. 62(1951):605-611.

61. P.A. Trudinger, “Microbes, Metals, and Minerals,” Minerals Sci. Eng. 3(1971):13-25; C.L. Brierley, “Bacterial Leaching,” CRC Crit. Rev. Microbiol. 6(1978):207-262; “Microbiological mining,” Sci. Am. 247(February 1982):44-53.

62. A. Okereke, S.E. Stevens, “Kinetics of iron oxidation by Thiobacillus ferrooxidans,” Appl. Environ. Microbiol. 57(1991):1052-1056.

63. Verena Peters, Peter H. Janssen, Ralf Conrad, “Transient Production of Formate During Chemolithotrophic Growth of Anaerobic Microorganisms on Hydrogen,” Curr. Microbiol. 38(1999):285-289.

64. Mark S. Coyne, “Lecture 24–Biogeochemical Cycling: Soil Mineral Transformations of Metals,” Agripedia: Introductory Soil Biology; “Lecture 3–Soil as a Microbial Habitat: Microbial Distribution,” Agripedia: Introductory Soil Biology.

65. “NIH Guidelines for Research Involving Recombinant DNA Molecules,” January 1996 revision.

66. Ralph C. Merkle, “Self-replicating systems and low cost manufacturing,” in M.E. Welland, J.K. Gimzewski, eds., The Ultimate Limits of Fabrication and Measurement, Kluwer, Dordrecht, 1994, pp. 25-32.

67. “Links to Earth Observing System (EOS) Data and Information.”

68. Paul E. Tiegs, “Design and Operating Factors Which Affect Emissions from Residential Wood-Fired Heaters: Review and Update,” 22 June 1995; Stephen Black, A.B. Donaldson, “Some Observations on Operation of a Diesel Engine With Ethanol and Ethanol-Water Blends and Combustion Air Preheat,” Spring 1998; A. Ngaloken Gintings et al, “The relationship between waste wood management and the risk of transboundary haze from forest fire,” 17 December 1998.

69. World Resources Institute, World Resources 1988-89, Basic Books, Inc., New York, 1988, p. 169; EPA, Federal Register 61(13 December 1996):657-63.

70. Sankar Chatterjee, The Rise of Birds: 225 Million Years of Evolution, Johns Hopkins University Press, Baltimore, MD, 1997.

71. Paul R. Ehrlich, David S. Dobkin, Darryl Wheye, “Adaptations for Flight,” 1988.

72. H. J. Morowitz, M. E. Tourtellotte, “The Smallest Living Cells,” Sci. Am. 206(March 1962):117-126; H.J. Morowitz, Prog. Theoret. Biol. 1(1967):1.

73. A. R. Mushegian, E. V. Koonin, “A minimal gene set for cellular life derived by comparison of complete bacterial genomes,” Proc. Natl. Acad. Sci. (USA) 93(17 September 1996):10268-10273.

74. R. Himmelreich, H. Hilbert, H. Plagens, E. Pirkl, B.C. Li, R. Herrmann, “Complete sequence analysis of the genome of the bacterium Mycoplasma pneumoniae,” Nucleic Acids Res. 24(15 November 1996):4420-4449.

75. C. B. Williams, Patterns in the Balance of Nature and Related Problems in Quantitative Ecology, Academic Press, London, 1964.

76. C. W. Sabrosky, “How many insects are there?” in Insects, The Yearbook of Agriculture, U.S. Department of Agriculture, Washington, DC, 1952.

77. “Numbers of Insects (Species and Individuals),” Department of Entomology, National Museum of Natural History.

78. Nelson Thompson, “Biology/Entomology 173. Insect Physiology, Spring 1998, Lecture 17: Respiration,” 6 November 1997; “Some biological problems involving diffusion.”

79. J. Storrs Hall, personal communication, 6 May 2000.

80. U.S. Bureau of the Census, Statistical Abstract of the United States: 1996, 116th Edition, Washington, DC, October 1996.

81. “…there are dozens of HIV-like viruses in wild monkey populations, and if natural transfer of AIDS viruses from chimpanzees to monkeys has already occurred, there is no reason why it should not happen again.” Beatrice Hahn, Howard Hughes Medical Institute scientist, quoted in: Declan Butler, “Analysis of polio vaccine could end dispute over how AIDS originated,” Nature 404(2 March 2000):9.

82. “Recycled Tires for a Building System,” 1999; “Annual Form 10-KSB Report,” The Quantum Group, Inc., 31 December 1998; “Return Trip: How To Recycle the Family Car,” 1994.

83. “Solar Radiation Data Manual for Flat-Plate and Concentrating Collectors: 30-Year Average of Monthly Solar Radiation, 1961-1990, Spreadsheet Portable Data Files,” DOE Renewable Resource Data Center.

84. George M. Hidy, The Winds: The Origins and Behavior of Atmospheric Motion, D. Van Nostrand Company, Princeton, NJ, 1967.

85. Evan R.C. Reynolds, Frank B. Thompson, eds., Forests, Climate, and Hydrology: Regional Impacts, United Nations University Press, Tokyo, Japan, 1988; see: “Effect of surface cover on land surface processes.”

86. Map of roughness parameter due to vegetation in the U.K.;.

87. PSUBAMS Model, “Dual roughness regimes,” April 1997.

88. Horace Robert Byers, Synoptic and Aeronautical Meteorology, McGraw-Hill Book Company, New York, 1937.

89. Mindaugas Zickus, “Influence of Meteorological Parameters on the Urban Air Pollution and its Forecast: Section 2.6.4 Vertical temperature gradient,” Ph.D. Thesis, 1999.

90. Joseph Morgan, Introduction to University Physics, Volume One, Allyn and Bacon, Inc., Boston, MA, 1963.

91. Reporting on Climate Change: Understanding the Science.Chapter 3. Greenhouse Gases, Some Basics,” Environmental Health Center, National Safety Council, Washington, DC, November 1994, ISBN 0-87912-177-7.

92. Robert J. Bradbury, personal communication, 8 May 2000.

93. B. Lobitz, L. Beck, A. Huq, B. Wood, G. Fuchs, A.S.G. Faruque, R. Colwell, “Climate and infectious disease: Use of remote sensing for detection of Vibrio cholerae by indirect measurement,” Proc. Natl. Acad. Sci. (USA) 97(2000):1438-1443.