SYNERGISTIC SELECTION: Endnote Outtakes

CHAPTER 1: EXPLAINING COMPLEXITY

  1. Due to space limitations, the endnotes in the published volume were abbreviated. A version with more substantive material is provided here.
  2. Lamarck 1914/1809, p. 292.
  3. “From the earliest traceable cosmical changes down to the latest results of civilization,” Spencer wrote, “we shall find that the transformation of the homogeneous into the heterogeneous is that in which progress essentially consists” (Spencer 1892/1852, vol 1, p. 10).
  4. Darwin 1968/1859, p. 318.
  5. Ibid., p. 348.
  6. Indeed, Darwin rejected the very idea of sharp discontinuities in nature. In The Origin he emphasized what he called the “Law of Continuity,” and he repeatedly stressed the incremental nature of evolutionary change, which he termed “descent with modification.” (Darwin 1882/1859, p. 133).
  7. Lloyd Morgan 1923, 1926, 1933. For a history of emergence theory, see Blitz 1992.
  8. Haldane and Huxley (1927, pp. 234-235) viewed an increase in size and complexity over time as the very essence of what they called “biological progress” (a characterization that is, of course, widely rejected today) and pointed out, among other things, the importance of modular organization and a “division of labour”. I thank Egbert Leigh for providing this reference.
  9. See Hutchinson 1965. Bonner, in his 1988 book, The Evolution of Complexity by Means of Natural Selection, explored at length the thesis that increased size and complexity were interrelated, and that these were functionally advantageous developments that had been favored over time by natural selection. See also Bonner 2003.
  10. Margulis 1970, 1990, 1993; also, Margulis and Fester, 1991. The term “symbiogenesis” was actually coined by the Russian theorist, Konstantin Mereschkovsky 1909.
  11. Margulis 1998; Margulis and Sagan 2002. See also Carrapiço 2010; Kiers and West 2105. Also important was Egbert Leigh’s (1977, 1991, 2010a,b) work on “the common good” and on the factors allowing group selection to override within-group selection (see also Leigh 1983). As Kiers and West (2015) put it: “Symbiotic partnerships are a major source of evolutionary innovation.” The extensive work on the role of development, and especially modularity, has further illuminated the evolution of biological complexity. See especially Gould, 1977; Jablonka and Lamb, 1995, 2014; Raff, 1996; West-Eberhard, 2003; Kirschner and Gerhart, 2006; and Wagner, 2014.
  12. Holland 1998, p. 2.
  13. Ibid., p. 3.
  14. Kauffman 1995, pp. 8, 25.
  15. Kauffman 2000, p. 5. It should be noted, however, that, in a more recent 2008 book, Kauffman’s views have evolved; he now fully embraces the Darwinian paradigm. See also Schneider and Sagan 2005.
  16. Thus, Francis Heylighen and his colleagues (1999) claimed that evolution leads to the “spontaneous emergence” of systems with higher orders of complexity. Mark Buchanan (2000) discerned a “law of universality” in evolution – from our cosmic origins to economic societies – as a consequence of the phenomenon of “self-organized criticality” as proposed by Per Bak and his colleagues (Bak and Chen, 1991). Steve Grand (2001) viewed the emergence of networks in nature as a self-propelled, autocatalytic process. Albert-László Barabási (2002) invoked “far reaching natural laws” that, he believes, govern the emergence of networks. Niels Gregersen and his coauthors (2002) saw an “innate spontaneity” in the emergence of complexity. A similar argument can be found in the book by Daniel McShea and Robert Brandon (2010). They posit a “first law’ of biology — an inherent tendency in evolution for living systems to diversify and become more complex over time, unless constrained by natural selection. (See also the critical review by Bromham, 2011.) Biophysicist Harold Morowitz (2002) comes somewhat closer than most theorists of this school to a view that is compatible with the Darwinian paradigm. Recognizing that variability is inherent in the living world at every level, Morowitz posits that there are “pruning rules” that shape the forms that arise out of the many possibilities in evolution (p. 55). However, Morowitz did not specify what these pruning rules are (perhaps he meant natural selection) and finds himself in sympathy with the anthropologist/priest Teilhard de Chardin (and others) in believing that there is “something deeper” in the “orderly unfolding” of the universe. Also noteworthy is the nascent effort in what has been called “synthetic biology” to map and perhaps explain the major transitions in evolution in terms of laws derived from the theoretical work in robotics, artificial life and advanced simulation models. See especially Solé 2016a,b. See also Corning and Szathmáry 2015; Smart 2015.
  17. McShea (2015) aspires to find “some single principle or some small set of principles” that could explain the evolutionary trend toward greater complexity. Likewise, biologist Deborah Gordon (2007) laments: “Perhaps there can be a general theory of complex systems, but we don’t have one yet.” The recent theoretical work devoted to this subject is recounted in Corning 2013; also, Corning and Szathmáry 2015. It should also be noted that there is a long-running debate about how to define complexity. See especially Corning 2005, pp. 92-93; also, Corning and Szathmáry 2015; McShea 1996. For more on this issue, see Chapter Seven;
  18. Corning 1983.
  19. The biologist Steven Frank, in a 1997 journal article, made the same point: “Synergism creates associations between [gene] loci, and statistical association may have [selective] consequences similar to physical linkage.”
  20. A number of convergent views about the role of synergy in evolution will be quoted in various places throughout the book. They have been assembled here.*In a new edited volume devoted to Cooperation and Its Evolution, the co-editors, in their introduction, observe that there must be “higher profits” for cooperation to occur, and they see “this profit as the synergy of collective action” (Sterelny et al. 2013, p. 3.)*The biologist Steven Frank, in a 1997 journal article points out: “Synergism creates associations between [gene] loci, and statistical association may have [selective] consequences similar to physical linkage.”*Anthropologist Joe Henrich points to the rise of what he calls our “collective brains” — our shared cultural systems. “Our collective brains arise from [the] synergies created by the sharing of information among individuals.” He characterizes our evolution as “a broad and synergistic process” (Henrich 2016, pp. 212, 250).*The interplay between synergy, relatedness and reciprocity are highlighted in a recent article by Van Cleve and Akçay 2014. They conclude: “We suggest that measuring nonadditive [synergistic] interactions at both the payoff (fertility or survival) and fitness levels will prove important for explaining the importance of social behaviors…”*Biologist Klaus Jaffe, in a 2001 paper on the relative importance of haplo-diploidy, assortative mating and synergy in the evolution of social behaviors concludes that “economic considerations rather than genetic ones are critical in explaining the emergence and maintenance of sociality.”*Economists Samuel Bowles and Herbert Gintis (2011) observe in their book-length study of the subject that the key to cooperation is “a high ratio of benefits to costs…the net benefits.”*Pigliucci and Müller (2010) conclude in their edited volume, “[The] shift of emphasis from statistical correlation to ‘mechanistic’ causation [like synergies] arguably represents the most critical change in evolutionary theory today.”*In a comprehensive volume on The Ants (1990), and in a subsequent popularization, co-authors Bert Hölldobler and Edward O. Wilson conclude: “The amazing feats of the weaver ants and other highly evolved species comes not from the complex actions of separate colony members but from the concerted actions of many nestmates working together…The colony is the equivalent of the organism, the unit that must be examined in order to understand the biology of the colonial species.”*Biologists David Sloan Wilson and Edward O. Wilson (2008) stress the potential synergistic value of social information in a recent article on “Evolution ‘For the Good of the Group.’” Referring to a phenomenon they characterized as “the group mind,” Wilson and Wilson observed that “the collective benefits of making a wise decision can be great and the within group costs can be low.”*Biologist David Sumpter (2010), in his volume on Collective Animal Behavior, deploys the concept of synergy extensively as an explanatory concept (and model) Sumpter addresses collective behaviors across a broad range of species and behaviors, from migration to food acquisition, information sharing, decision making, collective defense, and risk sharing. He focuses especially on the relationship between mechanisms and their functions. Equally important, he utilizes his own and others’ game theory models to analyze in detail the relationship between costs and benefits, and the functional consequences of cooperation or defection.

    In contrast, consider the argument by Kenneth Weiss and Anne Buchanan in The Mermaid’s Tale (2009) that the fundamentally cooperative aspect of life is due to “slippage” in the competitive dynamic of natural selection – in other words, an under the radar phenomenon. As they put it: “The weak nature of most natural selection most of the time solves a major problem that was raised by the extensively cooperative nature of life, because weak selection tolerates the slippage that enables new kinds of cooperation to emerge” (p. 241). On the contrary, I see cooperation as entirely consistent with natural selection.

  21. Biologist Patrick Bateson (2013).
  22. A word is in order here about the relationship between synergy and the concept of “emergence”.   The two are emphatically not synonymous. Synergy refers to the functional results or outcomes produced by cooperative relationships and interactions of all kinds, effects that are not otherwise attainable.   Emergence is a term that has only very recently “re-emerged” in the scientific dialogue (as noted above, it was also popular in the 19th century), and it is often very unclear what it stands for. How do you know it when you see it? It has been used in various ways, often without being defined by the user, though it commonly seems to refer to the physical properties, or physical appearance of some object or phenomenon. (See the extended discussion of this issue in Corning 2005.)   My argument is that it is the functional consequences – the synergies – that have been the drivers behind the evolution of complexity over time. This is spelled out in much greater detail in the chapters ahead. For a lucid discussion of emergence, see Capra and Luisi 2014.
  23. For more on this theory, see Corning 1983, 2003, 2005, 2013.
  24. Maynard Smith and Szathmáry 1999, pp. 22-23. They wrote: “Co-operation will not evolve unless it pays. Two co-operating individuals must do better than they would if each acted on its own…Behavioural examples are easy to think of, but the principle is relevant at all levels…Peter Corning, in a book called The synergism hypothesis published in 1983, reviewed the role of synergy in social and biological evolution. We had not seen his book when we wrote The major transitions in evolution [1995], but we are happy to acknowledge that he foreshadowed this part of our argument, often using the same examples.” See also Corning and Szathmáry 2015. It should be noted that Klaus Jaffe (2001) also independently recognized the role of synergy in evolution. As he put it, “economic considerations rather than genetic ones are critical in explaining the emergence and maintenance of sociality.”
  25. Maynard Smith 1982a; also, 1983. In Aristotle’s words, “The whole is something over and above its parts, and not just the sum of them all…” Aristotle, 1961/ca. 350 B.C., Book H, 1045: 8-10.
  26. Plotkin 2010, p. 139.
  27. A word is in order here about how this theory relates to the theoretical enterprise known as “synergetics”. Two different theorists are associated with this term. One is the American engineer/inventor/polymath Buckminster Fuller, who coined the term in relation to his work on the design, behavior and transformation of physical systems. (See his book on Synergetics http://www.rwgrayprojects.com/synergetics/synergetics.html). Fuller called it “the geometry of thinking” and defined it as a system of measurement focused on the relationships between things. Following the ancient Greek mathematician, Pythagoras, before him, Fuller believed synergetics illuminated a “cosmic logic.” Fuller was also famous for inventing a multi-purpose geodesic dome that used tetrahedrons as structural elements. The other theorist associated with synergetics is the German physicist Hermann Haken, whose work was inspired by how lasers aggregate and empower light photons. Haken developed a rigorous approach to understanding the dynamics of self-organizing thermodynamic systems that he refers to as “the science of structure.” Particularly notable is his “order parameter” concept, which describes how organized physical systems may be governed. (Among his many books, see Haken 1983.) The major distinction in both versions of synergetics is that they are focused on the “how” questions related to the physical and biological realms – how do systems work and how do they change over time? The synergism hypothesis, in contrast, is focused on the “why” question. Why have living systems evolved and continue to evolve? To repeat, it is an economic theory that is located within a Darwinian selection paradigm.
  28. Darwin The Origin of Species, p. 444. http://darwin-online.org.uk/Variorum/1859/1859-490-c-1860.html (accessed 25 February, 2015).
  29. Quoted in Brandon 1969.

 CHAPTER 2: A NEW VIEW OF EVOLUTION

  1. Michod, 1999, p. xi.
  2. Nowak (with Highfield), 2011, p. xviii.
  3. See Corning 2003; also, Corning 1983, 2005, 2007a, 2013; Corning and Szathmary 2015.
  4. Aristotle, 1961/ca. 350 B.C., Book H, 1045:8-10. To repeat Aristotle’s original phrasing: “The whole is something over and above its parts, and not just the sum of them all.”
  5. The scientists Klaus Jaffe and Gerardo Febres (2016) have recently highlighted the thermodynamic aspect of synergy in the natural world. This is a useful approach. Living systems consume energy and display an increase in physical order (or “negentropy” in the jargon of the physicists). However, Jaffe and Febres err in seeking to define synergy strictly in thermodynamic terms; they equate the thermodynamic aspect of a living system with the physical/biological properties and their functional effects. To do so, they utilize the standard thermodynamic measures of Gibbs free energy and Claude Shannon’s information entropy – a purely statistical indicator related to the complexity of communications “messages”. It’s like defining an automobile in terms of its fuel consumption and the total number of its parts. The thermodynamics may be relevant, but it is the functional, “economic” payoffs that explain the presence of synergy in living systems, as we shall see. Jaffe and Febres also err in characterizing synergy narrowly in terms of the classic non-additive formulation – the whole is greater than the sum of its parts (1+1=3) – and they use water as an example. But this nicely illustrates the problem. A water molecule requires two hydrogen atoms and one very different, much heavier atom of oxygen coupled with a catalyst. It’s not 1+1, and the result is not “greater” quantitatively than the sum of its parts; it’s qualitatively different. “Synergy” is just a word. It can be defined in various ways. But one “unambiguous” definition is combined, or “cooperative” functional effects that are not otherwise attainable. This is at once more accurate and more inclusive.
  6. Allen 1983.
  7. Van Soest 1994; also, see: http://en.wikipedia.org/wiki/Cellulose (last modified 21 August 2014); see also: http://science.jrank.org/pages/1335/Cellulose-Cellulose-digestion.html#ixzz3CBTQMiuT (accessed 2 September 2014).
  8. Raven 1992.
  9. Spribille et al. 2016. Another example is Azolla, a small aquatic fern with some remarkable properties. It can double in size in two or three days; it can fix nitrogen directly from the air and therefore does not need to be rooted in the soil (it’s typically found floating in ponds and quiet streams, although it does have roots and can also thrive in wet soils); it serves as a “green manure” in many of the rice paddies of China and Asia, where it can boost yields by as much as 150 percent and eliminate the need for crop rotation; it benefits the climate by taking lots of CO2 out of the atmosphere; it can also provide feed for livestock, serve as a renewable biofuel, clean up waste water, or even deter mosquitos. (Under some circumstances, though, it can also become a noxious weed.) The secret of its success, it turns out, is that it is actually a three-way symbiotic relationship between the Azolla plants, a blue-green algae (cyanobacterium) called Anabaena, and various bacteria. Biologist Francisco Carrapiço (2010a), who has studied these plants extensively, characterizes the Azolla symbiosis as a superorganism; it’s an obligate partnership that can be traced back 80 million years. The key to it all, it seems, is the extraordinary dual talent of Anabaena. Sheltered inside a specialized cavity provided by the Azolla leaves, the Anabaena resemble strands of beads, with two distinct kinds of cells. One set produces energy through photosynthesis while the other (called heterocysts) use this energy to capture free nitrogen and convert it into ammonium as a feedstock for the Azolla plant. In return, the Azolla plant provides a protective environment and essential nutrients for the Anabaena, especially fixed carbon (sugars). (The role of the bacteria remains unclear.) It’s like adding a supercharger to the engine of your car. See also Adams et al. 2013; Carrapico 2010b; also http://theazollastory.com; http://waynesword.palomar.edu/plnov98.htm; http://theazollafoundation.org; https://en.wikipedia.org/wiki/Azolla (last modified 13 July 2017).
  10. See Corning 1996; also, Ridley 2001. Weiss and Buchanan, in their book length treatment of the cooperative aspect of living systems, speak of the “cooperative genome.” They argue that reproduction is fundamentally a team effort. “Biological traits represent cooperation on a grand scale” (p. 225).
  11. It should be emphasized that Synergistic Selection is a concept that is now being utilized with increasing frequency by evolutionary biologists, as reviewed in Corning and Szathmáry 2015. It should also be stressed that, like the concept of natural selection itself, Synergistic Selection is not viewed as some external selecting agency, or force. It is an “umbrella term” for an open-ended sub-category of causal influences in evolution. It highlights the distinctive properties of a class of influences that have shaped the evolution of cooperation and complexity over time.
  12. Darwin 1968/1859, p. 133.
  13. Ibid., p. 229.
  14. Ibid., p. 459. To be sure, Darwin also used a more nuanced phrasing in other places (see Chapter Three), and, in a later edition, he inserted a qualifier before the famous passage quoted above: “It may metaphorically be said…”
  15. Wilson 2013, pp. 50-51. Wilson is not alone in calling natural selection a “force”. See also Weiss and Buchanan 2009 (p. 226); also, Plotkin 2010. Others speak of “selection pressures” as if there were some external physical force influencing selection. See for example, Odling-Smee et al. 2003; Lieberman 2013.
  16. Among the nay sayers about Mayr’s concepts, see especially Laland et al. 2011, 2013; Calcott 2013a,b. However, I still find the proximate/ultimate distinction – properly interpreted — to be useful. The point is not to abandon the distinction but to refine and clarify what it should mean. Proximate causation refers to all of the various functional questions related to how a trait might have evolved, including its utility, how it develops during ontogeny, and the mechanics of how it works. Ultimate causation refers to the consequences of the trait — how all of these functional aspects may affect differential survival and reproduction (evolution) over time.   It should be emphasized that proximate/functional causation and ultimate “causation” (i.e., natural selection) are in reality deeply interpenetrated, and interactive. Laland and his colleagues rightly emphasize that very often there is “reciprocal causation.”   But the very idea of reciprocal causation also implies that there are in fact two distinct spheres of causation. I believe this formulation also accords with the four different questions about biological causation proposed by ethologist Nikolaas Tinbergen in the 1960s. His famous quartet includes the adaptive value of a trait, its ontogeny or development, how it works (or the mechanics), and how it evolved. See the review in Bateson and Laland 2013; also, the discussion in D.S. Wilson 2015.    As always, Aristotle seems to have had the first word on this subject, in the Physics and the Metaphysics, with his famous four-fold classification scheme: material causes, formal causes, efficient causes, and final causes. See https://en.wikipedia.org/wiki/Four_causes (last modified 24 November 2015).
  17. Mayr 1961, 2001. The term “selective retention” was coined by psychologist Donald T. Campbell (1974). A theoretical issue that is closely related to the proximate/ultimate distinction involves the relationship between the concept of rational “utility” in economic theory and the concept of “fitness’ in evolutionary biology. See especially the edited conference volume by Okasha and Binmore 2012. The basic problem is that both of these concepts are ill-defined and have been much-debated over the years.
  18. For an in-depth discussion of the role of behavior as a shaping influence in the evolutionary process, see Corning 2014.
  19. Laland et al., 2011, 2013.
  20. Grant and Grant 2014; also, Lack 1961/1947; Weiner 1994. We now know that variations in a single gene are responsible for producing these functional differences. The sequencing was reported by Lamichaney et al. 2015. Recent work suggests also that these birds are not actually true finches and may not all be distinct species. Other recent work reminds us that the shapes of bird beaks are also greatly influenced by nondietary physiological factors. See Bright et al. 2016.
  21. Warren 2010.
  22. Corning 1983; also 2003, 2005, 2007a, 2013.
  23. Maynard Smith and Szathmáry 1995, 1999. As noted in the Preface, the term “synergistic selection” was actually coined (and modeled) by Maynard Smith in 1982a; also, Maynard Smith 1983. It has been used with increasing frequency in evolutionary biology in recent years. See the review in Corning and Szathmáry 2015. It should be noted that an earlier book by Leo Buss (1987) on the evolution of complexity also invoked synergy, but his primary focus was on overcoming the problem of conflicts between different levels of biological organization. He did not advance the concept of synergy as a primary causal agency. See also the discussion of Buss’s work in Plotkin (2010), and the independent work in this area by Jaffe 2001, 2016.
  24. See especially Noble 2013; also, Corning 2014.
  25. Hamilton 1964a,b; Maynard Smith 1964.
  26. Maynard Smith (personal communication). Maynard Smith asserted that Haldane later changed the number to 10, in order to show a profit. Thus, Haldane deserves credit for the idea behind kin selection. His informal math presaged Hamilton’s work by many years.   As he wrote in his pioneering textbook The Causes of Evolution (Haldane 1932, p. 131): “Insofar as it makes for the survival of one’s descendants and near relatives, altruistic behavior is a kind of Darwinian fitness and may be expected to spread as a result of natural selection.”
  27. There has recently been a highly technical (and rancorous) debate about what has been called a “stretched version” of Hamilton’s original rule, where inclusive fitness is applied to various contexts beyond those involving genealogically related actors to encompass any shared genes, however derived, or even any social influence that affects fitness. See Leigh 2010b; Queller 2011; Marshall 2015; Jaffe 2016. For critiques, see Nowak et al., 2010; Allen et al. 2013; Birch 2014; Birch and Okasha 2015; Corning and Szathmáry 2015 D.S. Wilson 2015; also, the extensive comments in Abbot et al. 2011.
  28. Maynard Smith 1982b; also, Binmore 2005.
  29. It seems that others are now converging on this insight. In a new edited volume devoted to Cooperation and Its Evolution, the co-editors, in their introduction, observe that there must be “higher profits” for cooperation to occur, and they see “this profit as the synergy of collective action” (Sterelny et al. 2013, p. 3.) The interplay between synergy, relatedness and reciprocity are highlighted in a recent article by Van Cleve and Akçay 2014. They conclude: “We suggest that measuring nonadditive [synergistic] interactions at both the payoff (fertility or survival) and fitness levels will prove important for explaining the importance of social behaviors…” Another example, is the biologist David Sumpter (2010) in his book Collective Animal Behavior, where deploys the concept of synergy extensively as an explanatory concept (and model) and analyzes collective behaviors across a broad range of species and behaviors, from migration to food acquisition, information sharing, decision making, collective defense, and risk sharing. He focuses especially on the relationship between mechanisms and their functions. Equally important, he utilizes his own and others’ game theory models to analyze in detail the relationship between costs and benefits, and the functional consequences of cooperation or defection. (For a number of other examples, see the outtake at my website for Chapter 1. Endnote 20.)
  30. A vexed debate about the status and utility of kin selection theory and various broader interpretations of “inclusive fitness” has been raging within evolutionary biology for the past several years. See especially Nowak et al. 2010; Abbot et al. 2011; Bourke 2011; Allen et al. 2013; Birch 2014; Birch and Okasha 2015; also, Corning 2013; Corning and Szathmáry 2015. Particularly notable is the attempt by Klaus Jaffe (2016) to meld inclusive fitness theory and the economic factors associated with the Synergism Hypothesis.   He emphasizes the potential role of “assortation” – various influences that might bias the propensity to cooperate. Jaffe (2010) has also sought to quantify the potential energy consumption efficiencies in socially organized species.
  31. Further discussion and innumerable examples can be found in Corning 1983, 2003, 2005, 2007a 2013; also, Corning and Szathmáry 2015. A thoughtful analysis of synergism by the mathematical biologist David Sumpter can be found in his 2010 book Collective Animal Behavior, p. 236ff. Also noteworthy is the model developed by Cornforth et al. (2012) for synergistic benefit functions in microbes that produce public goods molecules that are shared with others. They found that the synergies typically exhibit a sigmoidal pattern. They are very much dependent on group size, with diminishing returns beyond an optimum number. Likewise, Taylor and Maciejewski (2012) developed an inclusive fitness model showing that frequency dependence plays a critical role in producing synergistic fitness effects.
  32. The term was inspired “music minus one” — a popular set of recordings (typically a string quartet or quintet) with one instrument omitted, so that the listener can fill in with their own accompaniment. “Thought experiments” are used quite frequently by scientists as a way of testing ideas. As biologist Richard Dawkins (1982, p. 4) points out: “Thought experiments are not supposed to be realistic. They are supposed to clarify our thinking about reality.”
  33. Peace and Grubb 1982.

 CHAPTER 3: HOW COOPERATION TRUMPS COMPETITION

  1. From Huxley’s notorious Romanes lecture, “Evolution and Ethics,” in 1893. Reprinted in Nitecki and Nitecki eds., 1993.
  2. Williams 1993.
  3. Dawkins 1989/1976, p. 2.
  4. From a letter to J.D. Hooker, 13 July 1856. http://www.darwinproject.ac.uk/letter/entry-1924 (accessed 17 February, 2015).
  5. Darwin 1882/1859 (Sixth Edition). Interestingly, in The Descent of Man (1874/1871) Darwin’s emphasis shifted. Here he used the term “social” more than 100 times, the term “mutual” 32 times, “co-operation” four times, and “competition” only 11 times.
  6. Spencer, http://en.wikipedia.org/wiki/Survival_of_the_fittest (last modified 25 February, 2015).
  7. As the nineteenth century steel magnate (and later philanthropist) Andrew Carnegie assured us in an inflammatory essay known as the “Gospel of Wealth”: “While the law [of competition] may be sometimes hard for the individual, it is best for the race, because it ensures the survival of the fittest in every department. We accept and welcome, therefore…great inequality of environment, the concentration of business, industrial and commercial, in the hands of the few, and the law of competition between these, as being not only beneficial, but essential for the future progress of the race” (Carnegie 1992/1889).
  8. Malthus 1798, Ch 2, 18. “Population, when unchecked…increases in geometrical ratio…[while] the means of subsistence, under circumstances the most favorable, could not possibly increase faster than in an arithmetic ratio.” It has been said that Thomas Carlyle’s characterization of economics as the “dismal science” was inspired by Malthus’s pessimistic essay.
  9. Darwin 1968/1859, p. 459.
  10. For a detailed discussion of this history, see Corning 1983. A similar world view can be found in Western, capitalist economics. For instance, the economist Jack Hirshleifer, in a classic 1978 article in The American Economic Review, asserted that “competition is the all-pervasive law of natural economy interactions.” However, he assures us that “political economy” in human societies acts to constrain the war of each against all with third-party enforcement of contracts, so that competition can serve the beneficial ends of stimulating progress and “harmonizing” the market place (compare Adam Smith’s “hidden hand” metaphor).
  11. Leigh 2010a.
  12. Bowles and Gintis 2011, p. 197.
  13. For short summaries of this subject see https://en.wikipedia.org/wiki/Competition_(biology) (last modified 1 June 2015); also, https://en.wikipedia.org/wiki/Competition (last modified 11 June 2015); and https://en.wikipedia.org/wiki/Intraspecific_competition (last modified 9 April 2015). Our perspective on the shaping role of competition also continues to shift. Recent doubts about its role in producing species diversity are discussed in Singer 2014.
  14. Indeed, much competition is an indirect result of the capacity for living organisms to expand their numbers over time and exploit the available space and resources. These days many theorists see direct competition as less significant as a driver of evolution than ecological and other external influences. See for example Sahney et al. 2010.
  15. Schopf et al, 2015. The authors of this paper describe it as “extreme evolutionary stasis.”
  16. See especially Carrapiço 2010; Pereira et al. 2012.
  17. Darwin 1859/1968, p. 459.
  18. Ibid., p. 116.
  19. For recent examples of this constricted paradigm, see Sachs et al. 2004; Lehmann and Keller 2006; West et al. 2011. Indeed, in a full-throated defense of “traditional” neo-Darwinian inclusive fitness theory, James Marshall (2015, p. 26) concludes that “The occurrence of cooperation in the natural world can thus be explained by standard natural selection theory acting on individuals’ direct fitness.” As noted earlier, the many exceptions challenge the rule – from non-kin without green beards to horizontal gene transfers in bacteria.
  20. Maynard Smith 1982b, 1998.
  21. A brief history of this paradigm can be found in Binmore 2005; also, Nowak (with Highflield) 2011. Nowak provides a detailed discussion of the many different “solutions” to the Prisoner’s Dilemma. In retrospect, the assumptions/specifications underlying the Dilemma seem to be highly contrived. The original “story” behind the Prisoner’s Dilemma involves two gangsters and a District Attorney who knows they are guilty of a serious crime but needs a confession from either one of them in order to convict them. So he offers each of them separately (and unknown to each other) the following options:• If you confess but your partner does not, you can go free.
    • If you fail to confess but your partner does, you will get the maximum sentence.
    • If you both confess, you will get lesser sentences.
    • If neither of you confesses, you will be set up and convicted on a lesser charge.Got that? It’s no wonder that most social scientists could not see how this model was broadly applicable to economic and social life. See also https://en.wikipedia.org/wiki/Prisoner%27s_dilemma (last modified 19 February 2016).
  22. Maynard Smith and Szathmáry 1995, p. 261.
  23. Less well known is the work on a more positive alternative model known as the stag hunt. The basic idea is that if two hunters collaborate to hunt for a stag they can do better than if each one hunts for hare on his own. See Skyrms 2004.
  24. Binmore 2005, p. 63.
  25. Ibid.
  26. Hunt 1998.
  27. Nowak 2006; also, Nowak (with Highfield) 2011.
  28. The term “primordial pizza” has also been used to characterize one of the alternative theories about the origin of life, namely, that it originated on land, perhaps in a layered form.
  29. Darwin 1874/1871. See also Smaldino 2014.
  30. See Crow 1979. Weiss and Buchanan (2009) use the metaphor of a lock and key, where a single defect in the key can render it inoperative.
  31. Clutton-Brock and Parker 1995, p. 209.
  32. Ratnieks and Visscher 1989.
  33. Sherman, et al. 1991, 1992.
  34. de Waal 1982, 1996.
  35. See especially Gintis 2000a,b; Gintis et al. 2003; Fehr and Gächter 2000a,b; Falk et al. 2001; Sethi and Somanathan 2001; Fehr et al. 2002; Fletcher and Zwick 2004; Bowles and Gintis 2011.
  36. See also the discussion in Bowles and Gintis 2011; also, Sterelny et al. 2013.
  37. Koestler 1967.
  38. A wide-ranging – and very readable – discussion of the role of cooperation in evolution, with a special focus on the molecular and cellular level in living systems, can be found in Weiss and Buchanan (2009). They note, for example, that almost every trait embodied in the genome of a complex organism involves a system of genes linked by a network of communications “signals” (pp. 223-225).

CHAPTER 4: EVOLUTION AS A ‘COMBINATION OF LABOR’

  1.  Smith 1964/1776, Book 1, Ch. 1, p.7. See also https://books.google.com/books?id=rBiqT86BGQEC&printsec=frontcover#v=onepage&q&f=false  Smith was also the first economist to point out that overseas trade can expand the potential market for a division of labor, foretelling one of the most significant advantages that England enjoyed during the industrial revolution (see Chapter Nine).
  2. Stigler 1951. The article focused on what Stigler called Adam Smith’s “theorem” that the benefits of a division of labor depend on “the extent of the market” – how many customers are available to purchase the product, so that the costs can be recovered and the benefits of increased production do not go to waste. He also noted that Smith’s theorem was not a complete explanation for the division of labor. Many other factors may be involved. We will return to this issue.
  3. Durkheim 1997/1893. Quoted in “Division of Labour” http://en.wikipedia.org/wiki/Division_of_labour (last modified 29 October 2014). See also http://sites.middlebury.edu/individualandthesociety/files/2010/09/division-of-labor.pdf Actually, Herbert Spencer, considered by many of his contemporaries to be the preeminent thinker of the nineteenth century, foreshadowed and may even have inspired Durkheim’s argument. Indeed, Durkehim’s classic contains 43 references to Spencer, far more than any other author. See Corning 1982. In his essay on “The Development Hypothesis” (1892/1852), Spencer equated “progressive” evolution with increasing “functional differentiation” and “functional integration.”
  4. Bonner 2003.
  5. Anderson and Franks 2001. See also Franks 1989.
  6. Simpson 2012.
  7. Ratcliff et al. 2012. McShea and Brandon (2010) have elevated this (presumed) tendency into an evolutionary “law”. The trouble is that it’s a law that can also go into reverse, with many documented cases of simplification – or even extinction. I suspect what is really at work here is that this “diversifying tendency” is simply another dimension of the fundamental variability that exists in the natural world – as Darwin himself highlighted. Sometimes there can be “hotspots” of variation. It’s more likely that functional synergies have been the drivers of this trend; diversifying variations with adaptive properties are favored by natural selection. Luis Zaman and his associates have found evidence that even the antagonistic interactions associated with co-evolution (say between predators and their prey) may favor increased complexity. See Zaman et al. 2014.
  8. Hölldobler and Wilson 2009, p. 84.
  9. E.O. Wilson 2013, p. 121.
  10. Plato 1946/380 B.C., Book 2, 369a,b, 370b,c, 372c.
  11. Smith 1964/1776.
  12. The following is synthesized from Bergquist 1978; George and George 1979; Ricketts et al. 1985; and Curtis and Barnes 1989.
  13. Smith 1964/1776, Book 4, Chap. 2.
  14. A reviewer of this book in manuscript pointed out that sponge cells are actually totipotent, meaning that adjacent cells could revert to an amoeboid state and then migrate to replace the missing cells. As an example of functional differentiation, however, sponges still provide a valid example.
  15. See Corning 2003.
  16. The French writer/philosopher Voltaire is frequently quoted on this point: “God is always on the side of the big battalions.” However, this is actually a famous misquote; it was taken out of context. What Voltaire wrote was that “It is said that…” but then he went on to disagree, writing that, on the contrary, God is on the side of the best shooters. https://en.wikiquote.org/wiki/Voltaire#Misattributed (Last modified 27 March 2015).
  17. Bonner 1988. See also Shapiro 1988.
  18. Dugatkin 1999, pp. 14-15.
  19. It should be noted that biologists Carl Anderson and Nigel Franks, in their article on “Teams in Animal Societies” (2001), used the term as a synonym for the division of labor. I prefer to distinguish between the two concepts to highlight the distinction between a task divided and roles that are combined to produce an emergent whole.
  20. Margulis and Sagan 2002; also, Mayr 1974. See also http://en.wikipedia.org/wiki/Mixotricha_paradoxa (last modified 18 July 2014).
  21. Isack, and Reyer 1989. This classic study was recently reinforced and amplified by a study among the Yao people in Mozambique. See Spottiswoode et al. 2016.
  22. Allee 1951/1938.
  23. Maddox 1990; Hardison 1999.
  24. El Naggar et al. 2010.
  25. Le Maho 1977. Of course, heat-sharing has also been practiced by humans for many centuries. It’s even mentioned in the Old Testament.
  26. Gould and Gould 1995.
  27. Von Wagner 1954.
  28. Earthworms dramatically alter the soils they inhabit in a multi-generational process. It’s believed that they were originally aquatic creatures and that they have made many adaptations, including major changes to their environments, in order to survive in a terrestrial habitat. See Odling-Smee et al. 2003, especially pp. 11-12, 374-376; also, Laland et al. 1999.
  29. Wright and Jones 2006.
  30. The following data are from the U.S. Department of Agriculture. See https://www.agclassroom.org/gan/timeline/farm_tech.htm (accessed 7 December 2014).
  31. From “A Four Year Plan for England,” BBC Broadcast March 21, 1943. See http://www.ibiblio.org/pha/policy/1943/1943-03-21a.html (accessed 7 December 2014).
  32. For further exploration of the concept of cybernetic “control information,” see Corning and Kline 1998a,b; Corning 2005, 2007b. The term “cybernetics” refers to the science of communications and control in goal-oriented systems. A key concept is “feedback” – the capacity to monitor and adjust the behavior of the system in order to attain, or maintain, its desired goal-state. A classic example is a household thermostat. Perhaps it is unintended, but many theorists these days use the term “feedback” in inappropriate ways that may obscure and even misrepresent the underlying causal dynamics in the natural world. Thus, biologist Conrad Waddington used the term to characterize a “circularity” in the relationship between an organism and its environment. Christopher Wills, in The Runaway Brain, speaks of a brain-body-environment “feedback loop.” James Lovelock characterized the Earth and its biota as being participants in a self-regulating, homeostatic “superorganism” that is maintained by “automatic feedback.” Geneticist David Thaler, one of the leading advocates for the notion of an “intelligent” genome, argues that there is, as he puts it, “feedback between the generators of genetic diversity [mutations, transpositions, etc.] and the environment that selects among variants.” Meanwhile, at the other end of the evolutionary scale, economist Brian Arthur has developed what he calls a “positive feedback” model of economic evolution.However, feedback is in fact a technical term in cybernetics and the information sciences. A cybernetic system is, by definition, a purposeful, goal-oriented system that is controlled by information flows. Feedback, sensu stricto, refers to information that is used by the system to monitor its behavior and make adjustments as necessary to achieve its goals or maintain a pre-specified state (homeostasis). Positive feedback will tend to encourage more of the same behavior, while negative feedback will discourage, modify or terminate the action. The classic example is a household thermostat, which senses room temperatures and turns the furnace on or off accordingly. To use the systems scientist William T. Powers’s formulation, feedback “signals” are compared to the internal “reference signals,” and it is the relationship between the two signals that determines what the behavior of the system will be.Any usage of the term “feedback” that departs from this goal-oriented, information-driven model is at best metaphorical and at worst misleading. Though perhaps inadvertent, it amounts to sneaking a hidden teleology into the natural world. Sometimes the term misrepresents a pattern of interactive, reciprocal causation. Lovelock’s “Gaia” is a case in point. Where is the “reference signal” in Lovelock’s superorganism? At other times, the feedback processes are inappropriately enlarged to include the inanimate world. When an organism produces changes in the environment, those changes may produce feedback (in the strict sense) to the organism that will influence its behavior over time. But there is no “circularity” involved; it is a one-way street. The environment does not react to feedback, only to the physical actions that produce environmental changes. A reciprocal feedback “cycle” can only occur when two different cybernetic systems respond to each other.Thus, to cite one example, when one of Lamarck’s giraffes “selected” acacia leaves and initiated a new feeding pattern for that species, the direct cause was the immediate nutritional rewards — the “reinforcements”. There was (presumably) also positive feedback in the strict sense between the giraffe’s digestive system and its brain, which increased the probability that the behavior would be repeated. But the feedback was a secondary, indirect cause of the behavior. If the new behavior contributed to the differential survival of longer necked giraffes, feedback played only a supporting role. It was the nutritional benefits that mattered. A lucid introduction to the science of cybernetics and its important contributions can be found in Capra and Luisi 2014.
  33. Darwin 1965/1873.
  34. See Capra and Luisi 2014. See also http://en.wikipedia.org/wiki/Quorum_sensing (last modified 9 November 2014).
  35. See especially Bonabeau et al. 1999; Eberhart et al. 2001; Couzin 2007; Woolley et al. 2010; Henrich 2016.
  36. Seeley 2010.
  37. D.S. Wilson and E.O. Wilson 2008. See also Wilson and Wilson 2007; D.S. Wilson 2015; also, Franks 1989; Couzin 2007.
  38. See S.J. Gould 2002. See also the lucid discussion in Capra and Luisi 2014.
  39. What if the ship that was promoted as “unsinkable” had been provided with a sufficient number of life boats? What if the ship-builder had used a higher grade of steel, or better rivets? What if the President of the White Star Line had not been aboard the ship on its maiden voyage and had not been pressuring the captain for a speed record? What if any of the numerous telegraph warnings about icebergs had been heeded? What if the lookouts had been provided with binoculars, an incredible oversight? What if the duty officer on the ship’s bridge that night had not reversed the engines, which slowed the vessel as he was attempting an emergency turn? What if a nearby merchant ship had not ignored the many distress signals and had instead come to pick up survivors? What if…? The tragedy of the Titanic was in fact a colossal synergistic effect. If just one of those “what ifs” had been different, this terrible disaster might not have occurred. Among the many sources on this infamous historic event, perhaps the most comprehensive and up-to-date is the volume by Lynch (1992).
  40. For example, a very different form of synergy was identified by biologist Mary Jane West-Eberhard (2003, pp. 591-592). She noted that there can be times when two distinct selection processes may reinforce each other. Thus, behaviors associated with reproductive competition (sexual selection) might also be selectively favored for ecological success. The so-called “hover wasps” provide an illustration. Their helicopter-like behaviors are used for both competitive displays and for foraging.
  41. Spencer 1897. Spencer’s definition was strictly functional. He spoke only of an analogy between structures and functions in organisms and societies (see vol. 1, p. 592).
  42. Wheeler 1928. Biologist Warder C. Allee also embraced the term in his 1931 book, Animal Aggregations.
  43. Williams 1966, p. 220, p. 8.
  44. E.O. Wilson 1975, p. 383.
  45. D.S. Wilson and Sober 1989. A defense of the superorganism concept in my 1983 book, The Synergism Hypothesis, was not widely noticed. For a more detailed history and discussion of the concept, see the chapter on “Synergy and the Evolution of Superorganisms” in my 2005 book, Holistic Darwinism.
  46. Wilson and Sober 1989, p. 339. That same year, honeybee specialist Thomas Seeley (1989) published an article on “The Honeybee Colony as a Superorganism.”
  47. Hölldobler and Wilson 1994, p. 107.
  48. Ibid., p. 122. Emboldened by Hölldobler and Wilson’s unapologetic usage of the term, biologists Robin Moritz and Edward Southwick subsequently published a book-length monograph on Bees as Superorganisms: An Evolutionary Reality (1992).
  49. The following discussion is derived from the definitive rendering in Hölldobler and Wilson 2009. We are deeply indebted to them for a lifetime of work, and publications, on insect societies. See also Hölldobler and Wilson 1990.
  50. Currie et al. 1999; Currie 2001.
  51. See the discussion of this subject in Seeley 1995.
  52. In fact, sponges display several different forms of synergy — functional complementarities, a division/combination of labor, synergies of scale, and even structural (gestalt) synergies. For instance, the shape of the (classic) sponge, with its exit opening located at the top, utilizes physics to help pull water through its cavity, rather like the updraft in a chimney. As a result, a sponge can typically process a quantity of water equal to its own volume in less than ten seconds.

CHAPTER 5. A TALE OF TWO THEORIES

  1.  Lamarck 1914/1809. Actually, his formal name was Jean-Baptiste Pierre Antoine de Monet. He later inherited the title “Chevalier de Lamarck” from his father.
  2. See Cannon, 1955, 1959; also, Corsi 1988. Pietro Corsi’s meticulous reconstruction of Lamarck’s life and work reveals a scientist of diverse interests (for instance, he was also a pioneer in meteorology) who became an early advocate for successive “transmutations” of life and made the evolution of complexity a major theme in his work. However, his views, and writings, were also enmeshed in personal rivalries with other prominent scientists and were deeply contested. On one major score, in retrospect, he was mistaken. He denied the extinction of species – as a rule – in favor of a theory of transformation over time. In fact, he was only half right. See also Burkhardt 1959.
  3. Gould 1999a,b. Gould quotes the following key passage from Lamarck’s 1820 book: “Let us consider the most influential cause for everything done by nature, the only cause that can lead to an understanding of everything that nature produces…. This cause resides in the power that circumstances have to modify all operations of nature, to force nature to change continually the laws that she would have followed without [the intervention of] these circumstances, and to determine the character of each of her products. The extreme diversity of nature’s productions must also be attributed to this cause.”
  4. Cited in Noble 2013. Noble states: In his introduction to Harvard’s re-publication in 1964 of The Origin of Species, Ernst Mayr wrote (pp. xxv–xxvi) “Curiously few evolutionists have noted that, in addition to natural selection, Darwin admits use and disuse as an important evolutionary mechanism. In this he is perfectly clear. For instance… on page 137 he says that the reduced size of the eyes in moles and other burrowing mammals is ‘probably due to gradual reduction from disuse, but aided perhaps by natural selection’. In the case of cave animals, when speaking of the loss of eyes he says, ‘I attribute their loss wholly to disuse’ (p. 137). On page 455 he begins unequivocally, ‘At whatever period of life disuse or selection reduces an organ…’ The importance he gives to use or disuse is indicated by the frequency with which he invokes this agent of evolution in the Origin. I find references on pages 11, 43, 134, 135, 136, 137, 447, 454, 455, 472, 479, and 480.”
  5. Darwin 1968/1859, p. 459.
  6. Quoted in Noble 2013.
  7. Though Lamarck was often accused, even by Darwin, of proposing that new habits arise as a result of spontaneous “volition” or “desire”, he said no such thing. This misapprehension was apparently the result of a mistranslation of the French word besoin; the word volonté was used by Lamarck only in relation to some “higher” animals. In fact, Lamarck viewed behavioral changes, by and large, as a matter of challenge and response, of externally stimulated creativity rather than a spontaneous impulse. See also Cannon 1955, 1959.
  8. Lamarck 1914/1809, p. 114.
  9. Darwin, 1968/1859, p. 215.
  10. Ibid.
  11. Gissis and Jablonka, eds. 2011, pp. xi-xii. See also Corsi 1988.
  12. Lamarck 1914/1809, p. 122.
  13. Although the term Organic Selection was coined by child psychologist James Mark Baldwin, it should properly be called the Lloyd Morgan/Baldwin Effect. The basic claim of Organic Selection theory was that, in the course of evolution, the first step in producing systematic biological changes might well be a change in behavior, especially among the more ‘plastic’ species. When an animal is in some way able to modify its behavior so that it can ‘select’ a new habitat, or a new mode of adaptation, after a number of generations the change might precipitate congenital changes ‘in the same direction’ that would undergird and perfect the new adaptation. This would occur not because the changes are somehow stamped into the offspring but because the new environment creates a ‘screen’ that would selectively favor individuals with the relevant ‘somatic variations’ (Baldwin, 1896a,b,c). Baldwin also made ambitious claims for the role of Organic Selection. However, his writings (and his terminology) were anticipated by the British comparative psychologist Conwy Lloyd Morgan in two of his books Animal Life and Intelligence (1991), and Habit and Instinct (1896a), and in his 1896 lecture at the New York Academy of Science, which Baldwin attended. Following Lloyd Morgan’s verbiage, Baldwin in his writings spoke of behavioral ‘accommodations’ that could keep an animal alive and so allow its offspring to ‘accumulate’ biological ‘variations’ determining evolution in ‘subsequent generations.’ Indeed, in his 1902 book Baldwin stressed the possibility of varying relationships between learned accommodations and ‘coincident [congenital] variations’, as did Lloyd Morgan (1896a,b). Biologist Patrick Bateson (2013) reports that both Lloyd Morgan and Baldwin were anticipated by Douglas Spalding (1873).
  14. Paleontologist George Gaylord Simpson (1953) was responsible for coining this term in a now classic paper. However, Simpson chose to define the Baldwin Effect much more narrowly than Baldwin and the other Organic Selection theorists had done. He claimed that it pertained only to parallel ‘genetic changes with similar results’. Thus, to use one of Simpson’s examples, if a callous acquired by a human hand through use made an appearance in the hand of a newborn and was favorably selected, this would be an instance of the Baldwin Effect. No wonder Simpson concluded that the Baldwin effect was unproven and, in any case, of no great importance. However, subsequent Baldwin champions bypassed Simpson’s definition in favor of more expansive interpretations (e.g., West-Eberhard 2003, pp. 24–25; 151–153), just as later evolutionists (Simpson among them) came to see behavior as a major change agent in evolution without reference to Baldwin or Lloyd Morgan, much less Lamarck.
  15. See Waddington 1962. This history is reviewed at length in Corning 1983; also Corning 2014.
  16. Weismann 1904.
  17. See https://en.wikipedia.org/wiki/Hugo_de_Vries (last modified 20 February 2016).
  18. Huxley 1942.
  19. D.T. Campbell 1974. To be sure, the Modern Synthesis also succeeded in bringing such life science fields as ecology, paleontology, botany, systematics, and cytology under a gradualist, selection-oriented theoretical umbrella, and it inspired biologists to address such important theoretical questions as just how speciation actually occurs in the natural world and how biological diversity is maintained over time? It was also realized that evolution is a process that occurs in interbreeding populations of organisms, not just isolated individuals.
  20. Crick 1970.
  21. J.H. Campbell 1994, p. 86. As West-Eberhard (2003, p. 143) notes: “If pressed to name the mechanism behind the origin of new traits…most biologists would answer genetic mutations.”
  22. Dawkins 1989/1976, p. ix. This famous prefatory remark was only one of many provocative statements in the book. Here are two other examples: “I shall argue that a predominant quality to be expected in a successful gene is ruthless selfishness” (p.2); “Be warned that if you wish, as I do, to build a society in which individuals cooperate generously…you can expect little help from biological nature…we are born selfish” (p.3). In retrospect, Dawkins’ inspired metaphor was deeply subversive. Dawkins himself backpedaled toward the end of his famous book (where he explained that it is really cooperative “teams of genes” that evolve together). In later writings he averred that the “gene’s eye view” is only one way of viewing the natural world. Nevertheless, the selfish gene perspective greatly influenced an entire generation of theorists and, alas, misled the general public as well (especially anti-Darwinian Creationists and political conservatives). Evolutionary theorist Eva Jablonka (2004) points out that Dawkins’ focus on the “replicators”, especially in a later book on The Extended Phenotype (1982), did help to purge some sloppy thinking from evolutionary theory, but at a high cost.
  23. See Corning 2005; Noble 2006; Capra and Luisi 2014.
  24. This quote is from an earlier article by Shapiro 2009.
  25. Shapiro 2011, p xvii. See also Jablonka and Lamb 1995, 2014; Jablonka et al. 1998; also, Noble 2006, 2013; Pigliucci and Müller 2010; Bateson and Gluckman 2011; Gissis and Jablonka 2011; Jablonka 2013. Shapiro summed up these changes in a 2009 article: “Since the elaboration of the central dogma of molecular biology, our understanding of cell function and genome action has benefited from many radical discoveries. The discoveries relate to interactive multimolecular execution of cell processes, the modular organization of macromolecules and genomes, the hierarchical operation of cellular control regimes, and the realization that genetic change fundamentally results from DNA biochemistry. These discoveries contradict atomistic pre-DNA ideas of genome organization and violate the central dogma at multiple points. In place of the earlier mechanistic understanding of genomics, molecular biology has led us to an informatic perspective on the role of the genome. The informatic viewpoint points towards the development of novel concepts about cellular cognition, molecular representations of physiological states, genome system architecture, and the algorithmic nature of genome expression and genome restructuring in evolution.”
  26. Shapiro 2011, p. 2.
  27. Noble 2013. See also Noble 2012; Jablonka and Raz 2009. For an update to Rcihard Lewontin’s classic analysis of the various “units” of natural selection, see Godfrey-Smith 2015.
  28. Jablonka et al. 1998; also Jablonka and Lamb 2014. Jablonka and Raz (2009) reported that there were then more than 100 documented examples of epigenetic inheritance in 42 different species. It is now apparent that various forms of epigenetic inheritance are ubiquitous.
  29. Jablonka 2013. Jablonka and Lamb (2014) cite some 102 studies encompassing 42 different species.
  30. West-Eberhard 2003, p. 20. She argues (p. 144) that “the most important initiator of evolutionary novelties is environmental induction.” What is sometimes referred to as the “plasticity first” hypothesis was recently evaluated in depth by biologists Nicholis Levis and David Pfennig (2016). They cited some 21 documented cases in various species, ranging from plants to insects, fishes, amphibians, reptiles and birds.
  31. For examples, see Plotkin 1988; Weber and Depew 2003. For a detailed review, see Corning 2014.
  32. Bateson 2004, 2005, 2013.
  33. In the index to their book on Animal Traditions, Eytan Avital and Eva Jablonka (2000) list well over 200 different species. See also Tagkopoulos et al. 2008; also, Breed and Moore 2010.
  34. Heinrich 1995. Heinrich’s (1999) book, The Mind of the Raven, provides extensive evidence for the mental abilities of these remarkable birds.
  35. Gilroy and Trewavas 2001. See also the comprehensive treatment by Trewavas (2014).
  36. E. O. Wilson 1975, p.172; also, Beck 1980; McGrew 1992. Chimpanzees are particularly impressive tool users. They frequently use saplings as whips and clubs; they throw sticks, stones and clumps of vegetation with a clearly hostile intent (but rather poor aim); they insert small sticks, twigs and grasses into ant and termite holes to “fish” for booty; they use sticks as pry bars, hammers, olfactory aids (to sniff out the contents of enclosed spaces), and even as toothpicks; they also use stones as anvils and hammers (for breaking open the proverbial tough nuts); and they use leaves for various purposes — as sponges (to obtain and hold drinking water), as umbrellas (large banana leaves are very effective), and for wiping themselves in various ways (including chimpanzee equivalents of toilet paper and “sanitary napkins”). See also Wrangham et al. 1994. Not only are chimpanzees proficient as tool-users but they can also make tools. They break off small tree branches and strip them to fabricate ant “wands”; they use their bodies for leverage when they break down larger sticks to make hammers; they work leaves into sponges; and they carefully select stones of the right size and shape for the job at hand and will then carry them to their work-sites.
  37. Weigl and Hanson 1980.
  38. Palameta and LeFebvre 1985.
  39. John et al. 1968.
  40. Cited in Byrne 1995, p.58.
  41. Byrne and Whiten 1988; Whiten and Byrne 1997; also, Gibson and Ingold 1993.
  42. Mayr 1960, pp. 371, 377-378.
  43. The experiment is described in Jonathan Weiner’s wonderful book, The Beak of the Finch (1994), which recounts in fascinating detail the long-range study by Peter and Rosemary Grant (and their colleagues and their daughter) in the Galápagos Islands. See pp. 182-184. See also Grant 1986, 1991; Grant and Grant 1979, 1989, 1993, 2002, 2014.
  44. Adam Weiss (2015), in a critical paper on “Lamarckian Illusions,” disagrees. However, his definition of Lamarckism is very constricted and thus distorted, focusing on a single example.
  45. E.g., D.S. Wilson 1997a,b; Sober and Wilson 1998.
  46. E.g., Pigliucci and Müller 2010; also Mesoudi et al. 2013; Jablonka 2013; Jablonka and Lamb 1995, 2014; Jablonka et al. 1998.
  47. Noble 2013. See also Noble et al. 2014. Noble and his co-authors highlight three specific contributions: a reintroduction of “function” into the causal mix, the addition of higher order organizing principles, and accounting for organismal systems properties.
  48. For a discussion of group-level selection, see Farine et al. 2015. Jablonka and Lamb (2014) point out that there are three distinct processes at work in evolution: variation, transmission, and differential selection. All three of these processes are greatly influenced by developmental, phenotypic, and environmental influences. However, the old saying that the truth often runs well in reverse can also be applied to evolution. Sometimes a genetic change can be a direct, deterministic cause of natural selection. Consider the “raspberry gene” in the fruit fly Drosophila melanogaster, a recessive mutation that is responsible for producing both raspberry-colored eyes and a sharp decline in reproductive success when it is present at both of its two gene loci (the so-called homozygous condition). In a classic set of experiments that pitted the homozygotes against heterozygotes (where only one of the two loci had the mutant gene), as well as individuals without the mutation, the selection coefficient against the homozygote was a powerful .5. After only ten generations, the frequency of the homozygotes in the experimental population was reduced from 50 percent to about 10 percent. The specific mechanism responsible for this decline was suggested in subsequent experiments with another mutant form. It was found that, when the raspberry gene mutation was expressed (in the homozygote), it was responsible for subtle changes in the Drosophila’s stereotyped courtship behavior, which in turn drastically reduced its mating success. See Merrell 1949, 1953; also, Bastock 1956.
  49. Noble et al. 2014.
  50. Pigliucci and Müller 2010. See also Laland et al. 2015.
  51. s Pigliucci and Müller conclude in their edited volume, “[The] shift of emphasis from statistical correlation to mechanistic causation arguably represents the most critical change in evolutionary theory today” Ibid., p. 12.  They stress that the concept of “variation” in evolution has been greatly expanded beyond genetic mutations and that natural selection is no longer treated as strictly an external agency but includes internal “generative” changes as well. I go one step further. In fact, natural selection is located in the relationships and interactions between an organism and its environments(s) and the consequences for differential survival and reproduction — as discussed in Chapter Two.
  52. Watson and Szathmáry 2016. See also the comments and the authors’ responses in Trends in Ecology and Evolution, 2016, 31(12): 891-896.
  53. Kauffman 2008. An in-depth treatment of the role of “agency” – purposeful, goal-related behavior – in evolution can be found in Walsh 2015.
  54. In a detailed examination of several current textbook-length treatments of evolutionary theory, philosopher of science Alan Love (2010) found that they are in fact quite inclusive and synthetic and have moved well beyond a gene-centered approach.

CHAPTER 6: THE MAJOR TRANSITIONS IN EVOLUTION

  1. Leigh 1995.
  2. Leigh 1977, 1983, 1991.
  3. Maynard Smith and Szathmáry 1999, pp. 22-23.
  4. See especially Corning 2014. The science of semiotics is also relevant here. See especially Fernández 2016.
  5. See for example, Kiers and West 2015.
  6. For a critique of traditional information theory and a functional alternative, see Corning and Kline, 1998a,b; also, Corning, 2007b. For a perspective on the synergistic nature of human language and its integration with other cognitive abilities, see Szathmáry and Számadó, 2008.
  7. For an up-to-date overview, see the fiftieth anniversary issue of the Journal of Theoretical Biology in 2012 with some 20 articles devoted to various aspects of cooperation. For a critical review of the work on sanctions, see Frederickson, 2013.
  8. Some of the many important contributions to this theme include Jablonka 1994; Jablonka and Lamb 2005; Queller 1997, 2000; Bonner 1999, 2006; Michod 1999, 2005, 2007; Michod and Herron 2006; Kauffman 2000; Frank 1995; Okasha 2005, 2006; Nowak et al. 2010; Bourke 2011; Van Cleve and Akçay 2014; Kiers and West 2015.
  9. Corning and Szathmáry 2015; Szathmáry 2015
  10. See especially Calcott and Sterelny 2011.
  11. Corning and Szathmáry 2015.   Here is what we said: “Our view is that there is no one correct way to measure complexity, or the major trends and transitions in evolution; they can be defined in different ways for different purposes. As a way of characterizing the broad evolutionary trend culminating (temporally at least) in humankind, we suggest two alternative methodologies (at least in theory). One is structural: A synthetic complexity scale compounded from the number of levels of organization (inclusive of social organization), the number of distinct ‘parts’, the number of different kinds of parts, and the number of interconnections among the parts. The other method is functional: A complexity scale derived from the number of functionally discrete ‘tasks’ in the division/combination of labor at all levels, coupled with the quantity of ‘control information’ that is generated and utilized by the system. (Control information is defined as the capacity to control ‘the capacity to do work’ in a cybernetic process; it is equivalent to the amount of thermodynamic work that a system can perform.)”
  12. McShea and Simpson 2011.
  13. For the record, I differ slightly from Maynard Smith and Szathmáry, who use a dualistic definition that includes new forms of information (see below). I also agree with Ereshefsky and Pedroso (2015) that a new level of reproductive specialization need not be involved. This is especially relevant for human evolution. (See Chapters Seven and Eight.)
  14. The term has many fathers. Among others, see especially Buss 1987; Ghiselin 1997; Michod 2007; also, West et al. 2015; Ereshefsky and Pedroso 2015.
  15. See Heckman et al. 2001; also, Field et al. 2015.
  16. Schrödinger 1944.
  17. For a critique of misuses of the second law, see Corning and Kline 1998a.
  18. Schrödinger’s characterization is seductive. It has been cited on innumerable occasions over the years. But there are problems. In the first place, negative entropy reduces the complexities of living systems to a monolithic thermodynamic process and conflates thermodynamic order with functional “organization” — purposive designs for adaptation in a great variety of specific environments, including a number of different energy regimes and levels of organization. Metabolism is only one aspect of the many-sided problem of earning a living in the natural world. In effect, Schrödinger truncated the challenges associated with physical organization, self-development, self-maintenance and reproduction into a single parameter, thus distorting the very nature of the evolutionary process. Schrödinger has also been excused by his many fans over the years for saying some things that, in his own word, were “absurd.” Organisms are able to circumvent the inherent tendency of real world processes to dissipate energy and become ever more entropic over time, Schrödinger explained, by “extracting order” from their environments. He also spoke of organisms “sucking orderliness” from their environments, and he identified as their most distinctive feature their “well-ordered” state. Although it is often said that organisms feed upon energy, Schrödinger declared, this is “absurd….what an organism feeds upon is negative entropy” (Schrödinger 1945, p. 72). Schrödinger’s vision thus misrepresents the energetics of living systems, which have developed ingenious and highly efficient (i.e., profitable) mechanisms for capturing or harvesting available energy in various forms and then using it for various purposes, from doing useful work to building biomass. Contrary to Schrödinger’s assertion, it is more accurate to say that organisms feed on available energy and create thermodynamic, structural and functional order than to say that they feed on order (cf., Morowitz 1968, p. 19; Perutz 1987; also Corning and Kline 1998a,b).
  19. Corning and Kline 1998a,b; Corning 2007b. See also the brief history and discussion of the science of cybernetics in Capra and Luisi 2014.
  20. Maturana and Varela 1980/1973. As Capra and Luisi (2014, p. 132) put it, “Life is a factory that makes itself from within.”
  21. For an insightful discussion of autopoiesis, see Capra and Luisi 2014. For a brief review of how the concept is used and misused, see http://en.wikipedia.org/wiki/Autopoiesis (last updated 30 March 2015).
  22. See Kauffman 2008.   Needless to say, the issue of how to define life has become a theoretical quagmire. It is definitely a matter of opinion. See especially the pioneering theoretical work by the late Robert Rosen (1970, 1985, 1991). A recent attempt to do, so using E-coli bacteria as a “model”, can be found at http://www.ncbi.nlm.nih.gov/pubmed/25796394
  23. Gánti 2003/1971; 2003. A more recent, independently developed formulation of this idea is the so-called CMP model, which stands for Container, Metabolism, and Program. See Bedau 2102.
  24. See especially Morowitz 1992. The theory developed by David Deamer, Harold Morowitz and others postulates a key role for what they call “amphiphiles” — elongated fatty molecules that are like lipids in modern cells. Amphiphiles, which evidently were present in the prebiotic environment, have the unique ability to align themselves with respect to water and can self-assemble into “vesicles.” Certain “recourse” would nonetheless have been available to be selectively transported into the vesicle to catalyze and sustain the beginnings of life, according to this theory. See especially Deamer 1978; Deamer and J. Oro 1980; Morowitz 1978, 1981, 1992; Morowitz et al., 1987.
  25. Kauffman 1971, 1986; Hordijk et al., 2011; Vasas et al., 2102. See also Segbroeck et al., 2009; Bedau 2012. Another autocatalytic model, called the “hypercycle”, was proposed by Manfred Eigen in the early 1970s. See especially Eigen and Schuster 1977, 1979. Maynard Smith and Szathmáry later observed in The Major Transitions that this model was unstable due to its susceptibility to “cheating”. Szathmáry (2015) points out that many theorists confuse the hypercycle and chemoton models. They are not the same thing.
  26. See Corning 2003, 2007, 2014; see also Mayr 1974. Also noteworthy is the fact that Jablonka (2002) develops a functional definition of information, similar to my cybernetic “control information”, which implicitly recognizes its teleonomic foundation. Capra and Luisi (2014) formulate the same dynamic in terms of the concepts of autopoiesis and “cognition”.
  27. The much-debated challenge of how to define life has been inconclusive. I take the contrarian view that life is not a discreet, static set of traits or attributes but a moving target — a dynamic process that has been evolving (and changing) for some 3.8 billion years – so far.
  28. An excellent brief summary of this concept and the research behind it can be found at: http://en.wikipedia.org/wiki/RNA_world (last updated 27 May 2015).
  29. Indeed, the vital role of DNA in biosynthesis is possible only because of the highly coordinated role that three distinct forms of RNA play in “transcribing” and utilizing the nucleotide bases (a combination of labor). The so-called messenger RNA makes a copy of the relevant DNA sequence from a given chromosome in the cell nucleus and transports it to a “manufacturing” site, a ribosome in the surrounding cytoplasm (or cell body); transfer RNA brings the appropriate amino acids from elsewhere in the cytoplasm to the ribosome and links them to the RNA “template”; and ribosomal RNA plays a key role in fabricating the protein by lining up the amino acids in the proper order. Without the actions of RNA, DNA would be impotent.   Biologist James Darnell and his co-authors, in their well-known textbook Molecular Cell Biology, conclude that the development of several distinct functions by RNA was probably “the molecular key to the origin of life” (Darnell et al., 1990, p.88).
  30. Szathmáry 2015. Others vehemently disagree (e.g. Capra and Luisi 2014). Some of the controversy involves different definitions of what the term means. For an up to date review and assessment, see Kun 2015.
  31. Mazur 2014.
  32. Miller 1953; Miller and Urey 1959.
  33. Cairns-Smith 2009/1968; 1987/1982.
  34. However, the idea was recently revived with a proposal, backed by some experimental work, that ancient clay based “hydrogels” could have served as an alternative to lipids as containment vessels. See Yang et al. 2013; also, http://www.natureworldnews.com/articles/4784/20131106/life-evolved-clay-researchers-find.htm
  35. Wickramsinghe 1974.
  36. See especially Callahan et al., 2011; also, the review at http://en.wikipedia.org/wiki/Panspermia (last updated 27 May 2015).   See also Crick 1981; Hoyle 1983.
  37. Wächtershäuser 1988, 1990, 1992.
  38. See Tunnicliffe 1991.
  39. See especially Gold 1992, 1999; Martin and Russell 2003.
  40. Wächtershäuser 2006, 2007; also, Wächtershäuser and Adams 1998.
  41. Russell 2006; Martin and Russell 2003, 2007.
  42. Lane 2009.
  43. Martin and Russell 2003, 2007; Martin et al. 2008; Lane 2009; also, Koonin and Martin 2005.
  44. Nick Lane (2009), the well-known biochemist and author at University College, London, sees alkaline vents as the ideal environment for the full sequence of evolutionary steps that was required to move from primitive RNA molecules to DNA and ultimately complete replicating (and evolving cells). Lane concludes: “What began as simple affiliations between molecules became, in this world of naturally proliferating cells, selection for the ability to reproduce the contents of whole cells. It became selection for self-sufficiency, and ultimately for autonomous existence” p.55.
  45. Patel et al., 2015; also, Powner et al., 2009; Lilley and Sutherland 2011.
  46. Letter to J.D. Hooker, 1 February 1871. See Francis Darwin 1887, Vol. 3, p. 18.
  47. Wolfenden et al., 2015; Carter and Wolfenden 2015.
  48. Carter quoted in http://phys.org/news/2015-06-evidence-emerges-life.html. (accessed 6 June 2015). David Deamer (2011) disagrees with the hot pond scenario, citing experiments in which clays and metal ions blocked chemical interactions. On the otherhand, William Martin and his co-workers, in a pioneering DNA analysis (Weiss et al. 2016), believe they can trace the earliest common ancestor of contemporary living organisms to deep sea vents some 4 billion years ago. More research is obviously needed.
  49. With a somewhat different model in mind, Capra and Luisi (2014, p. 142) agree: “Life is the synergy of the three domains.” They emphasize the role of “cognition”, an alternate way of characterizing the capacity for teleonomic/cybernetic interactions with the environment. Also noteworthy is the novel thesis advanced by Eugene Koonin (2016), leader of the genomics research group at the National Institutes of Health. He proposes that the parasitic “pressure” of “selfish” genetic elements – plasmids, transposons, and viruses – played a major role in the evolution of complex living organisms. It was a defensive response.
  50. Equally important, even at the most basic biochemical level, life is fundamentally a synergistic effect.   The well-known origins-of-life theorist Mark Bedau (2012) calls it “cooperative chemistry.” He notes that the three key functional elements – what he refers to as Containment, Metabolism and Program (or CMP) – were very likely to have been chemically integrated as they evolved. As he puts it, each part was “created and sustained by the whole system itself.”
  51. The following discussion is derived from Corning 2003, 2005; see also the excellent reviews and references in http://en.wikipedia.org/wiki/Bacteria (last updated 2 May 2015), and http://en.wikipedia.org/wiki/Archaea (last updated 15 May 2015).
  52. Christina Warinner, Prehistoric Human Biology as Inferred from Dental Calculus, in the 2016 CARTA public symposium on Ancient DNA and Human Evolution, https://carta.anthropogeny.org/node/156412 (accessed 29 May 2016). Sears 2005.
  53. See Margulis and Sagan 2002.
  54. Price 1991.
  55. Gorbach 1990; O’Hara and Shanahan 2006; Zoetendal et al., 2006.
  56. See the review in Crespi 2001. A useful theoretical analysis can be found in West et al., 2006.
  57. The assumption that ancient stromatolytes were formed by the same processes that are observed in modern counterparts has been challenged by a mathematical analysis that suggests a purely physical origin — a “spontaneous” sprouting and aggregation process under the right conditions. See Grotzinger and Rothman 1996; also, Pendrick 1997. Moreover, many ancient stromatolytes do not contain evidence of microbial activity. But others do. And there is no doubt about contemporary stromatolytes. See also Allison et al., 2001.
  58. Bonner 1988. See also the discussions of collective behavior in bacteria by Shapiro 1988; also, Shapiro and Dworkin 1997.
  59. http://en.wikipedia.org/wiki/Bacteria (last updated 2 May 2015).
  60. The following discussion is an updated version of the material presented in Corning 2003, pp. 57-59.
  61. The case for this is argued in some detail in Lane and Martin 2010. As they put it: “Virtually every ‘eukaryotic’ trait is also found in prokaryotes, including nucleus-like structures, recombination, linear chromosomes, internal membranes, multiple replicons, giant size, extreme polyploidy, dynamic cytoskeleton, predation, parasitism, introns and exons, intercellular signaling (quorum sensing), endocytosis-like processes, and even endosymbionts. Bacteria made a start up virtually every avenue of eukaryotic complexity, but then stopped short. Why?” They calculate that eukaryotes can support perhaps 200,000 times as many genes. Equally important, they possess orders of magnitude more energy per gene, which determines how much they can do in building proteins and doing other work. See also Lane 2014.
  62. The combination of mitochondria and chloroplasts in plants enables them to generate some 15-20 times as much available energy (net of entropy) as do prokaryotes. See Margulis and Sagan 1995.
  63. Biologist Thomas Cavalier-Smith has pointed out that phagocytosis in turn depended on the prior development of internal membranes and the cytoskeleton. This allowed for the loss of a rigid outer cell wall in favor of a flexible external membrane that could create pockets. Cavalier-Smith 1981, 1991, 2013.
  64. Mereschkovsky, K.C., 1909.
  65. Quoted in Khakhina 1992a. See also Khakhina 1992b; Margulis and McMenamin eds. 1993.
  66. Wallin 1927, p. 8.
  67. See especially Margulis 1970, 1981.
  68. See especially de Duve 1996.
  69. Reviewed in Margulis and McMenamin 1990; also, Margulis and Dolan 1999 and references therein.
  70. See Margulis et al., 2000.
  71. Reviewed in Szathmáry 2015. See also Cavalier-Smith 2009, 2013.
  72. Martin and Müller 1998. See also Vogel 1998.
  73. Martin and Mentel 2010. Another, purely speculative, mutualistic scenario is based on the functional relationship between photosynthesis, glycolysis and respiration. It could perhaps be called the “waste hypothesis.” Early eukaryotes that were able to engage in photosynthesis also generated large quantities of a waste-product, oxygen, which is toxic to organic material at high concentrations. At least some of these eukaryotes also utilized glycolysis, which (as mentioned earlier) produces quantities of a second waste product, pyruvic acid. Then along came a purple bacterium (an ancestral mitochondrion) that had developed a talent for using waste oxygen and pyruvic acid, via respiration, to meet its energy needs. The partnership was mutualistic from the start, and in time it became mutually obligatory. It is significant that some ancient “archaezoa” that are not related to our own direct ancestors apparently engaged in photosynthesis but did not have mitochondria. In other words, the two organisms may have evolved independently and their subsequent symbiosis (a functional complementarity) empowered them to go on to bigger and better synergies.
  74. Lane 2009, pp. 111-112. See also Lane 2014.
  75. Queller 1997, 2000; see also Queller 2004.
  76. Bell 1985; Michod 1999, 2005, 2007, 2011; Michod and Herron 2006.
  77. Bonner 2006; also, Bonner 1999. Of course, large size may be a severe disadvantage if the food supply is short, and bacteria have managed to do alright despite their small size.
  78. Halstead 1988.
  79.  http://en.wikipedia.org/wiki/List_of_muscles_of_the_human_body (last updated 5 June 2015).
  80.  http://en.wikipedia.org/wiki/List_of_organs_of_the_human_body (last updated 21 May 2015).
  81. This supplants an earlier estimate of 51 areas in the so-called Brodman Scheme. See Glasser et al. 2016. https://en.wikipedia.org/wiki/Human_brain#Functional_divisions (last updated 8 June 2015); also   http://en.wikipedia.org/wiki/List_of_animals_by_number_of_neurons (last updated 9 June 2015).
  82. Szathmáry 2015.
  83. Jablonka and Lamb 2006; see also Ginsburg and Jablonka 2010.
  84. Tagkopoulos and Tavazoie 2008; Mitchell et al., 2009.
  85. Szathmáry defends the inclusion of humankind on the list of transitions on the basis of what he calls our four distinctive traits: (1) language, (2) human cooperation, (3) eusociality (meaning a non-reproductive caste, such as grandparents), and (4) cultural group selection. Furthermore, the “rudiments” of human language are viewed as having coevolved in Homo erectus with a new form of cooperative behaviors, namely “confrontational scavenging” (competing with other species for access to carrion). See Szathmáry 2015; see also Bickerton and Szathmáry 2011. The problem with using this scenario as a justification for labeling human evolution as a major transition is that all four of these signature traits can be found in some form in other species as well (assuming the “rudiments” of language mean intentional signaling). Even the scavenging scenario has been challenged. In fact, I will argue in the next two chapters that cooperative behaviors and social communications predated Homo erectus by at least one million years, to the time when our remote australopithecine ancestors made the transition from the safety of the trees to earning a living in a far more dangerous terrestrial habitat, a radically new niche.
  86. Economist Paul Seabright, in his 2004 book, A Company of Strangers, explores in depth the phenomenon of widespread cooperation among non-kin as a hallmark of our species, with exchange and markets) as a key institution and trust as a key lubricant.
  87. The meaning of this term is much debated. Some theorists attribute it only to humankind (my preference) while others equate it with the concept of a superorganism and apply it to any species that has achieved a complex social organization, such as leaf cutter ants. See especially Turchin 2013, 2016; Gowdy and Krall 2014, 2015.
  88. D.S. Wilson 2015, p. 49.

 CHAPTER 7: THE SELF-MADE MAN: AUSTRALOPITHECINE ENTREPRENEURS

  1. Wade 2006, p. 1. See also Bokma et al. 2012.
  2. Paleoecology has developed an array of sophisticated new tools in recent years, including stable isotope analyses, feces analyses, and the analysis of ancient micro-climates, and a great deal of ingenious and painstaking work has been done to extend the little evidence we do have.
  3. Corning 1983; also, 2003, 2005.
  4. Boyd and Richerson 1985, 2005; Richerson and Boyd 2005; Kingdon 1993; Henrich 2016. See also Fisher and Ridley 2013.
  5. Gene-culture coevolution theory, also known as “dual inheritance theory,” is attributed to three major contributions in the 1980s by Lumsden and Wilson 1981; Cavalli-Sforza and Feldman 1981; and Boyd and Richerson 1985. For an overview, see https://en.wikipedia.org/wiki/Dual_inheritance_theory (last modified 8 June 2105). See also Henrich (2016). He also prefers the “culture-gene” version of the term.
  6. For the important work on “niche construction theory” and “ecosystem engineering,” see Odling-Smee et al. 1996, 2003, 2013; Laland et al. 1999; O’Brien and Laland 2012.
  7. Lewin 1997.
  8. Darwin 1874/1871, p.148.
  9. Perhaps most famous was Raymond Dart (1953, 1959), who portrayed our ancestors as blood-thirsty “killer apes.” Our ancestors, Dart wrote, “seized living quarries by violence, battered them to death, tore apart their broken bodies, dismembered them limb from limb, slaking their ravenous thirst with the hot blood of victims and greedily devoured livid writhing flesh.” Quoted in Stanford 1999, p. 107. See Dart 1953, 1959. Dart’s vision of humankind as the descendants of killer apes may have been overdrawn, but the basic idea resonated with a generation that had just survived the first of two savagely destructive world wars.   Particularly noteworthy is Alexander’s “autocatalytic arms race” hypothesis. For a recent update, see Flinn et al. 2005. For an in-depth theoretical treatment of the role of collective violence in the natural world generally, including hominins, see Corning 2007a.
  10. Washburn and Lancaster 1968, reprinted in Ciochon and Fleagle 1993, p. 219. The criticism leveled at this scenario over the years provides another illustration of the tendency to misinterpret what a theorist has said. Washburn and Lancaster did not argue that the hunting mode of life initiated the process of human evolution but rather that it emerged much later, with Homo erectus.
  11. Sally Linton’s seminal paper appeared in 1971, but the debate was ignited by Adrienne Zihlman and Nancy Tanner in their provocative 1978 paper.
  12. Zihlman and Tanner also asserted that the hunting of large animals with crude weapons at close range is a dangerous activity that most likely did not emerge until much later on, when hominin groups expanded into temperate climate zones with more seasonal food resources. In their female gathering scenario, the males were also viewed as being rather superfluous; they had no significant economic role. Zihlman and Tanner speculated that the males might also have been subject to female sexual selection for sociability and docility!
  13. Isaac 1978. See also Isaac 1981, 1983.
  14. Binford 1987. Binford 1987; also, Potts and Shipman 1981; Potts 1984, 1988; Shipman 1983, 1986.
  15. Lovejoy 1981; see also Lovejoy 2009.
  16. Subsequent studies by other anthropologists provided some support for Lovejoy’s thesis. Alan Mann (1981), for instance, highlighted the potential for more systematic exploitation of high-protein food resources (eggs, insects, invertebrates, reptiles, birds, small animals) as a way to augment and enrich the hominin diet. In other words, a dichotomous choice between big game hunting and plant foods is too simplistic; it is more likely that these creatures were opportunistic omnivores. Duane Quiatt and John Kelso, 1985, stressed the potential economic benefits of nuclear family units. Jane Lancaster, 1975, 1978, emphasized the mutual advantages that could be achieved by an economic division of labor. Paul Turke 1984 pondered the significance of hidden ovulation and the synchronizing of menstrual cycles in human females. And Milford Wolpoff 1999a, pp 210-211, citing fossil evidence of a high mortality rate and early deaths among our remote hominin ancestors, suggested that a high frequency of orphaned juveniles would have favored the selection of more elaborate kinship bonds and social support networks, possibly involving male as well as female relatives.   For an alternative theory of allomothering, see Hrdy 2000, 2009; Hawkes 2003; see also Kramer and Otárola-Castillo 2015. See Chapter Eight.
  17. Wrangham et al. 1999; Wrangham 2009; Wrangham and Carmody 2010; Gowlett and Wrangham 2013; also, Pennisi 1999.
  18. See the various contributions in Vrba, et al., eds. 1995; also Kingdon 1993; Foley 1994,1995; Allen, et al. 1999; McManus, et al. 1999; Taylor 1999; Alley 2000; Keeling and Whorf 2000; Stanley 2000; Richerson et al. 2001; deMenocal 2004, 2011; Richerson and Boyd 2013; Potts and Faith 2015.
  19. Vrba et al., eds. 1995 proposed that climate changes were the necessary and sufficient cause for the emergence of Homo erectus. According to their “pulse hypothesis,” climate changes are the primary drivers for various waves of speciation and extinction. See also Stanley 1992.
  20. Bateson and Hinde 1976; Bateson 1988. See also Foley 1995.
  21. Washburn and Lancaster 1968; Blumenshine 1987; Shipman and Walker 1989; Wrangham and Peterson 1996; Stanford 1999; Wolpoff 1999a; Wood and Collard 1999.
  22. See Blumenshine 1987; Blumenshine et al. 1994.
  23. Among others, see Domínguez-Rodrigo and Barba 2006, 2007; Blumenshine et al. 2007; also, Bunn 2001; O’Connell et al. 2002; Bickerton and Szathmáry 2011; Számadó 2011.
  24. See the various contributions in Vrba, et al. eds. 1995; also, Kingdon 1993; Foley 1994,1995; Potts 1996, 1998a,b, 2012; Allen, et al. 1999; McManus, et al. 1999; Taylor 1999; Alley 2000; Keeling and Whorf 2000; Stanley 2000; Richerson 2001; deMenocal 2004, 2011.
  25. Potts 1996, 1998a,b, 2012; Potts and Faith 2015. The most important ecological influence on the course of human evolution, it now seems evident, was a major long-term change along the Eastern edge of the subtropical forest belt that spans the mid-section of Africa, in what is now (principally) Tanzania, Kenya and Ethiopia. Here a combination of factors gradually reshaped the landscape. One was a rifting process, due to the movement of the huge tectonic plates that run the length of the continent. The other was a shift in global climate patterns. Over time these influences greatly altered the terrain and the ecology of East Africa. Immense tracts of dense tropical rain forest areas — safe havens with plentiful resources for arboreal primates — were gradually converted into a more broken, “mosaic” pattern with many “patches” of woodland, an abundance of lakes and streams, large areas of marshland and more open savanna areas. These environmental changes created both a challenge and an opportunity — a gradual reduction in the traditional primate homeland, a growing array of more seasonal plant materials and, equally important, the appearance of many animals that were more suited to woodland and open grassland areas. Many of these areas ultimately came to be populated with large herds of herbivores (some 20 different species), along with a formidable number of carnivores — saber toothed cats, lions, leopards, hyenas, wild dogs, cheetahs and more.
  26. See Kerr 1996; Hill 1987; also, Foley 1994; Potts 2012.
  27. White 1995, p. 378. White added: “It seems likely that the end of this time window was marked by technological innovation involved with the production of the Acheulean [tool] industry.”
  28. deMenocal 2004, 2011. At the risk of offending the professional paleoanthropologists, I have taken the liberty of lumping together the current (shifting) list of about a dozen early hominin species as being categorically distinct from the dozen or so species of Homo, ranging from Homo rudolfensis to Homo neanderthalensis for the sake of clarity in explaining the major functional/adaptive changes over time to this audience. Some of the differences between “early” and later species will be discussed later on.
  29.  Homo erectus is the name that has traditionally been used to identify the entire genus, or grade. More recent African fossil finds that are significantly different from the original Eurasian fossils have inspired the use of Homo ergaster for what are most likely to have been our direct ancestors. Here, for convenience sake, I will use the original generic term. For a brief description of Ardipithecus ramidus see White et al. 2009.
  30. See especially Anderson 1986; Cheney and Wrangham 1987; Dunbar 1988; Cowlishaw (1994); Iwamoto, et al. 1996; Wrangham and Peterson 1996.
  31. The evidence of predation against early hominins is discussed by Brain (1981) (1985), where he details his studies at Swartkrans. There is also the hominin fossil known as OH-7, the 12-13 year-old with a gnawed foot. See also the review of predation on primates by Isbell (1995), and Foley (1995) and the citations in Endnote 37 above. J. Lee-Thorp, et al. (2000) single out leopards (Felis), sabre-toothed cats (Merantereon), and hyenas (Crocuta), as likely to have been systematic predators on primates and australopithecines of that era. There is some concrete evidence that saber-toothed cats preyed on australopithecines. https://en.wikipedia.org/wiki/Dinofelis (last updated 9 June 2015).
  32. This assumption is derived from our common ancestry with chimpanzees, another primate that follows a pattern of male-based, multi-male/multi female groups. See especially the analyses in Lee 1994; also Wrangham 1987; Silk 2011; Schultz et al. 2011. Foley and Gamble (2009) point out that all of the major social developments in hominins over time can also be found (though not all together) in other social animals. One reviewer of the manuscript for this book questioned why male based groups would have been important. True, baboon troops are female based, but they also depend (structurally and functionally) on a precarious, often highly contentious male dominance hierarchy. A savanna baboon troop is also very small and is not subject to the many predators that challenged our early hominin ancestors. Both the context and the limited evidence we have suggest (to me, at least) that, to be successful over the long term (and outstrip baboons!), australopithecines would have benefitted from having larger groups, more intense male cooperation, and a more egalitarian, consensual decision making pattern in order to execute a more complex foraging and food-procurement strategy, as well as group migration and reproductive cooperation. Whether the required level of male-male cooperation was achieved by “family selection” or group selection, or both is certainly arguable.
  33. Wolpoff 1999a, pp. 217-219.
  34. First proposed by Morton 1927. See also Kurtén 1984;Coppens and B. Senut 1991; Wolpoff 1999a. Köhler and Moya-Solá (1997) report a recent find of a nine million year old fossil specimen, which they suggest may have used bipedalism for ground foraging in their sheltered Sardinian environment. For a different point of view, see Isbell and Young 1996.
  35. Thorpe et al. 2001.
  36. Kingdon 2003.
  37. Kingdon believes that intensive terrestrial squat foraging would likely have been highly productive, judging by today’s very similar lowland East African forest floors, which contain a rich variety of fruits, nuts, insects, snails, small invertebrates and more than 100 species of small mammals and reptiles. Efficient squat feeding on the rear haunches would have encouraged the necessary anatomical changes for bipedalism and would have freed up the hands to become effective foraging tools.
  38. Wrangham 1987.
  39. Diamond 1992.
  40. See Richmond et al. 2001; and Kivell and Schmitt 2009.
  41. Sockol et al. 2007.
  42. Zollikofer et al. 2005. Indeed, the discovery some years ago of several 5.2-5.8 million-year-old Ardipithecus fossils in the Middle Awash area of Ethiopia indicates that the shift toward Australopithecus began far earlier than had previously been supposed. Some features, from their bipedal gait to their teeth, clearly presage Australopithecus afarensis (Lucy and her cousins) some two million years later. See also Lieberman 2013.
  43. For our purpose, the assortment of very early hominin fossils that have been given different names (Ardipithecus, Australopithecus and Paranthropus, etc.) can be lumped together under the general heading of australopithecine “grade”.   This will allow us to sidestep the controversies over labels, categories, and sequences and concentrate on the most general and important features.
  44. Silk 2011.
  45. Many other theorists over the years have also endorsed the group-defense model, including George Schaller, Alexander Kortlandt, John Pfeiffer, Carel van Schaik, Richard Alexander, Richard Wrangham, Joseph Henrich, and others. For instance, Van Schaik (1983), in an in-depth study of non-human diurnal primates, concluded that security was the most compelling explanation for group living.
  46. These “smart monkeys,” as baboon expert Shirley Strum (1987) calls them, are ubiquitous throughout Africa and the Near-East and are formidable competitors. See also Kummer 1968.
  47. See Alemsegad et al. 2006.
  48. For an analysis of the role of public goods in collective action across 135 animal species, see Willems et al. 2013. For an analysis of the role of cultural variation in group selection and human evolution, see Bell et al. 2009.
  49. For further analysis of this issue, see Cowlishaw 1994.
  50. The manifold advantages of size in evolution have been explored in depth by John Tyler Bonner (1999, 2003). See also the article by Adrian Bejan (2012): “Why the bigger live longer and travel farther: animals, vehicles, rivers and the winds.” http://www.nature.com/articles/srep00594
  51. See Gintis et al. 2015 and the references therein; also, Corning 2017.
  52. On this issue, see Wrangham 1999, 2009; also, the discussion in Stanford 1999, p. 100.
  53. Gintis et al. 2015.
  54. Harmand et al. 2015; also, McPherron et al. 2010. See also http://news.sciencemag.org/africa/2015/04/world-s-oldest-stone-tools-discovered-kenya (accessed 23 June 2015).
  55. See especially the discussions in Lewin 1993; and Kingdon 1993; Ambrose 2001.
  56. See especially Klein 1999; Wrangham, 1999, 2009.
  57. See Shumaker et al. 2011.
  58. Pruetz and Bertolani 2007.
  59. Gintis and his colleagues agree. They believe that the use of stones as weapons may have been underrated; see Gintis et al. 2015.
  60. Does this mean that the australopithecines were “killer apes” after all? I would argue that this image is greatly overdrawn. Warfare is not an instinct; it’s a cultural invention. If there are deep psychological predispositions for aggressive behaviors, they are also shaped by experience and modulated by higher-level cortical processes. The challenge for our ancestors — as it is for us — was survival and reproduction. Survival was the overall challenge, and this involved an array of continuing, inescapable “basic needs.”
  61. Rose and Marshall 1996.
  62. Wrangham 1987; Foley 1995; Stanford 1999.
  63. the recent writings on intergroup aggression among primates and evolving hominins, see especially Manson and Wrangham 1991; Wrangham and Peterson 1999; Stanford 1999; Corning 2007a; Gintis et al. 2015. Gintis and his colleagues argue that weapons may also have served as internal power “equalizers” that encouraged the evolution of a more egalitarian social structure.
  64.  See Wolpoff 1999.
  65. Parrish 1996; Tokuyama and Furuichi 2016. A detailed analysis can be found in Wolpoff 1999a, pp. 141-142, 271-273.
  66. Discussed in depth in Sober and Wilson 1998.
  67. Holloway 1975, 1983a, 1996, 1997; Tobias 1971, 1985; see also Falk 1998; Falk, et al. 2000; Ambrose 2001.
  68. See especially Conroy, et al. 1988; Donald 1991; Aiello and Dunbar 1993; Dunbar 1996, 1998, 2001; Rilling and Insel 1999; Dor and Jablonka 2000; Ambrose 2001; Tomasello 2001, 2005, 2008.
  69. Dunbar 1996, 1998, 2001, 2009; also, Dunbar and Shultz 2007; Schultz et al 2011. see also Sewall 2015. As Sewall notes: “Cognition and communication both can be essential for effectively navigating the social environment and thus, social dynamics could select for enhanced abilities for communication and superior cognition. Additionally, social experience can influence both the ability to communicate effectively and performance in cognitive tasks within an individual’s lifetime, consistent with phenotypic plasticity in these traits.”
  70. Tomasello 2014; also Tomasello 2008. Tomasello explains that shared intentionality involves a “we” orientation, which affects our relationships to others and to our environment. He believes this provides the basis for our collaboration and our cultural institutions, as well as our communications and language.
  71. See Kingdon 1993.
  72. Teaford and Ungar 2000. See also the suggestive parallel in Oreopithecus bamboli, described in Alba, et al. 2001.
  73. Toth 1987a, b; Toth, et al. 1993; Schick and Toth 1993; also, Ambrose 2001.
  74. Steele 1999; Sterelny 2012; Morgan et al. 2015.
  75. Potts 1998a,b.

CHAPTER 8: THE SELF-MADE MAN: FROM HOMO ERECTUS TO HOMO SAPIENS

  1. Indeed, as more fossil evidence has accumulated in recent years, some hotly contested distinctions have begun to blur. For instance, new evidence of variability and overlapping traits in Australopithecus afarensis and Australopithecus africanus specimens has led some theorists to conclude that these hominins were merely subspecies rather than sharply divergent lineages. Likewise, the distinctions between australopithecines and early Homo (H. rudolphensis, H. habilis) do not seem as sharp as they once did. Not only did these ancestral hominins coexist for at least one million years but it now seems that the earliest systematic use of stone tools for hunting/scavenging actually occurred in australopithecines. Indeed, as more fossil evidence has accumulated, some hotly contested distinctions have begun to blur. It should also be noted that there is an unresolved debate among biologists about whether to view the evolutionary process – and human evolution in particular — as being incremental in nature or as a series of “transitions” (see especially Foley 2016, and Foley et al. 2016). I would argue that both are true. The ongoing process is indeed incremental but the functional outcomes may result in discontinuities — threshold effects and new synergistic combinations.
  2. http://www.nytimes.com/2015/09/11/science/south-africa-fossils-new-species-human-ancestor-homo-naledi.html?smprod=nytcore-ipad&smid=nytcore-ipad-share  New analyses indicate that the small-brained H. naledi may also have co-existed with early humans, perhaps 250,000 years ago.
  3. Asfaw, et al. 1999. See also the report on meat-eating in these hominins in De Heinzelin, et al. 1999.
  4. Harmand et al., 2015; see also Alemseged et al. 2006; McPherron et al. 2010. Some theorists liken these early, multi-use tools to Swiss army knives, but this is not really a good analogy. The Swiss army knife consists of a set of functionally specialized blades and other tools that have been bundled together. A better analogy is the famed Bowie knife of the American frontier. See S. Latham 1973.
  5. See especially B. Campbell 1985, pp. 396-398.
  6. Dunbar 1996; 1998; 2001, 2009; Ambrose 2001; Sewall 2015.
  7. Foley 1995; Lewin 1993; Rodman and McHenry 1980; McHenry 1992; Steudel 1994, 1996; Klein 1999; Wolpoff 1999a; Pontzer et al. 2010.
  8. See Washburn and Lancaster 1968; Blumenshine 1987; Shipman and Walker 1989; Wrangham and Peterson 1996; Stanford 1999; Wolpoff 1999a,b; Wood and Collard 1999; Klein 1999. Wilkins et al. 2012 report evidence for hunting with hafted stone spears at least 500,000 years ago. See also Sahle et al. 2013.
  9. See especially deMenocal 2004, 2011.
  10. By contrast, the more “robust” hominins of that era, with a distinctively different anatomy and dentition, may have deployed a feeding strategy that relied on a broad range of lower-quality plant foods.
  11. Well-crafted, light throwing spears have been dated back at least 400,000 years. See Thieme 1997; Ambrose 2001; Stringer 2012. Jabbing spears, that could seriously wound an animal, were no doubt used much earlier.
  12. See Pickering 2013.
  13. The so-called expensive tissue hypothesis is based on the evidence that the human brain represents about 2-3% of our total body mass, yet it consumes some 20% of our energy budget and even more, about 27%, in a developing infant. See R.D. Martin 1981; R.D. Martin and MacLarnon 1985; Leonard and Robertson 1994, 1997; Leonard et al. 2007. Aiello and Wheeler (1995) point out that there may have been a tradeoff — the increase in brain tissue associated with meat eating may have been offset by a reduction in the size of our guts (also expensive tissue) in response to dietary changes. However, larger overall size, extended childhood dependency, greater longevity and shortened birth-spacing must also be included in the bookkeeping.
  14. Aiello and Key 2001.
  15. Bramble and Lieberman 2004. They note that distance running uses only 30 percent more energy than walking (about 900 kilocalories per hour). Killing a duiker might yield 15,000 kcal in return, and a wildebeest might yield 240,000 kcal. See also Lieberman 2013; Stanford 1999; Bunn 2001; Bunn and Gurtov 2014; Henrich 2016.
  16. The work of Peter Ungar and his colleagues suggests that there was also omnivory and a high degree of local opportunism. See Ungar and Sponheimer 2011; also, Ungar et al. 2006. There is even possible evidence of repiratory tract changes to accommodate a more strenuous life-style. See Trinkaus 1987.
  17. Newman 1970; also, Wheeler 1985, 1991. Discussions of sweating and hair loss as major adaptive changes in evolving humans can also be found in Foley (1995) and Wolpoff (1999a), among others.
  18. See Hrdy 2000, 2009; also, Burkhart et al. 2009; 2014. See also the “grandmother hypothesis” advanced by Kristin Hawkes (2003) and the three-generation model of Kaplan et al. 2009. In an important comparison across 78 primate species and 65 carnivore species, Caroline Schuppli et al. (2016) found that access to a more complex and stable foraging niche was highly correlated with a long childhood and parental food provisioning.
  19. Meet the Alloparents, Natural History Magazine, April 2009   http://www.naturalhistorymag.com/htmlsite/0409/0409_feature.pdf
  20. Darwin 1874/1871. See also the discussion in Wrangham and Carmody 2010.
  21. An older analysis, but still useful is Clark and Harris 1985. A more recent analysis, based on archaeological evidence for an upsurge in natural fires, can be found in Parker et al. 2016.
  22. Bellomo 1994; B. Campbell 1985; Kingdon 1993. Kingdon also points out that fire was too commonplace and too powerful in its effects to have been overlooked as a potential tool until only a few hundred thousand years ago.
  23. Wrangham et al., 1999; Wrangham 2009. However, Gowlett and Wrangham (2013) use a more conservative 1.5 million years, while Stahlschmidt,et al. (2015) question the methodologies used in many of the findings and call for the use of new techniques for “microanalyses.” But see also Parker at al. 2016.
  24. Wrangham et al., 1999; Wrangham 2009; Wrangham and Carmody 2010; Gowlett and Wrangham 2013; Smith et al. 2015. Wrangham and his colleagues develop a case for the hypothesis that fire and cooking played a major causal role in the transition from Australopithecus to H. erectus. See also the analysis in N. Mann (2007) indicating that we evolved as true omnivores. The evidence for meat eating includes such things as our dentition, the morphology of our gastrointestinal tracts and deficiencies in the body’s ability to produce certain nutrients on its own. Also, see A. Mann 1981. A new hypothesis, that the processing and tenderizing of raw foods preceded cooking, has been proposed by Zink and Lieberman (2016). The problem is that this would have reduced chewing time by only 17 percent (by their own calculations) and would not have reduced digestion time or the need for a larger gut. At best it might have been only a first step.
  25. Wrangham 2009; see also Fonseca-Azevedo and Herculano-Houzel 2012; also, Herculano-Houzel 2012. As Ann Gibbons noted, (ScienceNOW, 22 October 2012): “Humans have more brain neurons than any other primate—about 86 billion, on average, compared with about 33 billion neurons in gorillas and 28 billion in chimpanzees. While these extra neurons endow us with many benefits, they come at a price—our brains consume 20% of our body’s energy when resting, compared with 9% in other primates. So a long-standing riddle has been where did our ancestors get that extra energy to expand their minds as they evolved from animals with brains and bodies the size of chimpanzees?”
  26. Wrangham believes that cooking may have precipitated the introduction of home bases and played a major role in changing the mating system toward stable pair-bonds (the nuclear family), leading in turn to such anatomical changes as the concealment of ovulation and the permanent receptivity of the females. Wrangham argues that this social change was not primarily related to male provisioning, as Lovejoy supposed, but to the defense by males of their hard-won food supplies. Wrangham calls it the “theft hypothesis.”
  27. A lengthy letter to the editor of the journal Science by anthropologist Ralph Rowlett (1999, p. 741) reinforced this argument. He concluded: “[Hominins] clearly had the pyrotechnical ability to cook tubers at least as far back as 1.6 million years ago, even if further research must determine exactly what was cooking.” Further unambiguous evidence tracing back at least 1 million years was reported Berna et al. 2012. See also Growlett and Wrangham 2013.
  28. Klein 1999; Wolpoff 1999a.
  29. Pinker 2010; also, Henrich 2016, p. 65ff. The idea of collective intelligence has gained support recently across a number of disciplines. See especially Couzin 2007; Woolley et al. 2010; van Schaik and Burkart 2011; Whiten and Erdal 2012.
  30. See especially Boehm 1993, 1997, 1999; also, Richerson and Boyd 1992; Gintis et al. 2105. For a recent critique and debate about what has also been called “strong reciprocity theory,” see Guala 2009 and numerous commentaries. See also Boyd and Richerson 2005; Richerson and Boyd 2005; also, the similar argument in Wilson and Wilson 2008.
  31. Indeed, we now know that social learning is widespread throughout the animal world, including even in fishes, songbirds and, of course, the primates. See the special issue of the Philosophical Transactions of the Royal Society B in 2011, especially Whiten et al., 2011.
  32. Gavrilets and Richerson (2017), in an important new paper, model the evolution of the capacity to internalize and follow costly norms. They view this as a crucial step in human evolution.
  33. Piaget 1932; Kohlberg 1981; Kohlberg et al. 1983; Tomasello 2001, 2005, 2009, 2014, 2016.
  34. Tomasello develops his “shared intentionality” hypothesis at length in a 2014 book, where he reviews and synthesizes recent work showing that humans have uniquely evolved cognitive adaptations for engaging in collaborative interactions. “Compared with other animal species,” he observes, “humans think in special ways” (p. 4). These cognitive skills are both derived from and facilitate relationships involving shared objectives, he believes.
  35. Henrich 2016 pp. 56ff.; p. 212, p. 250. He argues compellingly that “culture makes us smart.” We rely on “a large body of locally adaptive, culturally transmitted information that no single individual, or even group, is smart enough to figure out in a lifetime” (p. 12).
  36. Boyd et al. (2013), in a major review, argue that technological innovation is typically an iterative, cooperative process. They conclude; “The tools essential for life, even in the simplest foraging societies, are typically beyond the inventive capacities of individuals. They evolve, gradually accumulating complexity through the aggregate efforts of populations of individuals, typically over many generations” (p. 120).
  37. Boesch and Tomasello 1998; see also Tomasello 2001, 2009. For recent findings on the antiquity of various technological improvements, see Bar-Yosef and Kuhn 1999. Thieme (1999) also reported the discovery of 400,000 year-old Neanderthal spears. Some bone tools also date back at least 150,000 years. See Bower 1997; also, the broader perspective in Ziman, ed. 2000; also, Richerson and Christiansen eds. 2013. Stephen Shennan in this volume points out that the “ratchet” metaphor can also be misleading. There are also many cases where the ratchet slipped or even disengaged. There is nothing inevitable about the process. Another version of the ratchet effect, first suggested by Nicholas Humphrey and seconded by Alison Jolly, refers to the progressive evolution of our cognitive abilities. Humphrey likened the process to that of a self-winding watch. See Humphrey 1976.
  38. Henrich 2016, p. 57. He calls it a “Rubicon” where cultural evolution became the “primary driver” of our species genetic evolution. However, one must be careful not to reify culture as being some sort of independent agency or “driving force” (p. 317). The cultural innovations adopted by our ancestors were, for the most part, driven by their economic/functional benefits in terms of earning a living. The process became cumulative only insofar as it proved effective as a survival strategy. See also Thompson et al. 2016.
  39. Chomsky 2004.
  40. Pinker 1994; also, Pinker and Bloom 1990. See also the review in https://en.wikipedia.org/wiki/Origin_of_language (last updated 22 June 2015).
  41. This is discussed at length in Corning 2005; also, Corning 2007b.
  42. This aspect of language is discussed in some depth in Richerson and Boyd 2010; Richerson and Christiansen 2013; see also Tomasello 2005, 2008; Henrich 2016.
  43. Four insightful discussions of this approach are Jablonka 2002, Tomasello 2008, Sewall 2015, and Henrich 2016. There is also much of value in the recent book by Bickerton (2009), although I disagree with the thesis that recruitment for confrontational scavenging was the driver.
  44. Sherman et al. 1991, 1992.
  45. Suzuki et al. 2016.
  46. The case for language as a functional imperative in evolving hominins is argued in depth by Dunbar 1996, 1998, 2001, 2009; also, Henrich 2016.
  47. Mithen 2006. Also important, most likely, were body language and facial expressions, as Darwin himself (1965/1873) documented in his book The Expression of the Emotions in Man and Animals .
  48.  Snowdon 2001. Michael Corbalis (2003), Michael Tomasello (2008), and others have theorized that intentional gestures (body language) of various kinds likely preceded the emergence of verbal language skills among our ancestors. Tomasello notes that vocalizations in chimpanzees and other primates are mainly associated with the expression of emotions, whereas body language and various gestures are used for intentional communications. He also points out that human infants deploy non-verbal “signals” for communications before they learn how to talk. Therefore, he reasons, our verbal language has a gestural foundation. It arose from our evolving pattern of social cooperation – “shared intentionality” — along with our increasing skills for cultural learning. However, it should also be noted that vocal communication has the enormous advantage of freeing up the hands for other uses. And it’s not limited only to recipients who are in line of sight and paying attention. See also Stout and Chaminade 2011; Smit 2014.
  49.  The idea of self-domestication has been applied especially to our propensity to be obedient rule-followers. As Henrich (2016) puts it: “Natural selection shaped our psychology to make us docile, ashamed at norm violations, and adept at acquiring and internalizing social norms. This is the process of self-domestication” (p. 319). See also Leach 2003; Russell 2012.
  50. See especially Tomasello 2005, 2008. Also Mazda Y Farias-Virgens and Yevgeniya Sosnovskaya, Self-Domestication and the Evolution of Human Language. http://berkeleysciencereview.com/self-domestication-and-the-evolution-of-human-language/ (accessed 12 January 2016).
  51. See especially Levinson and Dediu 2013, Dediu et al. 2013; also, Szathmáry 2002, 2015; Deacon 1997; Szathmáry and Számadó 2008; Bickerton and Szathmáry 2011; Tomasello 2008.
  52. See the full discussion of this issue, and the evidence, in Deacon 1997.
  53. See Holloway 1983b. Deacon (1997) also points to the conclusions of endocast researchers Philip Tobias and Dean Falk that language adaptations can be discerned even in H. habilis.
  54.  See especially Levinson and Dediu 2013, Dediu et al. 2013; also MacLarnon and Hewitt 1999.
  55. See especially Goren-Inbar 2011.
  56. See especially Levinson and Dediu 2013, Dediu et al. 2013. A similar argument is developed in Richerson and Boyd (2010). Among other things, they point out that the very diversity and local specificity of human languages acts to limit communications to insiders and exclude those who are not members of a group, reinforcing thnocentrism.
  57. Aristotle 1946/ca. 350 B.C.
  58. Corning 2017.
  59. de Waal 1982; Strum 1987.
  60. Particularly noteworthy is the work of Mark van Vugt and his colleagues that has reformulated the field of leadership within an evolutionary/functional and psychological perspective. See especially van Vugt 2006; van Vugt and Ahuja 2011. Also, see the major cross-species study by J. Smith et al. 2016; also Kummer 1968; de Waal 1997; Conradt and Roper 2003. Distributed control provides another, as yet poorly understood model. See Gordon 2007
  61. Boehm 1993, 1997, 1999. See also Brown 1991.
  62. Gintis et al. 2015.
  63. Corning 2017. The concept of “prestige” as a basis for leadership is a well-established theme in anthropology. As Henrich (2016) observes: “Across human societies, prestige is consistently associated with great skill, knowledge and success in activities or tasks people care about. This prestige status readily forms a foundation for leadership in egalitarian societies” (p. 118).   Consensual leadership derives ultimately from the convergent self-interests of the leaders and followers. See also Cheng et al. 2013. For a critical analysis, see Chapais 2015.
  64. Gintis et al. 2015. See also Boehm 1996, 1997; Puurtinen and Mappes 2009; Richerson et al 2016.
  65. Anthropologist Richard G. Klein (2000) is the strongest proponent of a Apunctuational@ and biological cause. “The shift to a fully modern behavioral mode and the geographic expansion of modern humans were coproducts of a selectively advantageous genetic mutation” (pp. 17-18). Klein finds the human ability to innovate otherwise puzzling and speaks of the “sudden origin of the modern capacity for culture” (p. 24). I do not find it so inexplicable but instead view cultural evolution as a cumulative learning process, a ratchet.
  66. Wolpoff (1999a) and Klein (1999) interpret the available fossils and other artifacts in somewhat different ways. Wolpoff sees a multi-regional pattern of progressive changes. Klein sees a more unitary process with periods of relative stasis interrupted by punctuational changes. Yet both acknowledge that progressive changes were occurring; see also Ambrose 2001; Wilkins et al. 2012.
  67. On this point, see especially Mcbrearty and Brooks 2000.
  68. Wolpoff 1999a, p. 554. Wolpoff’s views were supported by subsequent discoveries of more sohisticated artifacts with much earlier dates in diverse regions. See Brooks, et al. 1995; Yellen, et al. 1995; McBrearty and Brooks 2000; Kuhn, et al. 2001; d’Errico and Stringer 2012. It is also probably safe to assume that many more artifacts have long since decayed or were buried under modern settlements.
  69. Klein 2000, p. 26; also, Klein 1999.
  70. Stringer 2012, p. 240.
  71. Wolpoff 1984, 1999b; others have been associated with Clark Howell and Christy Turner.
  72. Wolpoff 1999b; Wolpoff et al. 2000; Jurmain et al. 2008; Stringer 2012.
  73. See the 2016 CARTA symposium on Ancient DNA and Human Evolution, https://carta.anthropogeny.org/node/156412 (accessed 29 May 2016).
  74. See especially Klein 1999, 2000; Ehrlich 2000; Semino, et al. 2000; Bar-Yosef 2002; Stringer 2003, 2012; Mellars 2006; Liu et al. 2006; Roberts 2009. There is much variation in the estimates for inclusive dates. Between 150-350,000 years ago is also mentioned.
  75. It seems possible that some initial, tentative migrations may have occurred earlier. This would account for the human remains found at Skhul, Isreal, that date to about 100,000 years and the recent finding by Alan Thorne that an Australian fossil human, known as Mungo 3, now appears to be about 60,000 years old. See also Tucci and Akey 2016.
  76. Stringer 2012, p. 253. In a recent (2015) talk, Stringer described the process as a “coalescence” – a gradual accretion that was “assembled piecemeal.” See https://royalsociety.org/events/2015/10/major-transitions/
  77. Hublin et al. 2017.
  78. Lieberman 1998, pp. 85-97; see also Tattersall 1998; cf Deacon 1997. Many theorists in the waning years of the twentieth century were seduced by the discovery of the FOXP2 gene, which is strongly associated with language skills. At first it was believed that it was unique to humans – a mutant gene that gave us speech. Later research showed that a great many other species have variants of this gene and that it plays a general role in animal vocalizations, as well as other traits. However, the human version also has some unique features. What this may mean is still unknown. See https://en.wikipedia.org/wiki/FOXP2 (last updated 8 June 2015).
  79. Diamond 1995.
  80. Tattersall 2000, pp. 56-62.
  81. Klein 1999; see also the discussion in Wade 2006.
  82. Richerson et al. 2005; Mellars 2006; Potts 2012.
  83. Boyd and Richerson 2009; Richerson et al. 2009; Boyd et al. 2011; Richerson et al. in press; also, Henrich 2016.
  84. A number of other theorists — Richard Klein 1999, 2009, Luca Cavalli-Sforza et al. 1994, Jared Diamond 1997, Christopher Wills 1993, 1998, Paul Ehrlich 2000, Ian Tattersall 2002, Paul Mellars 2006, John Shea and Matthew Sisk 2010, Dniel Liberman 2013; Joseph Henrich 2016, and others — hold that the emergence of a more advanced technological package was an important factor in the modern human diaspora.
  85. Richerson et al. 2009; Boyd et al. 2011; Collard et al. 2013; Derex et al. 2013. Marean stresses the exploitation of marine resources. See https://royalsociety.org/events/2015/10/major-transitions/
  86. Richerson 2013; Henrich 2016; also, Powell et al. 2009. Henrich argues that the size of a group and its degree of interconnectedness played a crucial role; bigger groups have the potential for more rapid cultural evolution. However, this in turn required an economic system that could sustain larger numbers and encourage interconnectedness. What could be called the population hypothesis has also been criticized. See Andersson and Read 2016; also, Vaesen et al. 2016a,b; and the response by Henrich et al. 2016.
  87. Salisbury 1962, 1973.
  88. On hafted hand axes, see especially Clark 1992; Wadley et al. 2009. On the recent discoveries suggesting that bows and arrows were used as early as 64,000 years ago, see Lombard and Phillipson 2010; also, Mellars 2006. Tattersall (2000) notes that the more advanced, hafted axes had about 10 times as much cutting edge per pound of material as the more primitive hand axes. The manufacture of these advanced tools/weapons apparently involved a highly specialized process that could not readily be imitated. A form of glue — bitumen or a similar substance B heated to a very high temperature — was used for attaching the stone blades to wooden shafts. It was not easy to do. However, this does not mean that these implements were invented only for the purpose of making war. As we noted earlier, things that are devised for one purpose (both in biological and cultural evolution) may subsequently be adapted for quite different purposes. Indeed, it now seems that Neanderthals may also have developed (or perhaps borrowed) this tool-making technique as well. See also Farmer 1994; Shea and Sisk 2010; d’Errico and Stringer 2012; Marean 2015.
  89. The history of this theory, from Darwin to the present day, is reviewed in van der Dennen 1995, 1999. Among the many more in-depth recent works, see especially Alexander 1979; Keegan 1993; Crook 1994; Keeley 1996; Wrangham and Peterson 1996; Thorpe 2003; Otterbein 2004; Kelly 2005; Corning 2007a; Bowles 2009.
  90. A starting point for the extensive literature on ethnocentrism and xenophobia is Reynolds et. al. 1987; also, Shaw and Wong 1989; van der Dennen 1995, 1999; Eibl-Eibesfeldt and Salter, eds. 1998. In case it needs to be said, the assertion that we may have a psychological predisposition is not to condone it, or to accept it fatalistically. There is also good evidence that cultural means can be used to contain and even rise above our innate urges, but this is more likely to occur if we recognize them for what they are. Indeed, human populations often do live in peace with one another and even form alliances against common enemies.
  91. See Corning 2007a and citations therein. The provocative recent claim by anthropologist Douglas Fry and psychologist Patrik Söderberg (2013) that their new study of 21 contemporary forager “bands” suggests an absence of warfare among our ancestors (they found only personal disputes, not lethal collective violence) is refuted by anthropologist Carol Ember’s (1978) more extensive review of the many field studies documented in the Ethnographic Atlas. She found that frequent violent conflicts (some lethal, others not) were nearly universal. The data for 50 hunter-gatherer societies showed that 64 percent had engaged in warfare at least once every two years, while 26 percent did so less often, and only 10 percent were reported to have had none, or only very rarely. See also Otterbein 2004; Marlowe 2005.
  92. See Lahr et al. 2016. The authors found that at least 10 of the 12 skeletons found at a site near Lake Turkana appear to have died from violence. Also, the extensive evidence in pre-historic California reported in M. Allen et al. 2016.
  93. Simonti et al. 2016.
  94. See Corning 2007a.
  95. Diamond 1997, pp. 53-57.
  96. See especially Clark 1992; Farmer 1994; Ambrose 2001; Mellars 2006; Wadley et al. 2009; Lombard and Phillipson 2010; Shea and Sisk 2010; d’Errico and Stringer 2012; Lieberman 2013; Marean 2015.
  97. See Mellars and French 2011.
  98. Marean 2015.
  99. In a recent (2015) talk, Marean emphasized the role of dense and predictable food resources in stimulating territoriality and intergroup competition. See https://royalsociety.org/events/2015/10/major-transitions/
  100. The many different forms of synergy that are associated with collective violence in the natural world are explored in detail in my paper on “Synergy Goes to War” (Corning 2007a).
  101. See Corning 2007a; also, Lorenz 1966; E.O. Wilson 1975; Alexander 1979; Keegan 1993; Crook 1994; van der Dennen 1995, 1999; Wrangham and Peterson 1996; Stanford 1999.
  102. Robert Foley, in a 2015 symposium on the major transitions in human evolution, concluded that it is our massive capabilities and extraordinary impact that sets us apart. This transition was not a punctuational event, he said, but a combined effect of influences at many levels. See https://royalsociety.org/events/2015/10/major-transitions/
  103. The provocative theory of Paul Bingham (2000) should also be mentioned. Bingham believes there is a “simple, unitary, inexorable logic to the entire human story” (p. 255). What he proposes is, in essence, a synergistic nexus — cooperation by non-kin coalitions that were reinforced by policing and ethics, novel technologies that allowed for “killing at a distance” (which greatly reduced individual risks) and competition between groups. Perhaps so, but the logic of it seems “inexorable” only in retrospect.
  104. For an in-depth comparison, see Whiten 2011.
  105. These theorists include C. Loring Brace, Milford Wolpoff, Richard Klein, Ralph Holloway, Jonathan Kingdon, Robert Foley, Terrence Deacon, Michael Tomasello, Edward Wilson, Martin Nowak, Kevin Laland and John Odling Smee, Boyd and Richerson, Kim Sterelny, Curtis Marean, Herbert Gintis and his colleagues, Joseph Henrich, and others as well. The extended case developed by Richerson and Boyd (2005), Boyd and Richerson (2009) and Laland et al. (2010) and Henrich (2016) are especially noteworthy. Boyd and Richerson emphasize our ability to learn from one another and the capacity of culture to augment heritable variation between groups. Laland et al. review some of the various models of gene-culture co-evolution and detail the evidence for various human traits. Henrich stresses the importance of our “cumulative cultural evolution.”
  106. See especially Stoelhorst and Richerson 2013; Lieberman 2013; also, Bowles 2009; Bowles and Gintis 2011.
  107. Recent studies among contemporary hunter-gatherers reveal that these groups contain many unrelated individuals, and that they have extensive social networks – both internally and externally with neighbors and trading partners. It indicates that human cooperation has deep roots and goes far beyond inclusive fitness (Chapter Two). See especially Hill et al. 2011; Apicella et al. 2012.
  108. Tomasello et al. 2012.

CHAPTER 9: THE RISE OF COMPLEX SOCIETIES

  1. There is a debate among social scientists about how to define social complexity that parallels the debate among evolutionary biologists. A useful synthetic approach, recently suggested by a group of social scientists at a conference on cultural evolution, included such factors as population size, territory size, settlement density, economic specialization, trade networks, management mechanisms, number of levels of hierarchical control, and more. See Richerson and Christiansen 2013, pp. 87-116.
  2. Pfeiffer 1977.
  3. Diamond 1997. For a broad discussion of the role of technology in human evolution, see Boyd et al. 2013.
  4. See especially Bogucki 1999; Mithan 2003.
  5. Richerson et al. (2001) argue that agriculture became “compulsory” in the Holocene as a result of a “competitive ratchet” driven by competition between groups in the context of population pressures. See also the discussion in Sterelny 2013; also, the analysis of Winterhalder and Kennett (2009), utilizing the economic concepts of risk, discounting, economies of scale and transaction costs. Social networks and more elaborate patterns of exchange may also have played a significant role. See Apicella et al. 2012.
  6. See especially the analysis in Gowdy and Krall 2014.
  7. An analysis of this transition can be found in Flannary and Marcus 2012. What seems to hold the social contract together in complex societies, despite often extreme differences in wealth and poverty, is some combination of economic interdependence, orderly exchange (markets), the need for a common defense, social and cultural/communal bonds, coercion (there is no available alternative) and policing.
  8. Fagan 1998. See also Flannery and Marcus 2012.
  9. See especially the special issue of Current Anthropology in 2010 devoted to this subject, especially Bowles et al. 2010; Smith et al. 2010; Shenk et al. 2010; Gurven et al. 2010. Also, see Flannery and Marcus 2012. See also the important analysis of what is known as the “endowment effect” in facilitating the evolution of private property in various species by Gintis 2007.
  10. Flannery and Marcus 2012, pp. 67-71. Their example was derived from an in-depth study by archeologist Jeanne Arnold and her co-workers. Many details also come from eye witness accounts by early eighteenth century Spanish colonists.
  11. Boehm 1993, 1999. Recall that Boehm’s term refers to coalitions that actively contain and suppress aggressive individuals. See also Brown 1991.
  12. Flannery and Marcus 2012, p. 563.
  13. Ibid. Recent DNA studies have revealed that the earliest farming communities arose independently and only much later interacted (and interbred) See http://www.nytimes.com/2016/10/18/science/ancient-farmers-archaeology- dna.html?action=click&contentCollection=science&region=rank&module =package&version=highlights&contentPlacement=1&pgtype=sectionfront (accessed 18 October 2016)
  14. A notable example is the Indus River civilization known as the Harappans, from about 4500 to 4000 years ago, which evidently never had a centralized ruling class or extremes of wealth and poverty. See Maisels 1999.
  15. Trigger 2003, p. 375.
  16. Boehm 1993, 1999. Sterelny (2013) also argues that a shift from reliance on “social capital” (skills) to “material capital” (land) played a key role. But so did the transformation of the economic system from one that was based on collective action to one that was based on exchange and reciprocity among different specialists.
  17. See Flannery and Marcus 2012.
  18. For a brief history, see https://en.wikipedia.org/wiki/History_of_democracy (last modified 30 October 2105).
  19. Spencer 1852, 1897.
  20. Spencer 1897, vol. 1, 1 pp. 14-15.
  21. Parsons 1949/1937, p. 3.
  22. Childe 1951/1936.
  23. Wittfogel 1957.
  24. Cohen 1977.
  25. See especially Jared Diamond’s (2005) book length treatment. Also, see Corning 2005, Ch. 7.
  26. White 1959, p. 56.
  27. White 1949, p. 39.
  28. Alexander 1987.
  29. Carneiro 1981; also, Carneiro 1978.
  30. Carneiro 1970.   An example of this dynamic can be found in the history of the Zulu nation. See especially Gluckman 1940,1969; Morris, 1965; Corning 2003.
  31. Turchin 2009, 2011; Turchin and Gavrilets 2011; Turchin 2016.
  32. See Corning 2005, ch.7; also Corning 2007a.
  33. Gowdy and Krall 2014, 2015.
  34. A variation on this theme is physical chemist Ugo Bardi’s (2014) theory that the rise of complex states was driven by a quest for mineral resources – gold, iron, bronze, etc. Yes, but first you must feed the population.
  35. This material was obtained personally from displays at the Gold Mining Museum, Angels Camp, California, 2004.
  36. Ridley 2010, p. 4.
  37. See http://en.wikipedia.org/wiki/Industrial_Revolution (last modified 10 November 2014).
  38. Service 1971, p. 25.
  39. See also Corning 2007a.
  40. Political scientist Roger Masters (2008), in a systematic analysis of state formation, identified a number of important contributing factors, from a favorable environment to exceptional leaders, ethnic and religious commonalities, a common language, writing and record keeping, monetary systems, mutually beneficial economic activities, and, of course, military conquest. See also Richerson and Christiansen, eds. 2013
  41. To be clear, this theory does not make a claim to explain every aspect of culture, and cultural change. It applies to functional improvements over time in the “means of production” and reproduction – the collective survival enterprise. For extended discussions of the role of synergy in cultural evolution, see Corning 1983, 2005; also, see Plotkin 2010; Mesoudi 2011; Richerson et al. in press. Does the Synergism Hypothesis and Synergistic Selection therefore provide a general theory of cultural evolution? A prior question is how do I define culture and cultural evolution? As with many other theoretically important concepts, these can be defined in various ways, and there is a long-standing, sometimes heated debate about this issue in the social sciences. I prefer to define the term “culture” very broadly as a body of socially transmitted knowledge, skills, and artifacts that are passed down from generation to generation within a social group via learning and teaching (see Chapter Eight). Of course, any culture is subject to change over time. Song styles may differ; male beards may come and go; Leonardo da Vinci was succeeded by Jackson Pollack. But I define cultural evolution more narrowly as the sub-set of cultural changes – social, economic, political — that are related to biological adaptation and our basic needs (Chapter Ten).  Cultural evolution is thus a social/group phenomenon that affects the differential survival and reproduction of a group and its members over time. This definition keeps it within the framework of evolutionary biology and the concept of a society as a “collective survival enterprise.” Accordingly, the Synergism Hypothesis posits that functional synergies have been the drivers for the evolution of larger, more complex societies over time; synergies of various kinds have underwritten the evolving functional benefits.
  42. See https://en.wikipedia.org/wiki/Container_ship (last modified 19 July 2015); see also http://www.worldshipping.org/about-the-industry/liner-ships/container-vessel-fleet (accessed 24 July 2015). https://www.statista.com/topics/1367/container-shipping/ (accessed 26 March 2017).
  43.  https://en.wikipedia.org/wiki/World_population (last modified 23 July 2015).

 CHAPTER 10. THE NEXT MAJOR TRANSITION

  1. Quoted in Stringer 2012, p. 263.
  2. See especially Wills 1998; also, Stock 2002; Hawks et al. 2007; Cochran and Harpending 2009; Thomas 2015. See also the op-ed article in the New York Times by Menno Schilthuizen, Evolution Is Happening Faster Than We Thought. http://www.nytimes.com/2016/07/24/opinion/sunday/evolution-is-happening-faster-than-we-thought.html?emc=eta1 (accessed 24 July 2016). If further evidence of rapid evolution is needed, consider the diversity of domesticated dogs.
  3. Hawks et al. 2007; also, Stringer 2012, pp. 265-269; Lieberman 2013, p. 205.
  4. The evidence for such micro-evolutionary change as an ongoing process in human evolution was reviewed by Wills 1998; see also Wolpoff 1999a. New DNA evidence for genetic changes associated with the agricultural revolution were recently reported by Mathieson et. al. 2015. See also “Not what they were: Researchers can now watch human evolution unfold.” The Economist May 14, 2016, pp. 71-72.
  5. Laland et al. 2010. On the other hand, we also suffer from an array of what Daniel Lieberman (2013, pp. 168-174, 202) calls “mismatch diseases” – non-infectious ailments that are a product of a misfit between our Paleolithic bodies and modern life. Lieberman compiled a partial list of 49 of these that included such things as acne, dental cavities, sleep apnea, glaucoma, insomnia, scurvy, high blood pressure, and stomach ulcers. In addition, we suffer from more than 100 infectious mismatch diseases, most of which have arisen since the Agricultural Revolution.
  6.  https://en.wikipedia.org/wiki/Anthropocene (last modified 27 January 2016). At this writing, a working group of geologists have proposed to formally designate the beginning of the Anthropocene Epoch to 1964, when atmospheric nuclear testing spread traces of plutonium world wide. Waters et al. (2016) argue for the 1950s and the world-wide deposition of “technofossils” from manufacturing processes. Fire expert Stephen Pyne (2001, 2012) suggests that “The Anthropocene might equally be called the Pyrocene.” Some ecologists characterize humankind as “the world’s greatest evolutionary force.” See Hendry et al. 2016. For an analysis of the ecological consequences, see Boivin et al. 2016.
  7. There are a great many different kinds of negative synergy (or “dysergy”) in the natural world – cooperative effects that are deleterious to one or more participants, or to various “bystanders”. Whether synergy can be called “positive” or “negative” depends, quite simply, on the value that is assigned to it. For instance, cooperative hunting might be considered very beneficial by a group of predators, though the outcomes would be viewed as negative synergy by their prey. Parasitism provides innumerable examples of such a difference in perspective. (More on negative synergy can be found in Corning, 2003)
  8. This estimate was made by the United Nations’ Food and Agriculture Organization. See http://www.scientificamerican.com/article/only-60-years-of-farming-left-if-soil-degradation continues/ (accessed 22
  9. Among others, see Wills 1998; Stock 2002, 2007; Cochran and Harpending 2009; Lieberman 2013; Thomas 2015.   See also the op-ed article in the New York Times by Menno Schilthuizen, Evolution Is Happening Faster Than We Thought. http://www.nytimes.com/2016/07/24/opinion/sunday/evolution-is-happening-faster-than-we-thought.html?emc=eta1 (accessed 24 July 2016). If further evidence of rapid evolution is needed, consider the diversity of domesticated dogs. On the Indian drought, see http://www.bbc.com/news/world-asia-india-36089377 (accessed 5 May 2016).
  10. The current target for containing global warming falls short. The combination of warming (therefore expanding) oceans, rapid melting of the world’s 200,000 glaciers, and an acceleration in the melting of the vast Greenland and Antarctic ice sheets could have synergistic consequences of the worst kind for sea levels. The prominent glaciologist Eric Rignot warns that a 2-3 degree Celsius increase in global temperatures would probably “take out” the Greenland ice sheet. See Jon Gertner, “The Secrets in Greenland’s Ice Sheet.”   http://www.nytimes.com/2015/11/15/magazine/the-secrets-in-greenlands-ice-sheets.html?smprod=nytcore-ipad&smid=nytcore-ipad-share
  11.  The original version, by economist Herbert Stein, was: “If something cannot go on forever, it will stop.” See http://www.goodreads.com/quotes/467811-if-something-cannot-go-on-forever-it-will-stop (accessed 16 May 2016). A similar idea was later expressed by economist Kenneth Boulding: “Anyone who believes that exponential growth can go on forever in a finite world is either a madman or an economist.” See http://www.goodreads.com/quotes/627148-anyone-who-believes-that-exponential-growth-can-go-on-forever (accessed 16 May 2016).
  12. Ehrlich and Harte 2015.   They cite at least eight major “top of our list” policy changes, ranging from a global carbon tax to drastically reducing the use of pesticides and other chemicals in agriculture. They conclude “We find it hard to be optimistic.”
  13. Cited in Ehrlich and Ehrlich 2012.
  14. http://www.nytimes.com/2016/05/28/world/asia/text-of-president-obamas-speech-in-hiroshima-japan.html?_r=0 (accessed 28 May 2016).
  15. Kiers and West 2015. The authors say, in part: “Symbiotic partnerships are a major source of evolutionary innovation. They have driven rapid diversification of organisms, allowed hosts to harness new forms of energy, and radically modified Earth’s nutrient cycles…. The same problem—how to overcome the selfish interests of individuals to form mutually dependent cooperative groups—has arisen and been solved at several crucial moments in history across all orders of life…. Transitions in individuality are rare and require strict conditions: Partner interests need to be aligned and the benefits of more integrated cooperation must lead to mutual dependence…. How can group conflict be eliminated and loss of autonomy become favorable? …The resulting symbioses do not necessarily eliminate group conflict. Furthermore, although repression of competition is necessary, it is not sufficient to drive a major evolutionary transition, which requires mutual dependence.”
  16. Richerson and Boyd 1999. See also Boehm 1997.
  17. These fourteen categories are detailed and documented in Corning 2005, 2011.
  18. It should be noted that this list is convergent with U.N.’s Human Development Index, although it is more explicitly rooted in a biological framework. The important work on developing a Social Progress Index should also be noted here. It represents an effort to create an array of more than 50 social and environmental indicators to supplement our traditional focus on economic measures of well being, like the Gross Domestic Product. The so-called Social Progress Imperative, founded in 2010, has produced three annual reports, with ratings for 133 countries. This formulation draws on psychologist Abraham Maslow’s (1954) famous hierarchy of human needs, and it involves three dimensions: basic human needs, the foundations of well being, and opportunity. Although it is convergent with the basic needs that have been identified in the Survival Indicators Project (see Corning 2005, 2013), it is broader in scope and less focused on the biological challenge of reproducing the next generation. For more on the Social Progress Index, see http://www.socialprogressimperative.org/about/the-imperative (accessed 10 January 2016); https://en.wikipedia.org/wiki/List_of_countries_by_Social_Progress_Index (last modified 7 February 2016).
  19. Equality has been a socialist and liberal/progressive ideal ever since the Enlightenment. See especially Sandel 2009. Many theorists have focused on equality in terms of human rights, economic opportunities, or due process of law. However, economic egalitarians, beginning with Rousseau (1984/1762), have stressed especially an egalitarian distribution of the wealth and property of a society. From the perspective of the three fairness precepts associated with the biosocial contract paradigm, the radical socialist ideal is misguided. Or, better said, it must be strictly limited to the domain in which we are indeed equal – our basic needs. This leaves ample room for differentially rewarding our inequalities in talent, efforts and achievements. (The shortcomings of socialism are discussed in some detail in my 2011 book.)
  20. Frohlich and Oppenheimer 1992.
  21. See https://www.washingtonpost.com/news/wonk/wp/2016/04/25/bernie-sanders-is-profoundly-changing-how-millennials-think-about-politics-poll-shows/?tid=a_inl (accessed 10 May 2016).
  22. See https://en.wikipedia.org/wiki/Guaranteed_minimum_income (last modified 25 March 2017). Among others, supporters of the idea have included Thomas Paine, Henry George, Milton Friedman, Friedrich Hayek, Martin Luther King, Daniel Patrick Moynihan, and a manifesto signed by 1200 economists in 1968 led by James Tobin, Paul Samuelson, and John Kenneth Galbraith.
  23. Locke 1970/1690.
  24. https://www.marxists.org/reference/archive/smith-adam/works/moral/part02/part2b.htm (accessed 28 January 2016).
  25.  http://www.utm.edu/research/iep/b/bentham.htm (accessed 28 January 2016).
  26. It should also be noted that societies have long qualified this right, or hedged it in various ways, including the killing associated with war (or Jihad), capital punishment, euthanasia, killing in self-defense, etc.
  27. From a biological perspective, the purpose of life is to make more life. Reproduction is a basic need – a primary function in all living organisms. Of course, reproduction is also a strong human desire. As a practical matter, moreover, reproduction is essential to the continuity of society over time. If reproducing the next generation is ultimately a societal objective, the question then becomes, who shall contribute? There are only a finite number of options: everyone ad libitum, an equal number for everyone, a lottery, a “means test,” specific genetic and/or physiological qualities, “good behavior,” some form of social competition, or some other social qualification. In any case, producing the next generation is a basic need and a prime responsibility of the public trust, although currently overpopulation is a global problem. Although it’s still a third rail politically, there is a potential long term trade off that may be necessary – namely, global population control, or “mutual coercion mutually agreed upon” in Garrett Hardin’s memorable phrase. It is often forgotten that Hardin’s classic, widely reprinted and much debated article on “The Tragedy of the Commons,” in 1968, was focused on the problem of population control (see below). Although the so-called “demographic transition” — a tendency for population growth to slow with growing economic affluence – may have mitigated the problem, it remains uncertain as a long-term solution. See also https://en.wikipedia.org/wiki/Tragedy_of_the_commons (last modified 27 April 2016).
  28. Plato 1946/380 B.C.
  29. This is a very abbreviated version of a much longer discussion in my 2011 book, The Fair Society, Chapter Six. For a systematic critique of the neo-classical model, see Beinhocker 2006: Bowles 2006.
  30. The reality is, of course more complex. Wealth accumulation is a proven method for stimulating further innovation and entrepreneurship, but it also very often leads to conspicuous personal consumption, idled wealth, and other wasteful outcomes. Moreover, there is another alternative. Public taxes on wealth, if properly invested, can accomplish the same end.
  31. Smith 1964/1776, IV, 2, 9. Modern economists often become lyrical about “the superiority of self-interest” over altruism in economic life and the virtues of competition and the “profit motive,” while overlooking the fact that Smith’s rendering of the invisible hand was quite contingent. As he said, the invisible hand is not “always the worse” and “frequently promotes” the general welfare. But this is not a sure thing. Many of Smith’s acolytes also seem unaware of the cautionary warnings both in his masterwork and in his earlier work, The Theory of Moral Sentiments (1976/1759), where (as a Stoic and a Christian) he stressed the fact that everything in a free market depends on a moral foundation of trust, honest dealing and, as he himself put it, “justice”. Indeed, Smith was also a proponent of the Golden Rule.
  32. Mainstream economists might argue that utopian capitalism no longer reflects the orthodoxy that was once predominant. It’s certainly true that modern economics has become more heterodox and dynamic, with many new theoretical nodes. Some of these developments were discussed in Corning 2011, Chapter Four. An illuminating overview of this work can be found in the 2004 book by David Colander and his colleagues, The Changing Face of Economics: Conversations with Cutting Edge Economists.   Nevertheless, the orthodox neo-classical model retains a powerful hold on the discipline. See also Hodgson 2015.
  33. Bowles 2004, p. 208.
  34. Gowdy 1998, pp. xvi-xvii. Of course, many liberal economists over the years have challenged the neo-classical model, from John Maynard Keynes to Amartya Sen, James K. Galbraith, Samuel Bowles, Joseph Stiglitz , Geoffrey Hodgson, and others. Also important is the relatively new field of behavioral economics, a research enterprise that has been undermining the core neo-classical assumption about rational self-interest. One of the leaders of this new field, Richard Thaler, has called for a replacement of the Homo economicus model with Homo sapiens.
  35. From Corning 2011, Chapter Six.
  36. Picketty 2014/2013.
  37. Korten 2015/1995.
  38. See the insightful edited volume on this subject, Moral Markets: The Critical role of Values in the Economy, edited by Paul Zak 2008. See also Geoffrey Hodgson’s Conceptualizing Capitalism (2015); also, Beinhocker 2006.
  39. Pope Francis 2015. Papal Encyclical Laudato Si’ (“praise be to you”) p. 54. http://w2.vatican.va/content/francesco/en/encyclicals/documents/papa-francesco_20150524_enciclica-laudato-si.html (accessed 16 July 2015).
  40. Smith 1976/1759, Vol. I.i.5.5. See also note 4; also, Ritter 1954.
  41. Ibid., Vol. II.3.34.
  42. Ibid., Vol. VI.iii.II.
  43. Ibid., Vol. IV.i.10.
  44. In a nutshell, socialism is (in theory) cavalier about the Fair Society principle of equity – differential rewards (or punishments) for merit, and it is vague about the principle of reciprocity as well. See Corning 2011, especially Chapter Six.)
  45. An excellent general overview can be found in Kelly et al. 1997; also, Ackerman and Alstott 1999. Stakeholder capitalism as espoused by the philosopher/economist Edward Freeman (1984) and others has a strong moral cast to it. It is oriented especially toward giving workers and others a “voice” in the management of a company. Stakeholder capitalism as I envision it would have a more practical objective. It would be oriented to achieving a proper balance between the three basic social justice principles. There are no simple formulas for how to do so. It will always require a case-by-case approach and the spirit of compromise.
  46. In fact, the spirit of stakeholder capitalism long predates the term. Back in the 1950s, most American business firms endorsed a model of capitalism that was synonymous. For instance, the Chairman of Standard Oil of New Jersey, Frank Abrams, asserted that “The job of management is to maintain an equitable and working balance among the claims of the various interested groups…stockholders, employees, customers, and the public at large.” See Freeman 1984.
  47. Franklin Allen, Elena Carletti, and Robert Marquez, “Stakeholder Capitalism, Corporate Governance, and Firm Value.” (September 2009). http://fic.wharton.upenn.edu/fic/papers/09/0928.pdf. In a recent issue of The Economist (July 25, 2015), a “Schumpeter” column on “The enemy within” concluded that disgruntled employees are sometimes a more serious threat to a company than its competitors.
  48. See http://www.nytimes.com/2016/05/11/opinion/pressure-to-close-the-pay-gap.html?emc=edit_th_20160511&nl=todaysheadlines&nlid=27277596&_r=0 (accessed 11 May 2016).
  49. Government mandates, such as minimum wage legislation and regulations governing working hours, sick leave, etc., or the legal protections afforded by B-corporation status, also have the effect of leveling the playing field. Even Walmart, a firm that is notorious for paying poverty wages, must pay higher wages in jurisdictions that require it.
  50. Wood 2014. See also https://en.wikipedia.org/wiki/Public_trust_doctrine (last modified 11 September 2015).
  51. Both federal and state courts in the U.S. have also recognized the public trust in various rulings. Especially significant are the several landmark U.S. Supreme Court cases over the years in which the Court supported the doctrine. Indeed, between 1997 and May of 2008 the public trust doctrine was used in a total of 284 judicial decisions, including 34 federal cases and 250 in the states.   See David C. Slade, “The Public Trust Doctrine In Motion, State and Federal Cases 1997-2008.” International Submerged Lands Conference, The Evolving Public Trust Doctrine (September 24, 2009). www.gulfalliancetraining.com/…/Public%20Trust%20Doctrine%20In%2
  52. See https://www.washingtonpost.com/news/energy-environment/wp/2017/07/05/trashed/
  53.  Wood 2014, p. 132.
  54. Ibid., p. 128.
  55. Quoted in Wood, op cit., p. 128.
  56. Locke, 1970/1690.
  57. Wood, op. cit., p. 126.
  58. Brown 1994, p. 78.
  59. Cited in Wood, op. cit., pp. 129-130.
  60. Quoted in Brandon 1969.
  61. Norway has also used its profits to make many improvements in its infrastructure and public services, and it is now gradually winding down the oil sector of its economy.
  62. http://todayinsci.com/E/Einstein_Albert/EinsteinAlbert-PoliticsQuote500px.htm (accessed 28 January, 2016).
  63. Shriver 2016. For documentation, see Tainter 1988; Diamond 2005; Corning 2005, Chapter Seven. A new mathematical model reveals some of the underlying network dynamics. See Yu et al. 2016.
  64. Diamond 2005. As Diamond warns, “Globalization makes it impossible for modern societies to collapse in isolation” (p. 23).
  65. See the analysis in The Economist (August 1, 2015, pp. 12-13). It cites estimates by the International Atomic Energy Agency (IAEA) that the world will need to generate at least 11 times more wind power and 36 times more solar power by 2050. Among other recommendations, The Economist called for removing subsidies for fossil fuels world wide, imposing a gradually increasing carbon tax as an incentive for making a shift, and investing heavily in energy storage and transmission technologies. See also The Economist (Volume 424, Number 9049, 15 July 2017, p. 12); also   http://www.nytimes.com/2016/04/04/opinion/a-renewable-energy-boom.html?emc=edit_th_20160404&nl=todaysheadlines&nlid=27277596&_r=0
  66. See http://thinkprogress.org/climate/2016/05/09/3775606/used-second-life-electric-car-batteries/?utm_source=newsletter&utm_medium=email&utm_campaign=tptop3&utm_term=1&utm_content=53&elqTrackId=e7585cedc9b44b26950cb980701cc136&elq=f7aba5fa560e44ee861afaec283cfb5c&elqaid=30054&elqat=1&elqCampaignId=5581 (accessed 9 May 2016).
  67.  The Economist, December 19, 2015- January 1, 2016, p. 89.
  68. https://www.nytimes.com/2017/06/24/opinion/sunday/artificial-intelligence-economic-inequality.html
  69.  Battersby 2017.
  70. Hardin 1968.
  71. Ostrom 2009. Ostrom’s Eight Principles for effective common resource management have been used successfully in various situations. They are: 1. Define clear group boundaries; 2. Match rules governing use of common goods to local needs and conditions; 3. Ensure that those affected by the rules can participate in modifying the rules; 4. Make sure the rule-making rights of community members are respected by outside authorities; 5. Develop a system, carried out by community members, for monitoring members’ behavior; 6. Use graduated sanctions for rule violators; 7. Provide accessible, low-cost means for dispute resolution; 8. Build responsibility for governing the common resource in nested tiers from the lowest level up to the entire interconnected system.
  72. Of course, each of these historic acts of political creativity was built on foundation of precursor institutions – namely, the Articles of Confederation and the League of Nations. A similar foundation for global governance exists today in the United Nations.
  73. In an important article on how groups can evolve and prevail, David Sloan Wilson and Edward Wilson (2007) put it this way: “Selfishness beats altruism within groups. Altruistic groups beat selfish groups. Everything else is commentary.” A full-length elaboration on this thesis can be found in D.S. Wilson 2015.
  74. The problem of achieving global government has been debated by political scientists, and others, literally for generations, with opinions ranging from determinists who see it as “inevitable” (e.g. Wendt 2003) to those who view it as unlikely or even impossible (see Koenig-Archibugi 2010).   The moral case for it is argued by biologists David Sloan Wilson and Dag Olav Hessen in an online article on “Blueprint for the Global Village” https://evolution-institute.org/focus-article/blueprint-for-the-global-village/ (accessed 24 July 2016).
  75. D.S. Wilson 2015, p. 149.
  76. Jared Diamond (2005) asks why many societies historically have collapsed while others have endured. He finds that a common denominator is a failure to respond effectively, in any one or more of four ways, to a potentially catastrophic threat: a failure to anticipate the threat, a failure to perceive it when it becomes evident, a failure to act in response, or a failure to act effectively enough. Who is ultimately responsible for these failures? I would argue that the blame ultimately rests with those who have the power to inspire and lead, or to coerce effective societal change – a failure of leadership.  For a forward-looking article on developing a “science of intentional change,” see D.S. Wilson et al. 2014.
  77. Ehrlich and Ehrlich 2012.
  78. Lagi et al. 2015. Researchers at the New England Complex Systems Institute (NECSI), employing sophisticated modeling and analytical techniques, traced the root causes of these food price spikes to a combination of deregulated commodity markets, financial speculation, and the misguided corn-ethanol fuel program in the U.S., which removes some 5 billion bushels of corn from the market each year. It’s another tragic example of the law of unintended consequences. The authors of the study conclude: “A very strong social and political effort is necessary to counter the deregulation of commodities and reverse the growth of ethanol production. A concern for the distress of vulnerable populations around the world requires actions either of policymakers or directly of the public and other social and economic institutions.” Two other recent studies reinforce this climate-conflict nexus. See Schleussner et al. 2016; von Uexkull et al. 2016.
  79. From the documentary film, “E.O. Wilson, Of Ants and Men,” premiered on PBS 30 September 2015. http://www.pbs.org/program/eo-wilson/ (accessed 28 January 2016).
  80. It should also be noted that there is a school of contemporary theorists, including Robert Wright, Francis Heylighen, John Stewart, and others who portray evolution toward global cooperative organization as an inexorable trend. Stewart (2014), for example, sees a “large-scale” directionality in evolution that is “driven” by selection toward both increased diversification and increased “integration” (meaning more inclusive cooperative groups). This trend is likely to culminate, Stewart tells us, in a “global entity” – “a symbiotic community on the scale of the planet,” with all-encompassing goals and intelligence (a global brain). “An overall direction is unmistakable.” I side with those who see any such trend as an artifact of a cumulative, contingent process with myriad causes, not a result of what Stewart archly calls “entification”. See also Stewart 2000. To borrow a famous line from Theodosius Dobzhansky, the future is not vouchsafed by any law of nature. Extinction has been the rule in evolution.

 

Category: Publications