Holistic Darwinism: The New Evolutionary Paradigm And Some Implications for the Social Sciences

© Politics and the Life Sciences, 27(1), 22-54, 2008

ABSTRACT

Holistic Darwinism is a candidate name for a major paradigm shift that is currently underway in evolutionary biology and related disciplines.  Important developments include: (1) a growing appreciation for the fact that evolution is a multi-level process, from genes to ecosystems, and that interdependent “co-evolution” is a ubiquitous phenomenon in nature; (2) a revitalization of group selection theory, which was banned prematurely from evolutionary biology over 30 years ago (groups may in fact be important evolutionary units); (3) a growing respect for the fact that the genome is not a “bean bag” (in biologist Ernst Mayr’s caricature), much less a gladiatorial arena for competing  “selfish genes,” but a complex, interdependent, cooperating system; (4) an increased recognition that symbiosis is an important phenomenon in nature and that “symbiogenesis” is a major source of innovation in evolution; (5) an array of new, more advanced game theory models, which support the growing evidence that cooperation is commonplace in nature and not a rare exception; (6) new research and theoretical work that stresses the role of “nurture” in evolution, including developmental processes, phenotypic plasticity, social information transfer (culture), and especially the role of behavioral innovations as “pacemakers” of evolutionary change (e.g., niche construction theory, which is concerned with the active role of organisms in shaping the evolutionary process, and “gene-culture co-evolution theory,” which relates especially to the dynamics of human evolution); (7) and, not least, a broad effort to account for the evolution of biological complexity  from major transition theory to the Synergism Hypothesis. Here I will briefly review these developments and will present a case for the proposition that this paradigm shift has profound implications for the social sciences, including specifically political theory, economic theory and political science as a discipline. Interdependent “superorganisms,” it turns out, have played a major role in evolution – from eukaryotes to complex human societies.

INTRODUCTION

Thanks to the narrow, gene-centered focus of neo-Darwinism (the reigning paradigm in evolutionary biology for most of the past 40 years), not to mention biologist Edward O. Wilson’s imperialistic claims for sociobiology in the 1970s and the strident “genitis” of some evolutionary psychologists, I think it would be fair to say that most political scientists remain skeptical about the value of a Darwinian/biological approach to their discipline.  The only concessions might be at the margin, where some loose analogies can be found and, maybe, some consideration is given to genetic influences that affect the content of human nature or public policy issues.

This head-in-the-sand perception is outdated.  A major sea change has been occurring in evolutionary biology and related disciplines over the past decade that, I believe, also has profound implications for the underlying theoretical foundations of the social sciences, including our understanding of what are commonly defined as political behaviors.  This theoretical sea change in biology has contributed to the rise of such (relatively) new hyphenated interdisciplines as evolutionary anthropology, evolutionary economics, bioeconomics, human ethology and, not least, biopolitics.   I call this emerging new paradigm Holistic Darwinism (though it could also fairly be called Darwin’s Darwinism), and, as we shall see, it is much more compatible with the subject matter – and weltanschauung – of the social sciences than is generally appreciated.  We begin with the controversial issue of group selection.

THE PERILS OF GROUP SELECTION

The emotionally-charged group selection debate in biology — which celebrated an unofficial 40th anniversary in 2006 — provides a classic example of a controversy based largely on a misconception. To Darwin and many of his contemporaries, group selection was a perfectly respectable concept.  Indeed, it was Darwin who first proposed, in The Descent of Man (1874/1871), the then unexceptional idea that differential group selection may have played an important role in human evolution, along with what he called “family” selection (now known as inclusive fitness or kin selection theory) and individual reciprocities (now called mutualism and reciprocal altruism). Darwin’s “tripartite” explanation of human evolution was quite subtle, but his view of the role played by group selection is illuminated in this brief passage: “All that we know about savages, or may infer from their traditions and old monuments, the history of which is quite forgotten by the present inhabitants, show that from the remotest times successful tribes have supplanted other tribes” (p. 147).  Herbert Spencer, one of the outstanding theorists of the 19th century, expressed a similar view in The Principles of Sociology (1897/1874-82), and many of the pioneer anthropologists of that period also seemed to concur.

In the first half of this century, the founding fathers of modern genetics and population biology, notably including Haldane, Wright, Fisher, Morgan, Dobzhansky, and others (plus some non-geneticists like Huxley, Mayr, and Simpson) redefined evolutionary theory in quantitative genetic terms. However, the so-called “modern synthesis” was also deemed to be compatible with group selection of various kinds. For instance, Sewall Wright (1968-78) at the University of Chicago coined the term “interdemic selection” — i.e., selection between discrete breeding groups, or “demes” — and developed what he called a “shifting balance” model, which he believed was of the utmost importance in producing evolutionary changes. Ernst Mayr, likewise, speaks of evolutionary change as a population-level phenomenon, meaning that populations and species are the ultimate units of evolutionary change, not individuals. Mayr also developed what he calls the “founder principle,” which envisions small, reproductively isolated groups as a significant source of evolutionary innovation (Mayr 1963, 1976).  More recently, the paleontologist-popularizer Stephen Jay Gould (2002) championed a higher-level species selection paradigm. Meanwhile, various students of animal behavior, such as William Morton Wheeler and Warder C. Allee, stressed the cooperative aspect of animal behavior and social life. Wheeler (1927a,b) also promoted the idea of “emergent evolution,” and he borrowed from Spencer the idea that a socially-organized group can be likened to a “superorganism” (Wheeler 1928).

However, a theoretical “punctuated equilibrium” occurred in 1962.  In his subsequently much-maligned book Animal Dispersion in Relation to Social Behaviour, Vero C. Wynne-Edwards made himself a stalking horse, in Edward O. Wilson’s characterization, by propounding a seriously overstated version of the group selection hypothesis. Wynne-Edwards asserted that group-living animals regularly display behaviors that involve the curtailment of their own personal fitness for the good of the group (for example, through “conventional” controls on personal reproduction that serve to limit population densities). “The great benefit of sociality,” he claimed in a companion article in Nature (1963), “arises from its capacity to override the advantage of individual members in the interest of the survival of the group as a whole.” Some of Wynne-Edwards’s critics, playing loose with the facts, accused him of a Pollyanna-like naiveté that violated Darwinian theory, but in fact he clearly stated that altruistic, group-serving behaviors could arise only if natural selection were to operate between social groups “as evolutionary units.” Notwithstanding, Wynne-Edwards became a pariah in evolutionary biology and has been routinely chastised for his heresy ever since — rather like the treatment accorded to Lamarck.

Although the assault on group selection theory began with William D. Hamilton’s now classic papers on “The Genetical Evolution of Social Behavior” (1964a,b), it was fully-elaborated in George C. Williams’ New Testament — Adaptation and Natural Selection (1966). Williams’ near-legendary book was in many respects a therapeutic cold bath that served to purge evolutionary theory of some sloppy thinking. However, Williams also took an extreme position, from which he has since retreated, to the effect that selection at any higher level than that of an individual is essentially “impotent” and is “not an appreciable factor in evolution” (1966, p. 8; cf., Williams 1992).

Edward O. Wilson was more moderate by comparison in his discipline-defining volume, Sociobiology (1975), but he also (inadvertently) propagated a conceptual muddle that has caused no end of confusion and mischief in evolutionary theory.1 Wilson launched his massive synthesis with the startling assertion that altruism is “the central theoretical problem of sociobiology” (p. 3). The implication, which guided much subsequent work in this new interdiscipline, was that social life is founded on altruism. Therefore, cooperative behaviors are inherently a theoretical “problem” that can be overcome only under extraordinary circumstances — i.e., via group selection, kin selection and maybe Robert Trivers’s (1971) “reciprocal altruism.” In opposition to Wynne-Edwards, Wilson considered “pure” group selection — i.e., among non-kin — to be highly improbable, a rare occurrence confined to humans and perhaps a few other species. (His detailed, chapter-length discussion of group selection included a review both of the available evidence and of various formal models, but his conclusion was preordained by the assumption that “pure” group selection necessarily implied genetic altruism (1975.

Another broadside against group selection theory occurred when Richard Dawkins published his ideologically-tinged popularization with the cunningly anthropomorphic title The Selfish Gene (1989/1976). “I think ‘nature red in tooth and claw’ sums up our modern understanding of natural selection admirably,” Dawkins wrote with evident relish (1989/1976, p. 2). Not surprisingly, The Selfish Gene became a controversial best-seller. In retrospect, the selfish gene metaphor has proven to be a powerful heuristic tool. It has led to many new insights about the interactions within and among various functional units in nature and to much productive research. On the other hand, it also introduced a simplistic and seriously distorting perspective into evolutionary theory.

The short-term consequence of this rancorous theoretical debate was a wholesale rejection of the concept of group selection. Nevertheless, for the past 20 years or so David Sloan Wilson (lately with the collaboration of Elliott Sober and with parallel efforts from a growing number of other workers) has been attempting to resurrect group selection on a new foundation. What Wilson calls “trait group selection” (D.S. Wilson 1975, 1980; Wilson and Sober 1989, 1994; Sober and Wilson 1998) refers to a model in which there may be linkages (a “shared fate”) between two or more individuals (genotypes) in a randomly breeding population, such that the linkage between the two becomes a unit of differential survival and reproduction. Initially, Wilson assumed that one of the two was an altruist, for he was then intent on accounting for the evolution of altruism without recourse to kin selection. (John Maynard Smith developed a similar model, which he dubbed “synergistic selection.” See also Matessi and Jayakar 1976; Wade 1977, 1985; and the discussion in Dugatkin, et al., 1992.)

The current revival of group selection theory may perhaps be attributed, in considerable measure, to the growing recognition that it can also entail “win-win” processes. Cooperating groups might provide mutual advantages for their members, so that the net benefits to all participants outweigh the costs. In other words, cooperation is not equivalent to altruism and does not by definition require sacrifices, or genes for altruism. (I refer to it as “egoistic cooperation,” to distinguish it from altruism, and Maynard Smith has recently modified his usage of the term “synergistic selection” — originally associated with altruism — along the same lines.) This, in essence, is what game theory models of cooperation tacitly postulate (see Maynard Smith 1982b, 1984, 1989; Axelrod and Hamilton 1981, Axelrod 1984, inter alia.), which is why game theory formulations are largely indifferent to the degree of relatedness, if any, between the cooperators.  And game theory models of cooperation (as well as experimental research on the subject) have been growing exponentially over the past decade or so. (See especially Binmore 1994, 1998; Sigmund 1995; Gintis 2000a; Stephens et al., 2002; Sigmund et al., 2002; Bowles et al., 2003).

Moreover, game theory provides a window into a vastly larger galaxy of cooperative phenomena that, I submit, reduces the group selection controversy to a tempest in a teapot. This alternative formulation was originally developed in The Synergism Hypothesis: A Theory of Progressive Evolution (1983; also Corning 2003a, 2005).  It was also developed independently by Maynard Smith and Szathmáry (1995, 1999), and it is supported by an accumulating body of research findings across many different specialized disciplines, from molecular biology and microbiology to behavioral ecology, primatology and sociobiology — not to mention the social sciences. This alternative paradigm can be characterized as “Holistic Darwinism.” (See also Dugatkin and Reeve, 1994, and Dugatkin and Mesterton-Gibbons, 1996, on indirect “by-product mutualism” in evolution and Wilson and Dugatkin, 1997, on the role of “assortative interactions,” or behavioral selection, as a mechanism of group selection.)

HOLISTIC DARWINISM DEFINED

Holistic Darwinism is not an oxymoron. The term was coined as a way of highlighting the paradox that selfish genes are, without exception, “selected” in the context of their functional consequences (if any) for various wholes.  Holistic Darwinism is strictly Darwinian in its underlying assumptions about natural selection and the evolutionary process. It has no fundamental quarrel with the theoretical premise of “gene selfishness.” Rather, it involves a different perspective on the causal dynamics of evolution. In his preface to the second edition of The Selfish Gene, Dawkins uses the metaphor of a Necker cube — a two-dimensional drawing of a three-dimensional object that can be perceived in different ways — to characterize the intent behind his inspired metaphor: “My point was that there are two ways of looking at natural selection, the gene’s angle and that of the individual…It is a different way of seeing, not a different theory” (1989/1976, pp. x-xi).

Actually, there are more than two ways of looking at natural selection, and Holistic Darwinism focuses not on genes, or individuals, or even groups as units of selection but on the functional relationships among the “units” at various levels of biological organization, from genomes to ecosystems, and on their consequences for differential survival and reproduction. It involves refocusing the Necker cube on the “interactions” between genes, between cells, between organisms, and between organisms and their environment(s).  Perforce, Holistic Darwinism is also about the role of synergy — the combined effects produced by phenomena that “co-operate” (operate together) — as a major cause of evolutionary continuity and change.

It should be stressed at the outset that the term cooperation will be used here in a strictly functional sense; it refers to functional interactions. In this conceptualization, cooperation may or may not also be considered selfish or altruistic, mutualistic or parasitic, positive or negative. Such attributes involve additional, post-hoc judgments about the consequences of a cooperative relationship with respect to some separately specified goal or value. (Of course, in Darwinian theory the operative value is survival and reproductive success.) By the same token, a cooperative relationship may or may not be voluntary. Slavery, in nature and in human societies alike, involves a form of involuntary cooperation, and so (presumably) does the host’s role in a parasitic relationship.

Accordingly, a key point about cooperation as a functional concept is that it is found at every level of living systems. Beginning with the very origins of life, it is a common denominator in all of the various formal hypotheses about the earliest steps in the evolutionary process (reviewed in Corning 1996a). All share the common assumption that cooperative interactions among various component parts played a central role in catalyzing living systems.

Similarly, at the level of the genome, it goes without saying that genes do not act alone, even when major single-gene effects are involved. Indeed, the human genome sequencing project has established, among many other things, that there are in fact 1,195 distinctive genes associated with the human heart, 2,164 with white blood cells and 3,195 with the human brain (see Little 1995). The functional (morphogenetic) implications behind those numbers are awesome to contemplate.  As Richard Dawkins himself so eloquently put it in a later book, The Blind Watchmaker (1987/1986):

“In a sense, the whole process of embryonic development can be looked upon as a cooperative venture, jointly run by thousands of genes together. Embryos are put together by all the working genes in the developing organism, in collaboration with one another….We have a picture of teams of genes all evolving toward cooperative solutions to problems…It is the ‘team’ that evolves” (pp. 170,171).

The origin of chromosomes, likewise, may have involved a cooperative/symbiotic process (see Maynard Smith and Szathmáry 1993). Sexual reproduction, one of the major outstanding puzzles in evolutionary theory, is also a cooperative phenomenon, as the term is used here. Although there is still great uncertainty about the precise nature of the benefits, it is assumed that sexual reproduction is, by and large, a mutually beneficial joint venture.

As one moves upward in “the great chain of being” (to borrow that quaint anachronism), one finds further variations on the theme of functional cooperation.  Once upon a time bacteria were considered to be mostly loners, but no longer. It is now recognized that large-scale, sophisticated cooperative efforts — complete with a division of labor — are commonplace among bacteria and can be traced back at least to the origin of the so-called stromatolites (rocky mineral deposits) that, it is believed, were first constructed by bacterial colonies some 3.5 billion years ago (Shapiro 1988; Shapiro and Dworkin 1997; Margulis 1993; Bloom 1997). Shapiro suggests that bacterial colonies can be likened to multicellular organisms.

Eukaryotic cells can also be characterized as cooperative ventures — obligate federations that may have originated as symbiotic unions (parasitic, predatory or perhaps mutualistic) between ancient prokaryote hosts and what have now become cytoplasmic organelles, particularly the mitochondria, the chloroplasts and, possibly, eukaryotic undulipodia (cilia) and certain internal structures that may have evolved from structurally-similar spirochete ancestors (Margulis 1993). The phenomenon of symbiosis, by definition a category of cooperative relationships in nature, provides yet another example. Not only has the darker side of symbiosis   — parasitism — gained new prominence over the past decade or so but more benign commensalistic and mutualistic forms of symbiosis are also more widely appreciated (see below).

Sociobiology is also, by definition, concerned with cooperative relationships among conspecifics, interactions which can provide a variety of adaptive consequences for the participants. As shown by the many field studies and laboratory experiments that were inspired by inclusive fitness theory and game theory, the social interactions that occur in nature among members of the same species may be perturbed by free-riders, “defectors”, exploiters, conspecific parasites, etc., yet the fact remains that within-species cooperative behaviors are fairly common and encompass a broad array of survival-related functions, including: (1) hunting and foraging collaboratively, which may serve to increase capture efficiency, the size of the prey that can be pursued, or the likelihood of finding food patches; (2) joint detection, avoidance of and defense against predators, the forms of which range from mobbing and other kinds of coordinated attacks to flocking, herding, communal nesting and synchronized reproduction; (3) shared protection of jointly acquired food caches, notably among many insects and some birds; (5) cooperative movement and migration, including the use of formations that increase aerodynamic or hydrodynamic efficiency and reduce individual energy costs and/or facilitate navigation; (6) cooperation in reproduction, which can include joint nest-building, joint feeding and joint protection of the young; and (7) shared environmental conditioning.

Neo-Darwinian theory — as purified by the selfish gene perspective — attributes evolutionary change to competition among the “replicators” — the ultimate units of information transfer in evolution.  In the classical neo-Darwinian model, cooperation plays a decidedly subsidiary role. But if we shift our perspective and view evolution as an ecological and economic process — a “survival enterprise” in which living systems and their replicators are embedded — then differential reproductive success may be viewed as the result of a complex interplay of competitive and cooperative interactions (along with a variety of other factors), both within and among functionally interdependent units of ecological interaction. Our focus shifts to the activities of the “vehicles” (in Richard Dawkins’s terminology) or the “interactors” (in the terminology of David Hull, 1980) — and, more important, to the bioeconomic consequences of their functional interactions. (In short, we are now in the realm of the social sciences.)

It has been a cardinal assumption of neo-Darwinism that cooperation in nature is a  phenomenon that is at odds with the basic principle of gene competition, and that extraordinary conditions are required to overcome the inherent selective bias against the evolution of cooperation. This assumption is what accounts for the importance attached to inclusive fitness theory (or kin selection, in Maynard Smith’s term) and to game theory. However, a functional/ bioeconomic perspective on the evolutionary process challenges that point of view. Not only is cooperation (broadly defined) fairly common in nature but synergistic effects (the functional consequences of cooperation), it is argued, have played an important causal role in evolution, especially in relation to the evolution of complexity. To put it baldly, functional synergy explains the evolution of cooperation in nature, not the other way around.  In other words, functional groups (in the sense of functionally integrated “teams” of cooperators of various kinds) have been important units of evolutionary change at all levels of biological organization; “functional group selection” is thus a ubiquitous aspect of the evolutionary process.2 This is obviously a highly contentious assertion. So let me briefly summarize the evidence.

CONSIDER THE EVIDENCE

If cooperation in nature is not largely dependent on inclusive fitness, we would expect to find a significant degree of decoupling in the natural world between genetic relatedness and cooperation, and, in fact, there are at least four sources of evidence for this proposition. First, there is the entire domain of symbioses. Here we can observe a wide range of cooperative relationships that can only be accounted for in bioeconomic, cost-benefit terms. Kinship is largely irrelevant. Indeed, many types of symbioses, such as the estimated 20,000 species of lichen partnerships involving approximately 300 different genera of fungi, or the Rhizobium-like bacteria that form root nodules with some 17,500 species in 600 genera of plants, reflect a plethora of independent inventions. In other words, many different species may discover and utilize the same functionally-advantageous cooperative relationships. As Maynard Smith, 1989, has noted, extreme non-specificity is the rule among mutualists, whereas parasitism is highly specific. The case for “symbiogenesis” as a significant factor in evolution was documented by participants at a 1989 conference on the subject and in a subsequent volume edited by Margulis and Fester (1991).  (We will have more to say about symbiogenesis later on.)  Among the extensive evidence that was presented at the conference:

  • Mutualistic or commensalistic associations (not to mention parasitism) exist in all five “kingdoms” of organisms, as defined by Whittaker and modified by Margulis and Schwartz (1982). Most extant species may, in fact, be either a product of or currently involved in (or both) endo- or ecto-symbioses. Elsewhere, Bermudes and Margulis (1987) documented that 27 of 75 phyla in the four eukaryotic kingdoms (or 37%) exhibit symbiotic relationships.Silurian and Devonian plant fossils have been found to contain structures closely resembling the symbiotic “vesicles” produced by modern VAM (mycorrhizal) fungi (Smith and Douglas 1987), and over 90% of all modern land plants establish mycorrhizal associations (Lewis 1991).
  • Land plants may have arisen through a merger between fungal and algal genomes, as sort of inside-out lichens. In any case, it is evident that modern land plants represent a joint venture between fungi and green algae (Pirozynski and Malloch 1975; Atsatt 1988, 1991).
  • Approximately one-third of all known fungi are involved in mutualistic symbioses (e.g., lichens), many of which have conferred on their partnerships the ability to colonize environments that would not otherwise have been accessible to them (Kendrick 1991).
  • Virtually all species of ruminants, including some 2,000 termites, 10,000 wood-boring beetles and 200 Artiodactyla (deer, camels, antelope, etc.,) are dependent upon endoparasitic bacteria, protoctists or fungi for the breakdown of plant cellulose into usable cellulases (Price 1991).
  • Within the teeming communities of organisms that have recently been discovered in proximity to various sea floor hydrothermal vents, there are a number of symbiotic partnerships between chemoautotrophic (sulfur-oxidizing) bacteria and various invertebrates, which rely on the bacteria for their carbon and energy requirements (Vetter 1991).
  • Most bacterial cells congregate and reproduce in large, mixed colonies with many endosymbionts (virus-like plasmids and prophages) and ectosymbionts (metabolically complementary bacterial strains). These congregations call into question the classical notion of a species, in the sense of competitive exclusion and reproductive isolation (Sonea 1991; also Shapiro 1988; Shapiro and Dworkin 1997).

A second body of supporting evidence can be found in the various game theoretic models of cooperation between unrelated individuals, along with the substantial research literature that these models have inspired. (These will be discussed further below.) Third, there is the entire category of outbreeding reproduction, a class of cooperative behaviors which, by definition, falls outside of the inclusive fitness model. Finally, over the past decade or so there have been many field and laboratory studies of cooperation among conspecifics that are inconsistent with inclusive fitness theory and/or suggest that the particular behaviors in question are more satisfactorily explained in bioeconomic terms, although cooperation remains more likely to occur in closely-related, or at least familiar, animals.

A detailed summary of this discordant evidence (including 28 recent field and laboratory studies and seven reviews of the older literature) can be found in Corning (1996a) (see also the careful analysis by Goodnight and Stevens 1997). One particularly well-documented illustration is the food-sharing behavior among vampire bats (Desmodus rotundus), which clearly demonstrates the power of functional/bioeconomic factors to transcend the influence of genetic relatedness in shaping cooperative behaviors (Wilkinson 1984,1988, 1990). If gene competition were of overriding importance, the sharing of blood among vampire bats (their exclusive food-source) would be confined to close relatives. The reason is that blood sharing in this species has very high fitness value; an individual bat that fails to feed for two nights in a row will die. In field studies as well as controlled observations in captive groups over a ten-year period, Wilkinson found that blood-sharing both between relatives (matrilines) and non-relatives was extensive. Both relatedness and prior association proved to be important facilitators. Moreover, quantitative cost-benefit analyses showed that the cost to donors was relatively low (in effect, they were sharing their surpluses), while the fitness benefits to recipients was relatively high. When this was combined with the fact that the donors’ generosity was usually reciprocated later (i.e., “reciprocal altruism” sensu Trivers, 1971, 1985), there was a significant increase in the mutualists’ joint fitness. Wilkinson concludes: “Reciprocity is likely to be more beneficial than kin selection — provided that cheaters can be detected and excluded from the system” (1990, p.82).  (For a more recent example of non-kin cooperation, in red-winged black birds, see Olendorf et al., 2004.)

Two themes stand out in the many other examples that are described in Corning (1996a, and 2003): (1) the importance of bioeconomic cost-benefit considerations in cooperative relationships and (2) the presence of synergy — combined functional effects (payoffs) which are jointly produced and provide benefits to the cooperators that are greater than would otherwise be possible. As Maynard Smith and Szathmáry put it in The Major Transitions in Evolution (1995), if an individual can produce two offspring on its own but by cooperating in a group consisting of “n” individuals can produce “3n” offspring, it pays to cooperate. (An application of this perspective to avian species can be found in Emlen 1996.)

GAME THEORY REVISITED

Game theory models of cooperation, viewed in the proper light, are also consistent with Holistic Darwinism. Game theory suggests that the evolution of cooperative behaviors depends on an appropriate set of strategic circumstances, not genetic relationships. Although the focus has always been on the behavioral context and the strategies of the “players,” if one looks closely at the various game theory formalizations they tacitly depend on an interaction between the behavior of the players and the structure of the payoff matrix. And if one looks closely at the payoff matrices in some of the “classic” formulations, like Tit-For-Tat, the cooperative strategies in turn depend on synergy. In Axelrod and Hamilton’s (1981) model, mutual defection yielded one point each; asymmetrical cooperation (parasitism?) yielded 5 points for the “defector” and none for the cooperator; and mutual cooperation yielded a total of six points, evenly divided. Furthermore, defectors would be penalized in subsequent “rounds” (it was conceived as an iterated game) so that mutual cooperation becomes an increasingly rewarding option over time. In effect, this amounts to a quantification of synergy; the implicit economic benefits of the game are critically important.

But what about “cheating” or “defection” (the Prisoner’s Dilemma)? Maynard Smith and Szathmáry (1995) have proposed a response in terms of game theory, as illustrated in the two diagrams below. (I have taken the liberty of revising the payoff values that were utilized by Maynard Smith and Szathmáry to accord with a more explicit assumption about the object of the game, namely, that the oarsmen are both seeking to cross a river.) The left-hand figure in the diagram below involves a “sculling” model in which two oarsmen each have a pair of oars and row in tandem. In this situation, it is easy for one oarsman to slack off and let the other one do the heavy work. This corresponds to the classical two-person game.  However, in a two-person “rowing” model, each oarsman has only one opposing oar. Now their relationship to the performance of the boat is interdependent.  If one oarsman slacks off, the boat will go in circles. In this case, mutual cooperation becomes an evolutionarily stable strategy and defection is totally unrewarding; in the absence of teamwork, the boat will not reach its goal.

((Figure I))

Maynard Smith and Szathmáry conclude that the rowing model is a better representation of how cooperation evolves in nature: “The intellectual fascination of the Prisoner’s Dilemma game may have led us to overestimate its evolutionary importance” (1995, p. 261). Indeed, as Peck (1993, p. 195) observed: “The position of [stable] equilibria (and hence the frequency of cooperators) depends on the size of the various payoffs that define the Prisoner’s Dilemma game.”  (See also Dugatkin, et al., 1992; Brembs 1996.)

AN EVOLUTIONARY THEORY OF GOVERNMENT

If many forms of cooperation are functionally interdependent and thus self-policing, many more are not. The problems of cheating, defection, and “free-riders” — phenomena that the selfish gene metaphor has helped to illuminate — are real. But, in retrospect the problem may have loomed much larger in theory than it does in fact; our models may have been too pessimistic about the constraints on errant behavior in cooperative relationships. In effect, the games may have been unintentionally “rigged.” Consider some of the common assumptions in classical two-person games: The games are always voluntary and democratic; each player is free to choose his/her own preferred strategy, and the opposing player has no means available for coercing choices, or compliance. Also, the players are not allowed to communicate with one another in an effort to reduce the uncertainties in the interactions. Furthermore, defectors are usually rewarded handsomely for cheating while the cooperators are denied the power to prevent defectors from enjoying the rewards, much less punishing them for defection. Such “grade inflation” for defection biases the game in favor of cheating. Worse yet, in iterative games the players are forced to continue playing; they cannot exclude or ostracize a defector. They can only retaliate by themselves defecting and hoping thereby to penalize the other player (see also Binmore 2004).

A tacit rebuttal to this formulation was incorporated into a new kind of Prisoner’s Dilemma model developed by Nowak and Sigmund (1993) called “Pavlov,” which the authors suggested can outperform Tit-For-Tat. They called their strategy “win-stay, lose-shift,” and the significance of this innovation is that, in contrast with an iterated game in which the players must continue playing regardless of the outcome, in Pavlov they have the choice of leaving the game if they don’t like the results. In other words, a player may also have the power to exercise some control over the behavior of a defector by denying to that player future access to the game and its potential benefits. Punishments as well as rewards may be utilized as a means of keeping the game honest and, more important, as a means of restricting the game over time to mutual cooperators.

In addition to such suggestive formalizations, there is increasing evidence that a policing function does in fact exist in nature (among the outpouring of publications on this subject, see especially Boyd and Richerson 1992; Clutton-Brock and Parker 1995; Frank 1995,1996; Michod 1996; Fehr and Gächter 2000a,b, 2002; Gintis 2000b; Axelrod 2001, Falk et al., 2001; Henrich and Boyd 2001; Bowles and Gintis 2002; Boyd et al., 2003; Gintis et al., 2003; Binmore 2004).  As Clutton-Brock and Parker point out in the summary of their review article on the subject: “In social animals, retaliatory aggression is common. Individuals often punish other group members that infringe their interests, and punishments can cause subordinates to desist from behaviour likely to reduce the fitness of dominant animals. Punishing strategies are used to establish and maintain dominance relationships, to discourage parasites and cheats, to discipline offspring or prospective sexual partners and to maintain cooperative behaviour” (1995, p. 209). Evidence of a policing function has also been documented in social insects (Ratnieks and Visscher 1989), naked mole-rats (Sherman, et al., 1991), primates (de Waal 1996) and, needless to say, Homo sapiens, among others.

From a functional (synergy) perspective, if cooperation offers sufficient benefits it may be in the interest of some individuals to invest in coercing the cooperation of others. Inclusive fitness provides one possible explanation for punishment as a successful strategy in social groups. Another might be the sort of individual fitness tradeoffs referred to above. But group selection may also provide a mechanism. The enforcement of cooperation might have significant fitness-enhancing value for groups that are in competition with other groups, or other species. Maynard Smith’s (1982, 1983, 1989) “synergistic selection” model is relevant here.  The model suggests that, if cooperative interactions among two or more individuals — related or unrelated — produces selectively advantageous synergistic effects for all parties (on average), the cooperating “players” may become a unit of selection. A synergistic functional group might be favored in competition with other groups, or with ecological competitors from other species, or with the statistical probability of their survival and reproduction in the absence of cooperation. More broadly, synergistic selection can be defined in terms of gene combinations that enable/induce synergistic functional effects at various levels of biological organization. (For a model related to the multicellular level, see Michod 1996.)

SYNERGISTIC SELECTION

The concept of functional group selection, or synergistic selection, can be illustrated by returning to Maynard Smith and Szathmáry’s sculling and rowing models, as described above. What if the object of the game were changed? Rather than merely crossing a river (say), now the two oarsmen in each boat share the objective of winning a race against the other boat. Now it has become a functional group selection game (see Figure II). In this situation, if either oarsman were to defect, their “team” might lose the race; only all-out cooperation would provide rewards for either player. (Note that the two payoff matrices are now identical.) Now the sculling and the rowing games are functionally equivalent in the sense that the performance of either boat depends upon both oarsmen; they have both become “functional groups”; there is “synergistic selection.” Furthermore, it is irrelevant whether or not the oarsmen are related.

((Figure II))

Below are a few specific examples of synergistic selection:

  • In insects, Page and Robinson (1991) conducted an analysis of their own and other researchers’ data on the division of labor in honey bees, including a number of computer simulations, and concluded that natural selection operated on colony-level parameters. Oldroyd, et al. (1992a,b) also studied the genetics of honey bee colonies and concluded that colony performance was also influenced by the interactions among subfamilies, a colony-level parameter. Fewell and Winston (1992) conducted a study that examined the relationship between pollen storage levels in honey bee colonies (a group-level parameter) and individual forager efforts; not only was the correlation strong, but the researchers detected evidence of a homeostatic “set point.” And Guzmán-Novoa, et al. (1994) reported on a study that was focused on the relationship between colony-level natural selection and the level of effort associated with various components of the division of labor in honey bee colonies (see also Calderone and Page 1992).
  • An older study by Hoogland and Sherman (1976) examined in detail the influence of six possible disadvantages and three potential advantages of colonial nesting in 54 colonies of the Bank Swallow (Riparia), ranging in size from 2 to 451 members. Hoogland and Sherman concluded that the disadvantages were not very burdensome and, more important, that the maintenance of coloniality was most strongly associated with group-level defensive measures, which differentially benefitted the larger colonies. Although potential predators were not more frequent visitors to large groups, they were detected much more quickly and were mobbed by greater numbers of defenders; predators were also subject to more vocal commotion; and, bottom line, larger colonies were more effective overall in deterring predators.
  • Scheel and Packer (1991), in a study of female African lions, found that the average degree of relatedness among the animals had no bearing on their propensity to engage in group hunting. The key variable was the potential for synergy; successful hunting of larger prey required group hunting. And in a separate study by Packer, et al. (1990), it was concluded that the dynamics of female lion grouping were also strongly influenced by the need to defend their cubs (often a group-level function) and to compete against neighboring prides. In both situations, larger groups had an advantage.
  • Maynard Smith illustrated his 1982 article on synergistic selection with, among others, the examples of orb-web spiders (Metabus gravidus), where groups of 15-20 females may cooperate in building a joint web to span a stream where prey are abundant, tropical wasps (Metapolybia aztecoides) that establish joint nests, and coalitions of lion males that cooperate in taking over and holding a pride.
  • Finally, a recent study of spotted hyenas (Crocuta crocuta) by Russel Van Horn and his associates showed that, contrary to kin selection theory, individual matrilines commonly aggregate into larger clans of unrelated groups when confronted with dangerous competitors B including even matrilines that are closely related! (Van Horn 2004).  (It should also be noted that Wilson and Sober, 1994, in an in-depth target article on the subject, provide a compendium of over 200 references on group selection, of which 35 are identified as field or laboratory research efforts. See also the in-depth study of group selection in social bees in Moritz and Southwick 1992.)

Closely related to the notion of functional group selection, or synergistic selection, as an evolutionary phenomenon is the concept of “downward causation.” The term was actually coined by psychobiologist Roger Sperry (1969, 1991, inter alia) in connection with the functional organization and operation of the human brain — i.e., cybernetic control processes. (It may be  that psychologist Donald Campbell, 1974, developed the concept independently.) Sperry was fond of using as an illustration the metaphor of a wheel rolling down hill; its rim, all of its spokes, indeed all of its atoms, are compelled to go along for the ride.

We will use the term here in a slightly different sense. Downward causation in this context refers to the selective influences that have shaped the evolution of cooperative phenomena generally and complexity in particular. Why do selfish genes cooperate in ways that produce teamwork which, in turn, leads to interdependency? What compels them to subordinate their interests to the interests of the “whole”? To be specific, how did morphological castes and a division of labor evolve in army ants? How do reproductive controls evolve in mutualistic symbioses where, as Margulis (1993) points out, there must of necessity be reproductive synchronization if the relationship is to remain stable? (See also the discussion of “sociogenesis” in E.O. Wilson 1985; also Buss 1987; Smith 1992; and the examples cited in Leigh 1991.) Equally important, how can the potential for cheating among selfish genes (or selfish individuals) be constrained?

Downward causation in an evolutionary context refers to the fact that the functional (synergistic) properties of the whole become a selective “screen” — a significant influence on the differential survival/reproduction of the parts. Sometimes the parts might be disadvantaged (e.g., non-reproductive workers), and kin selection may help us to understand how such sacrifices for the common good may occur. But, as the evidence cited above indicates, kinship is not a sine qua non. The whole may also be sustained by fitness tradeoffs; that is, the costs may be offset by commensurate benefits. For instance, an animal that is at risk from predators might suffer a reduction in its relative reproductive fitness in a social group setting, but it may also enjoy greatly enhanced odds of survival and absolute fitness. (This may help to explain why defeated contenders for breeding privileges sometimes stay on in the group and may even serve as helpers.) To quote the born-again Dawkins once more: “In natural selection, genes are always selected for their capacity to flourish in the environment in which they find themselves…But from each gene’s point of view, perhaps the most important part of its environment is all the other genes that it encounters [his emphasis]…Doing well in such environments will turn out to be equivalent to <collaborating’ with these other genes” (1987/1986, p. 170,171).

In some cases, the whole may represent an unalloyed benefit for the parts with little or no costs.  Many cases of mutualistic symbioses seem to fit into this category. For instance, Margulis (1993) is adamant about the cooperativeness, promiscuity (and evolutionary significance) of bacterial colonies. (See also the parallel argument of Shapiro, 1988.) Thus, an isolated bacterium would be cut off from access both to the extensive gene swapping and the collective environmental “intelligence” (information) that commonly exists in bacterial colonies, not to mention the advantages of a division of labor and various collaborative efforts.  Social mammals may also exhibit many of these “higher level” properties.  Some of the most compelling recent field research has illuminated the surprisingly sophisticated social organization, mutualism and even culture in whales and dolphins (see especially Würsig 1988, 1989; Mann et al., 2000; Gygax 2002; Yurk et al., 2002; Whitehead and Rendell 2004).  Conversely, the power of a social group to isolate or ostracize a free-rider can be a significant deterrent and an agency of negative (downward) selection.

In any case, the synergies that result from cooperation may selectively “reinforce” cooperative behavior (to use the terminology of behaviorist psychology), and this may in turn differentially favor the evolution of relevant morphological/psychological characters over time. Thus, army ant sub-majors have acquired anatomical specializations that facilitate their role as “porters,” and humans have evolved psychological predispositions that help us to orchestrate (and even enjoy) our participation in group activities.

In sum, the relevant factors for explaining cooperative phenomena in nature (and in human societies) may include genetic relatedness, but kinship is neither necessary nor sufficient. The key lies in functional synergy and its bioeconomic consequences for differential survival and reproduction in a specific context; functional synergy is the frequently unappreciated common denominator in various models of cooperative behavior in sociobiology.

FROM “EVO-DEVO” TO THE BALDWIN EFFECT

Four current theoretical/research developments should also be highlighted briefly.  One is the work on what has been variously called developmental systems theory (Oyama et al., 2000), phenotypic plasticity theory (Rollo 1995; Pigliucci 2001; West-Eberhard 2003) and, simply, evolutionary development theory, or evo-devo (various authors but see especially Gould 2002). In effect, this movement represents an effort to meld the traditional, gene-centered evolutionary theory with the expanding body of evidence that developmental processes – which involve an inextricable interaction between an organism and its environment(s) – can also be an important source of evolutionary novelties – innovations, modifications or deletions.

Some adherents claim this is the path to a new evolutionary synthesis, much like the “modern synthesis” of the 1930s, which melded Mendelian genetics and Darwinian theory.   However, this claim falls short because it excludes another important development in evolutionary theory – a renewed focus on the role of behavior, and mind as an innovative agency in evolution.  The roots of this idea can be traced back to Jean Baptiste de Lamarck in his Zoological Philosophy (1963/1809).  Lamarck argued that, in the evolutionary process, changes in an animal’s habits often come first and that morphological changes may then follow.  This phenomenon – renamed the Baldwin Effect (after a turn-of-the-century advocate, psychologist James Mark Baldwin) by paleontologist George Gaylord Simpson (1953) – has gained an increasing number of adherents over the years. (Among others, see Roe and Simpson 1958; Mayr 1960; Corning 1983, 2003a, 2005; Bateson 1988; Plotkin 1988; Avital and Jablonka 1994, 2000; Deacon 1997; Weber and Depew 2003.)  I prefer to call this “mechanism” neo-Lamarckian Selection, in honor of the theorist who first recognized its importance.  But, under any name, behavioral innovations have often been the pacemakers of evolutionary change, in  Mayr’s (1960) terminology.

Another convergent trend in evolutionary theory involves the important work by Odling-Smee and his colleagues on niche construction theory (Laland et al., 2001; Odling-Smee et al., 2003).  These models, and the supporting evidence, show clearly that living organisms at all levels are not passive recipients of environmental conditions but actively shape their environments -and even the entire biosphere – to suit their needs.

Finally, there is a rapidly expanding body of work – most relevant to human evolution but not exclusively so – that is generally referred to as gene-culture co-evolution theory.  Theoretical work on this goes back to the 1980s (see especially Cavalli-Sforza and Feldman 1981; Corning 1983; Boyd and Richerson 1985; Durham 1991).  However, the last few years have seen a rapid increase in empirical work that is supportive of this paradigm, along with further theoretical refinements.  This has been stimulated in part by a growing recognition that culture (broadly defined as the social transmission of adaptive behavioral information) also exists in other species – from songbirds to cetaceans and, especially, our primate relatives.  (For more extensive discussions, see Corning 2003a, 2005; Hammerstein 2003; Richerson and Boyd 2004.)  Indeed, a case can be made for the view that, in some species (most notably humans), it might more accurately be called culture-gene co-evolution theory.

THE “SELFISH GENOME”

All of this rampant holism begs a question, however. Do wholes have goals that transcend the goals of the parts? Can wholes come to exercise a degree of autonomous control as wholes? In other words, can we postulate a “selfish genome?” The neo-Darwinian response, it appears, is a somewhat equivocal “no”. Richard Dawkins (1989/1976) became famous for the assertion that organisms are merely “robot vehicles” that have been blindly programmed to serve the interests of the genes, yet (as noted earlier) he also allowed that genes can be selected for their ability to serve the interests of the team. And George Williams (1966), while acknowledging the wholeness and unity of organisms, characterized many of the claims regarding superorganisms as figments of a “romantic imagination” (p. 220). In truth, some of these superorganismic claims were inflated, but Williams’s view of this issue was perhaps a bit too-jaundiced: “A wolf can live on elk only when it attacks its prey in the company of other wolves with similar dietary tendencies. I am not aware, however, of any evidence of functional organization of wolf packs” (pp. 217-218).

In contrast, Holistic Darwinism postulates that wholes at various levels of biological organization may evolve mechanisms that permit partially-autonomous control over the parts and their actions. Some insight into how superordinate controls can evolve in nature is provided in Egbert Leigh’s various discussions of how groups might act to contain or override individual advantages for the good of the group — what he calls the “parliament of the genes” (Leigh 1971, 1977, 1983, 1991; see also Michod 1996, 1997, 1999; Frank 2003; Rainey and Rainey 2003). Leigh’s argument, in essence, is that, if the potential payoffs (synergies) for each of the participants in a cooperative relationship are high enough, this could also provide an incentive for the imposition of “government” in the “public interest.” Leigh even draws on Adam Smith’s reasoning, not from The Wealth of Nations (1964/1776) but from the less well-known The Theory of Moral Sentiments (1976/1759). Although it is not widely appreciated, Smith argued for the necessity of a system of laws and appropriate means of enforcement in human societies to resist the dangers of unfettered self-interest. (pp. 86, 88-89, 340-341).

A key to understanding the evolution of “government” at various levels of biological organization may lie in what could be called the “paradox of dependency.” Although cooperative interactions may produce individual fitness-enhancing synergies, a tradeoff may be that the more valuable the benefits the more likely it is that the parts will become dependent upon the whole. As the benefits of cooperation increase, so may the costs of not cooperating. Wholes may then become obligatory survival units, one consequence of which may be that a decrement in the performance of the whole might result in the demise of the parts, and vice versa. An example can be found in a long-term study by Jeon (1972, 1983). A strain of Amoeba proteus was initially infected with bacterial parasites that were resistant to the hosts’ digestive enzymes. After 200 generations, or 18 months, a mutualistic relationship had become established, and after 10 years the symbionts had developed complete interdependence. (Jeon, 1992, has also illuminated some of the biochemistry associated with these changes.) It should also be noted that Margulis (1993) makes a similar argument with respect to the integration of symbiotic organelles in the ancestral eukaryotic cells. An obvious implication is that the incentives (both proximate and ultimate) for imposing government over the parts are likely to increase in relation to the degree of interdependency among the parts, and the advantages of operating as a “superorganism”.3

In fact, in what may appear to be an utter contradiction of classical neo-Darwinism, it may often be the case that it is in the interest of a gene, or an individual, to promote the well-being of an interdependent “other”, simply because functional interdependence means just that; it’s “one for all, and all for one” to borrow a legendary slogan. Consider this hypothetical example. If one of the two oarsmen in the rowing game (above) should suffer from thirst and dehydration in the summer heat (he forgot his water bottle), his partner might decide to share his/her water supply, in the interest of reaching their joint goal. Or, to cite a concrete example from nature, consider the exquisitely complex energy-production services that the mitochondria provide for eukaryotic cells, in their own direct self-interest. Or, for that matter, consider the innumerable situations in human societies where our well-being depends, unequivocally, upon the performance of others — airline pilots, railroad engineers, surgeons, and the other motorists that we encounter on the highways, to name a few. How do we explain these cooperative relationships? They have nothing to do with altruism, kin selection, reciprocal altruism or even (strictly speaking) tit-for-tat mutualism. They are sustained by “pure” self-interest.

One final point related to “downward causation,” “government” and the “selfish genome.” At the most basic level of biological organization (the genome itself), there is mounting evidence that the genes do not inhabit a “bean bag” (in Ernst Mayr’s felicitous caricature), and that morphogenesis is not a “mindless” process. Rather, it is an organized, cybernetic process which entails the extensive use of superordinate feedback controls (the very essence of a “teleonomic” system). In other words, selfish genes are only citizens on good behavior in the selfish genome, and the outlaws, tax evaders and parasites among them do not have a license to pursue their anti-social interests ad libitum.

EVOLUTION AS A MULTILEVEL PROCESS

There is one other aspect to Holistic Darwinism that should be mentioned briefly. It relates to the traditional distinctions between parts and wholes, individuals and groups, even “self-interest” and the “public interest” in political theory.  During the past decade or so, there has been a growing appreciation of the fact that evolution is a multi-leveled, hierarchical (some prefer “holarchical”) process, just as survival and reproduction is a multifaceted problem (see especially Koestler 1967;  Corning 1983, 2003a, 2005; Brandon and Burian 1984; Eldredge and Salthe 1984; Salthe 1985; Eldredge 1985, 1995; Buss 1987; Grene 1987; Wilson and Sober 1994; Maynard Smith and Szathmáry 1995; Michod 1996; Sober and Wilson 1998; Gould 2002). In essence, there is a recognition that natural section operates at various levels of biological organization — from genes to ecosystems — often simultaneously. One implication of this more complex view of evolution is that both competition and cooperation may coexist at different levels of organization, or in relation to different aspects of the survival enterprise. There may be a delicately balanced interplay between these supposedly polar relationships. To illustrate:

  • Eusocial insect species can generally occupy a broader spectrum of habitats and are often able to dominate and even exclude potential competitors among solitary and primitively social species, as noted earlier (see Hölldobler and Wilson 1990). Nevertheless, eusocial insect societies are not the harmonious communities that we once supposed. Among other things, there may be intense competition for breeding rights among potential queens and there is evidence of nepotism among the patrilines in polyandrous species.
  • A number of ant species establish pleometrotic colonies; multiple foundresses cooperate in initial nest construction and brood production. In at least one case, the desert seed-harvester ant Messor pergandei, a study by Rissing and Pollack (1991) has shown that pleometrotic colonies are able to prevail in direct ecological competition with single-foundress colonies; multiple-foundress colonies are able to produce a larger brood raiding force more quickly, and this apparently provides a decisive competitive advantage (group selection). However, other studies of these colonies suggest that one tradeoff may be internal competition among co-foundresses and their offspring — all very suggestive of human societies.
  • Members of African lion prides cooperate and compete with one another in a variety of ways: Females typically hunt large prey in groups, share food and may even share in guarding cubs and defending the pride. As Packer and Ruttan (1988) observe, there is evidence of synergy. For instance, a group of females can more effectively defend a kill against scavengers, including other groups. Likewise, a group of males can successfully defend access to a group of females, whereas single males cannot. However, there is also much intra-coalition competition among the males for mating privileges.
  • One of the more dramatic examples of the interplay between competition and cooperation concerns the Northern Elephant Seals (Mirounga angustirostris). Males of this species, which can weigh up to 4500kg., are legendary for their prolonged and bloody battles for dominance and mating privileges when they come ashore to breed in the winter and early spring. However, the males will only fight when estrous females have formed “harems” of 50 or more. And when the fighting is over, the “alpha” males commonly form coalitions with a half-dozen or more “beta” males, who will defend the perimeter of the harem against other marauding males (in return for which the beta males get limited mating privileges for themselves). Elephant seals generally feed at sea alone, and at great depths, but whenever they are ashore they congregate peacefully in tightly packed “rookeries” that facilitate defense and heat-sharing (a critically-important function in these animals). Males collaborate in this way during their summer moulting season; non-breeding males also aggregate into “loser groups” during the breeding season; females huddle closely together to share heat and defend their pups during the breeding season, and “pods” of weanling pups huddle for warmth and mutual self-defense before setting off on their initial feeding expeditions (Le Boeuf 1985; Le Boeuf and Laws 1994).

THE EVOLUTION OF HUMANKIND

Human evolution may provide a singular illustration of the synergistic, functional group selection hypothesis, and of Holistic Darwinism. In effect, the principles that were elucidated above can also be observed in the evolution of the human species, and in our cultural evolution as well. For various reasons, the evolution of humankind has often been portrayed as a process that is sui generis. Of course, this overlooks the fact that all of evolution can be said to be sui generis, given its historical and situation-specific causal dynamics. As Darwin himself put it in The Descent of Man (1874/1871), any evolutionary innovation depends upon many “concurrent favorable developments” that are always “tentative” (p. 150).  Nevertheless, the evolution of humankind is undeniably one of the more remarkable episodes in evolutionary history.

A number of suggestive and thoughtfully argued theories of human evolution have been advanced over the years. These theories were reviewed and critiqued in depth in Corning (1983, 2003), and a “synthetic” explanation was offered there that, in effect, combined Darwin’s “tripartite” selection theory of human evolution — i.e., family (kin) selection, mutualism (including reciprocity) and group selection — with the concept of functional synergism. As Darwin pointed out, and this point is crucial, the three modes of selection need not be opposed to one another; they can be complementary and mutually reinforcing. In addition, the Synergism Hypothesis posits, in essence, that it was the bioeconomic payoffs (the synergies) associated with various forms of social cooperation that produced the ultimate directional trend over a period of several million years, from the earliest bipedal hominids to modern Homo sapiens. That is, the synergies produced by various collaborative behavioral innovations provided “proximate” rewards or reinforcements (as the behaviorists would say) that were substantial enough to create a behavioral “pacemaker” (sensu Ernst Mayr, 1960) for the “progressive” evolution over time of our distinctive wardrobe of biological characteristics. In effect, we invented ourselves in response to various ecological pressures and opportunities — a paradigm that may be more widely applicable to evolutionary change than is generally appreciated (see Corning 1996a; also 2003). Here I can only summarize the argument.4

The traditional approach to explaining human evolution has been to propose a “prime mover” theory, which is typically portrayed as the “engine” that has powered the course of human evolution. Darwin, in The Descent of Man (1874/1871/), singled out the role of tool-making. E.O. Wilson (1975) stressed our primate “preadaptations” and speculated about the possible role of an unspecified “autocatalysis”. Bipedalism, which we now know preceded the development of the “big brain,” is currently viewed by many theorists as the “breakthrough” development (e.g., Johanson, Leakey, White, etc.). Major climate changes during the Miocene and Pliocene have also been suggested as important precipitating factors (e.g., Coppens, Vrba). Then there are the various competing “group theories”: group hunting (Dart, Washburn, Ardrey, Thompson, Stanford and others), group scavenging (Potts, Blumenshine, Shipman, etc.), female gathering (Zihlman and Tanner), the nuclear family and male provisioning (Lovejoy), collective defense against predation (Kruuk, Kortlandt, Alexander, etc.) and the ever-popular group conflict hypothesis, which traces back to Darwin and Spencer and which has been championed in this century by Dart, Keith, Ardrey, Lorenz, Bigelow, Otterbein, and Alexander, among others. In the latter stages of human evolution, climate change, population growth, food surpluses, the adoption of fire, language development, increased intelligence and warfare have also been singled out as prime movers by various theorists at various times.

Holistic Darwinism suggests the contrarian view that all of these factors were important but that none was sufficient — the engine is nothing without the car — and that the answer lies in the unique combination of factors that produced, over time, many compatible and mutually supportive cooperative effects (functional synergies). Indeed, objections can be mounted against every one of the factors cited above, taken individually.  For example, bipedalism is not unique to humans; birds are bipedal, after all, and kangaroo forelimbs have atrophied rather than becoming instruments for the skilled manipulation of tools. In fact, hominid bipedalism existed for some millions of years before the “big brain” emerged. Tool-making is also an insufficient explanation; we now know that many species make and use tools, especially our closest relatives the chimpanzees. And, again, crude stone tools were used by our hominid ancestors for perhaps a million years before the more refined and standardized Acheulean tool-kit appeared (Leakey 1994; Tattersall 1995). Group hunting and scavenging are also inadequate explanations, for the same reasons; our ancestors were hardly unique in this regard. The primeval gathering scenario and the nuclear family scenario are appealing, though difficult to support empirically, yet they are in any case insufficient because they overlook other factors — namely, the often serious threat from potential predators and the premium associated with meat-getting (via scavenging, hunting or both) in the more open and diversified environments in which some of the later hominid developments most likely occurred. Even the conflict hypothesis — which Alexander (1979) asserts is both necessary and sufficient — begs the question: why are there no nation-states composed of chimpanzees, which we now know can be quite warlike? The very absurdity of that idea highlights the fact that there had to be many other factors that “worked together” to propel the process. Indeed, the extensive hominid migrations out of Africa over time suggest that conflict avoidance might well have been a common adaptive strategy.

In a major critique of cultural evolution theories many years ago, anthropologist Elman Service came to this emphatic conclusion: “down with prime movers!” (1971, p. 25) The same can be said, equally emphatically, of the larger process of human evolution. Prime mover arguments invariably take for granted some, or all, of the other requisites for survival and reproduction. Very often they reflect a kind of ecological naiveté; they discount the many life-and-death challenges associated with living (and evolving) in a demanding and changeable environment over a period of several million years. But if no one factor alone can provide a sufficient explanation for the evolution of humankind, then what is sufficient?  The answer is that all of the important human traits were necessary and none were sufficient. In effect, there was a mutually-reinforcing synergy among the key innovations — combined effects that would not otherwise have been possible.

“MAN MAKES HIMSELF”

In sum, the dominant theme of human evolution may have been the expansion of various modes of social cooperation (including cooperative modes of competition), which have been rewarded with commensurate bioeconomic benefits. To reiterate, competition and cooperation are not mutually exclusive explanations for human evolution; both played an important role in shaping our evolution. Nevertheless, the thesis here is that increasingly potent (and selectively advantageous) forms of social cooperation may have given our ancestors their competitive edge.

As Edward O. Wilson (1975; 1985) has noted, a multifaceted group-living ecological strategy is a relatively rare occurrence in nature. We rightly admire the complex social organization of honeybees, naked mole-rats, army ants, killer whales, and a small number of other highly social species, including some of our close primate relatives. The synergies that have made such collective survival strategies rewarding for various social species are increasingly well documented. We are among that select company, and it has been the key to our evolutionary success. A human society can be characterized as a “collective survival enterprise.” We meet our basic survival needs through elaborate networks of social cooperation.5

We do not know, and likely never will know, the full story of our evolution as a species, although we are gradually adding more details to the outline and making better-informed guesses. However, there is reason to believe that behavioral changes in the direction of greater social cooperation for specific functional purposes were the “pacemakers” that precipitated supportive morphological changes. In a very real sense, as anthropologist V. Gordon Childe (1951/1936/) put it in the title of his famous book on the rise of civilization, the human species may have “invented” itself. The real key to human evolution, accordingly, was not any single prime mover but the entire suite of cooperative behavioral, cultural and morphological inventions — a synergy of synergies.

An oft-used (and important) illustration of this dynamic is the adoption by evolving hominids of the controlled use of fire (or, more broadly, various exogenous forms of energy). This is a uniquely human cultural invention and is still a major factor in our ongoing evolution — a point that various energy-oriented theorists have thoroughly documented (e.g., White 1943, 1949; Cottrell 1953; Odum 1971; Adams 1975, 1988). The earliest strong evidence for the use of fire by our hominid ancestors is identified with the Middle Pleistocene, perhaps 200-400,000 years B.P. However, some theorists argue on plausibility grounds, albeit with more fragmentary evidence, for a much earlier date.  (See especially the cautious review by James 1989 and the offsetting commentary by Lewis). The controlled use of fire by hominids (in effect, a cooperative animal-tool symbiosis) had enormous long-term benefits. Over the course of time, fire was most likely used as an effective means of defense against predators; it was a source of warmth that facilitated migration into colder climates; it might well have served as an insect repellant and as a means for obtaining honey from bee hives (as a bee suppressant); it probably became a weapon for driving and capturing prey animals; it was a means for shaping and hardening tools; it could be used for conditioning the environment (as in slash and burn horticulture); and, not least, it enabled our ancestors to add to their diets many foods that would otherwise have been toxic, indigestible, or possibly even infectious if eaten raw (Leopold and Ardrey 1972; Stahl 1984 and commentaries).  (See also the case that is developed for what they call the “cooking hypothesis” in Wrangham, et al., 1999; also Wrangham 2001.)

In any event, fire represented the functional equivalent of a major morphological development. With the acquisition of fire, our ancestors were able to expand their niche over time, which in turn changed the selective forces to which our ancestors were subject. Furthermore, fire most likely became another focal point of social cooperation. Fire-keeping was a “collective good” that required a division of labor — for gathering fire wood, fire tending, fire transport and, eventually, fire-making. In other words, this primordial hominid technology, like most human technologies, was at once a source of bioeconomic benefits and a generator of social cooperation and social organization.

How can this synergistic theory of human evolution be tested? Let us try out a few thought experiments: Take away fire (along with other energy sources in modern societies); we are utterly dependent upon exogenous forms of energy. Or, take away language; or bipedalism; or tools and technology. In short, there is no major adaptive modality that we could do without; they are all necessary parts of an interdependent, synergistic system.

An anonymous reviewer of this paper complained that this thought experiment is “inadequate.” We must specify which of the factors was/is most important and be specific about what happens when it/they are removed. Although thought experiments are quite acceptable in physics, they won’t pass muster in political science, it seems.

In the first place, this objection misses the fundamental point.  In a synergistic system, all of its major elements or parts are equally important. That’s what synergy is all about. It would be analogous to asking which part of an automobile is more important, the wheels, the engine, the steering mechanism, or the operator for that matter.  Remove any one of those “parts” and it is equally likely that the vehicle can no longer be operated successfully.

But more important, there have been innumerable in vivo tests of the Synergism Hypothesis over the past 10,000 years or more. An entire chapter of my 2005 book on Holistic Darwinism is devoted to political “devolution” – the declines and collapses that litter the historical record – and it has become increasingly clear, as a result of recent scholarship, that in many cases devolution can be attributed to the loss of a single necessary “part”.  Climate changes, especially droughts, are among the more dramatic and currently “relevant” causes. Likewise, drastic soil depletion is the most likely culprit in many early Mesopotamian societies, whereas deforestation and the exhaustion of their wood supply crippled an otherwise thriving Easter Island society. (Jared Diamond’s recent book Collapse, 2005, cites many other examples.)

A POST-NEO-DARWINIAN PARADIGM?

I believe that Holistic Darwinism can plausibly be viewed as a candidate for a post-neo-Darwinian theoretical paradigm.  It refocuses evolutionary theory on the “vessels” and their functional properties as the vanguard of evolutionary change. In fact, that is where natural selection as a causal dynamic actually occurs; to use an older turn of phrase, it is the phenotypes  which are “tested” in the environment. Holistic Darwinism shifts our focus from the anthropomorphic purposes of selfish genes in theoretical isolation to the evolved, emergent purposiveness of the living systems as wholes, and to the functional interactions and relationships (adaptations in specific environments) that result in differential survival and reproduction.  It also stresses the “synergistic selection” of various combinations.

Equally important, Holistic Darwinism de-emphasizes (without denying) the role of genetic mutations, recombinations, transpositions, etc., as sources of creativity in evolution and emphasizes purposeful innovations which may occur at the behavioral, cognitive, even social levels (inclusive of symbiogenesis). In this model, proximate “neo-Lamarckian selections” by wholes (i.e., adapting organisms and, in some cases, adapting groups) assume a much more important role in evolutionary change than is acknowledged in neo-Darwinism.

Needless to say, it also entails an explanatory theory relating to the emergence, elaboration and, often enough, the “devolution” of complex systems at all levels.  Like Darwin’s great concept of natural selection, the Synergism Hypothesis (and the concept of “synergistic selection”) represents an open-ended umbrella term that focuses our attention on a common property associated with the great diversity of complex biological systems in nature. It directs our attention to the “why” question, and it posits that synergy itself has been a major causal agency in the evolutionary process.

Finally, and perhaps most contentious of all, I maintain that Holistic Darwinism also encompasses human evolution and, indeed, the ongoing bio-cultural evolution of our species.  It is a seamless theoretical framework that does not require any additional causal principle or “mechanism” to account for humankind.6 We must recognize human culture, human economies, even human political systems for what they truly are — an augmentation of adaptive modalities that can be found in rudimentary form in many other species. (In this paradigm, the activities of humankind virtually everywhere on earth are viewed as an integral part of the evolutionary process, not something that is separate from it.) But, having recognized that, we must also acknowledge our uniqueness as a species. Complex human societies are as different from those of honeybees or naked mole-rats or even chimpanzees as complex multicellular organisms are from single-celled protists. The cumulative, synergistic effects of many differences in degree have produced a difference in kind — and a dynamic of rapid change at the behavioral/cultural level that is obviously unprecedented. This is an evolutionary development that Holistic Darwinism can fully comprehend. It is not even conceivable in a theoretical world that barely recognizes the existence of wholes.

I believe that it is time to refocus the Necker cube on the problem of explaining the evolution of complex systems in a way that is fully consistent with Darwin’s vision. In the long run, I believe that the Darwinian, functional explanation of complexity will prevail over various orthogenetic theories of “self-organization” (in reality a teleological “black box” that begs the “how” question), or theories that postulate a random “drunkard’s walk” (Gould 1996).  Gould’s argument is especially surprising, coming as it does from such a sophisticated and articulate student of evolution.  It is a formulation which tacitly ignores the functional costs and benefits related to the evolution of biological complexity; complexity is not a free lunch but a cumulation of adaptive innovations over the eons.  I can only second the conclusion of George Williams in the peroration of his famous book (1966, p. 273): “It may not, in any absolute or permanent sense, represent the truth, but I am convinced that it is the light and the way.”

SOME IMPLICATIONS FOR POLITICAL THEORY

The implications of this new paradigm for political theory should be obvious.  In a chapter in the new Handbook of Evolution on “ The Evolution of Politics” (Corning 2004), I argue (as others have) that what we term politics in human societies is an evolutionary phenomenon with roots that may trace back several million years.  Here I will summarize just a few of the key points.

It is appropriate, I believe, to divide political theory into two broad schools, the origins of which can be traced back the ancient Greeks.  One school generally adheres to a holistic (some say idealistic) vision that a society, or polity is an integrated whole — a superorganism — that has overarching interests, goals, and purposes and that entail interdependence and cooperation.  The other school (often characterizing itself, somewhat smugly, as realist) is focused on individual motives, interests and goals.  This school stresses the dynamics of competition and the struggle for power among individuals, or factions, as the quintessential characteristic of politics.

The latter view of politics received a boost from the then new science of ethology (the study of animal behavior) in the latter 1960s and 1970s, most notably from anthropologists Lionel Tiger and Robin Fox in their provocative popularization, The Imperial Animal (1971).  What Tiger and Fox did, and with a certain relish, was to equate politics in human societies with dominance competition in the natural world.  Thus politics is a world of winners and losers. The political system, they claimed, is synonymous with a “dominance hierarchy.”

At first glance, it may seem that Tiger and Fox were promoting the Machiavellian vision (actually a distortion of Machiavelli’s thinking) that politics is nothing more than a struggle for power.  Yet Tiger and Fox also recognized that dominance competition in nature has a purpose.  It is related to competition for scarce resources – nest sites, food, and especially obtaining mates. Tiger and Fox concluded that “the political system is the breeding system.”  Having thus flagrantly caricatured this ancient term, Tiger and Fox were then forced to concede that politics in human societies serves very different purposes.  It is more often associated with leadership, the division of labor and cooperative activities of various kinds.  It has become dissociated for the most part from breeding functions.  Unfortunately, Tiger and Fox did not bring this crucial distinction into focus.

A more coherent case for the proposition that human politics is related to dominance behaviors in other species was developed in a succession of works by the primatologist Frans de Waal, beginning with his Chimpanzee Politics: Power and Sex Among Apes (1982).  (See also de Waal 1989; Harcourt and de Waal 1992; and de Waal 1996.)  Drawing on his own extensive research in captive chimpanzees, as well as the many long-term field studies of these animals, de Waal offered a deeper, richer perspective on the issue.  The struggle for power and influence is ubiquitous among these animals, he acknowledged.  From the animals’ motivational perspective, this may well be an end in itself.  And, yes, the dominant animals may gain advantages in terms of such things as nesting sites and breeding privileges.

But there is much more to dominance behaviors than this. The competition for status very often involves coalitions and alliances; it is often a group process rather than an individualistic, Hobbesian “war”.  Indeed, there is much evidence that social constraints on dominance behaviors are common, both in these and other social animals; coalitions sometimes form to thwart the actions of a dominant animal.  And in bonobos (or pygmy chimpanzees), a loose female hierarchy seems to form the organizational backbone of the group; females often band together to constrain an aggressive male (de Waal 1997).

More important, stable dominance hierarchies in chimpanzees and other social animals also have functional importance for the group B maintaining peace, arbitrating disputes, limiting destructive competition, mobilizing collective action, even defending the group against outside threats.  The intense interdependence of social animals like chimpanzees and bonobos also leads to a degree of reciprocity and generosity, such as food sharing.  More recent work in chimpanzees, bonobos, orangutans and other socially-organized species also suggests that interpersonal social relationships and interactions can be very complex, and that cultural influences may also play an important part (see especially de Waal 1989, 1996, 1999, 2001, 2005, 2006).  In fact, there may even be a degree of democratic participation in various group decision-making processes (Conradt and Roper 2003).  Nor does one size fit all.  The dynamics may differ from one group to the next, or even within the same group over time.  (In addition to the de Waal references above, see especially Kummer 1968, 1971; E.O. Wilson 1975; Lopez 1978; Strum 1987; Dunbar 1988; Wrangham et al.,1994; Boesch and Tomasello 1998; Whiten et al., 1999; van Schaik et al., 2003.)   Also relevant is the anthropological evidence for what Boehm (1996, 1997, 1999) calls an “egalitarian syndrome” – a democratic political culture that is common to most small-scale human societies, like hunter-gatherers (see also Knauft 1991).

In sum, a more sophisticated ethological cum anthropological model implies that both the holistic (idealist) model of politics and the egoistic (realist) model have some validity; they are not mutually exclusive.  As de Waal (1996, pp. 9, 102) points out, we also need to ask “what’s in it for the subordinate?”  His answer: “The advantages of group life can be manifold…. increased chances to find food, defense against predators, and strength in numbers against competitors….Each member contributes to and benefits from the group, although not necessarily equally or at the same time…Each society is more than the sum of its parts.”  In other words, cooperative social groups produce mutually-beneficial synergies.

Accordingly, in the modern version of the ethological model, dominance behaviors may take on the functional attributes of leadership, and a dominance hierarchy may provide a framework for organizing various cooperative activities, including a division (combination) of labor (see Corning 1983; cf., Masters 1989; Grady and McGuire 1999; Rubin 2002).  Such organized political systems are characterized by overarching collective goals, decision-making, interpersonal communications, social control processes and “feedback”.  In short, political systems are cybernetic systems.  And the accumulating evidence supports the contention that cybernetic social processes – political processes – have also been an integral and necessary element in the evolution of human societies.  Politics is not simply an artifact of competing self-interests but a vitally-important functional element of the “collective survival enterprise.”

Though politics as we have defined it here often entails the pursuit of narrow self-interests (in accordance with the realist model), it also takes place within a larger context – the purposes and interests of the collective survival enterprise as an interdependent system (in accordance with the idealist model).  Both of these classical renderings of politics have merit; they are not, in fact, mutually exclusive.  Indeed, there is an inherent interplay, and very often a tension, between them.

Nevertheless, the reality of the human condition is that the superorganism is the key to our survival and reproduction, as it most likely has been for millions of years.  However, this vision of the public interest does not negate or ignore our individual self-interests.  Rather, it represents an aggregation of those interests into an immensely complex system of synergies based primarily on mutualism and reciprocity.  The superorganism serves our self-interests in a multiplicity of ways; it provides both public goods and what I call “corporate goods” (an often overlooked and a vastly underrated factor).7 And the public interest consists of preserving and enhancing these collective benefits.

In this light, it is clear that the state evolved as an instrumentality for self-government and the pursuit of the public interest – though its overarching purpose is all too often subverted.  Plato and Aristotle apprehended the overarching purpose of the collective survival enterprise (and its inherent vulnerability) in their conception of the polis, and Aristotle prescribed a mixed government under law as our best hope for ensuring that the public interest would be faithfully served.  Plato and Aristotle also recognized that a fair-minded form of justice is an essential element of the public interest; this is the only way to ensure the long-term stability and legitimacy (the willing consent) of the members of the community. (See Corning 2003b, 2005 and the plethora of recent studies on the psychology of “fairness” cited therein.).  Over the past 2000 years we have added very little to this vision that is fundamentally new, though we have made many important improvements in the machinery of self-government.

The bottom-line conclusion of Plato and Aristotle remains valid today, I believe.  For better or worse, our evolutionary future is dependent upon the goods and services that are provided (or not) by the collective survival enterprise, along with the decisions and actions that we undertake collectively (or not) in the public interest.  For this reason, the continuing quest for social justice, and the good life, remains the central challenge for every organized society, as well as for each one of us.  It is a goal worth striving for, because our own survival, and certainly that of our descendants, may very well depend upon it.  Nothing less than our evolutionary future is at stake. To paraphrase the American founding father, Benjamin Franklin, in the long run either we will survive together or go extinct separately.

SOME IMPLICATIONS FOR ECONOMIC THEORY

In his landmark book, The Great Transformation (2001/1944), Karl Polanyi produced a classic critique of the liberal (conservative) ideal of free market capitalism that still resonates today. According to liberal economic doctrine, if various impediments are removed so that a market economy can operate free of constraints and imperfections, it would not only be self-regulating and self-equilibrating but would lead to the most “efficient” utilization of capital, resources and labor. Indeed, for the most fervent of free market advocates, this ideal seems to be an end in itself.

Polanyi’s argument was that this model is utopian.  It can never work in the real world, and the history of the great 19th century “transformation” to industrial market capitalism (not to mention the history of the 20th century) proves it.  As Polanyi put it, markets are “embedded” in human societies, and the needs and wants of a society and its members cannot in the long run be subordinated to market efficiency.  This is not simply a normative statement, or a moral claim, moreover. It is an empirical reality.  Ultimately, people will rebel and governments will either need to intervene or else they will be replaced. Indeed, even capitalists do not, as a rule, want complete freedom from government interference. They want laws, and regulations, and property rights and, very often, government support in the form of protections and subsidies.

Economist Joseph Stiglitz, in a new foreword to a reprint of Polanyi’s book, notes that most economists these days concur with Polanyi’s diagnosis.  In addition, Stiglitz points out, advanced industrial economies like ours are even more constrained by the fact that the vast majority of the population depends on the marketplace for their livelihoods. So “labor” is not simply an “input” that can lie idle or find other uses if no jobs are available. Unemployment has inescapably destructive consequences.

What Holistic Darwinism adds to this critique is a theoretical foundation. To repeat, from an evolutionary/biological perspective, our basic vocation as individuals and families is survival and reproduction – and specifically the meeting of some 14 domains of “basic needs” (according to the Survival Indicators Project). These are biological imperatives. To repeat, we are all implicitly engaged in a “survival enterprise,” and, in a modern market economy, our individual needs have been aggregated into an extraordinarily complex, interdependent system – a collective survival enterprise. We are, in effect, parties to a “biological contract.”  Accordingly, the division/combination of labor and economic markets represent, at bottom, a strategy for meeting our basic biological needs (and often much more, of course). And if the markets fail to do so, for whatever reason, corrective actions are to be expected.  Markets exist to serve our needs, not the other way around.

Neo-Darwinians and social Darwinists might object that, to the contrary, capitalism is natural because it embodies our innately selfish and competitive natures – as Adam Smith himself suggested.  The problem with this model is that it overlooks our fundamental dependence on cooperation and the fact that, as noted earlier, we evolved over several million years in closely cooperating social groups. Equally important, modern societies remain deeply dependent on cooperation; we live in highly interdependent economic systems.  Competition may also be natural and inevitable, reflecting a duality in human nature, but it must also be subordinated to our collective needs.  And if capitalist markets fail to meet our needs, we have every right to cooperate in an effort to redress our grievances – whether it be through labor unions, or governments, or political movements or even revolution if necessary. As the American Declaration of Independence puts it, governments are “instituted among men” to secure our “inalienable rights,” and derive their “just powers” from “the consent of the governed.”  Furthermore, whenever any form of government “becomes destructive of those ends,” the people have the right “to alter or abolish it.”

SOME IMPLICATIONS FOR POLITICAL SCIENCE

Finally, Holistic Darwinism has some significant implications for political science as a social science. Very briefly, these include the following, among others:

First, Holistic Darwinism directs us to refocus the Necker cube on the fundamental biological problem of survival and reproduction, what Darwin called the “struggle for existence,” and, specifically, our suite of some 14 ongoing “basic needs” domains, as documented by the Survival Indicators project (see Corning 2005). It redefines the basic problem for any organized human society and its members. It also necessitates the use of a more expansive multi-level, multi-factorial (and multi-disciplinary) causal framework (see Figure III). And it requires a focus on relationships, and “systems”, including the dynamics of “synergistic selection” as a major causal agency in our ongoing evolution.

To repeat, the revival of the “superorganism” metaphor in evolutionary biology also lends support to the often-denigrated concept of a “public interest.” The public interest is nothing less than our shared interest in survival and reproduction, and in the well-being of the interdependent survival enterprise upon which we all depend. Both public goods and “corporate goods” (where many participants may benefit, though unequally) are major contributors. Moreover, these “goods” are ultimately defined in terms of the requisites (basic needs) for survival and reproduction. Political science should be unapologetic about addressing the relationship between politics and the collective survival enterprise.

Holistic Darwinism and the cybernetic theory of government also sheds light on the ongoing debate about democracy versus authoritarian rule.  To reiterate, cybernetic (decision-making, communications and control) processes are in fact essential elements of any complex, goal-oriented system.  Whether or not a given human political system/government is democratic or authoritarian in nature is therefore distinct from a consideration of its functional performance.  Either kind of system may be effective, or not, in relation to the ongoing survival enterprise. Using our basic needs as the criteria (the implicit goal), political science could utilize the cybernetic model in a much more concerted way. (For an elaboration on the cybernetic model, see Corning 2005.)

The collective survival enterprise in any human society is always contingent and sometimes precarious (contrary to the many “orthogenetic” evolutionary theories, which remain popular).8 What the great 20th century evolutionary biologist, Theodosius Dobzhansky, said of the overall evolutionary process applies also to human societies: “The future is not vouchsafed by any law of nature, but it may be striven for.” This calls for us to use an analytical approach that can encompass and account for both the rise and the decline of political systems, as well as major systemic changes over time.

Holistic Darwinism also points the way toward a healing of the deep theoretical (and ideological) cleavage between individualistic/competitive views of society and communal/cooperative models. There is in fact a duality that has deep roots in our evolutionary heritage. Political scientists should focus more closely on the dynamics of this duality.

Much of what the social sciences have in recent generations treated as sui generis and unique to human kind and human societies in fact has many analogues and homologues in the natural world: our competitiveness and our intense social cooperation, our selfishness and our altruism, our division/combination of labor, our moral propensities, our collective violence (see Corning 2007) and our political systems. Moreover, our politics are not a recent cultural overlay (or “veneer” in Frans de Waal’s term) but a behavioral pattern that, most likely, has a long evolutionary history. Political science can learn much about the nature and nurture of our species-typical behaviors by adopting a comparative (human ethology) perspective, as many scholars in the biopolitics movement have done over the years.

((Figure III ))

Although the field of biopolitics has been a very loosely defined and multi-faceted endeavor over the past 30-odd years, its practitioners have been united in accepting the “ground-zero” premise of the biological sciences that biological survival and reproduction is the fundamental problem of the human species, as with all other species. Holistic Darwinism fleshes out this assumption in a way that provides a much more robust, realistic, and balanced theoretical foundation than was the case with the selfish gene-centered paradigm of neo-Darwinism. Accordingly, the bottom line implication of Holistic Darwinism is that political scientists should, first and foremost, view human societies, including their economic and political systems, from an evolutionary/biological perspective. And if this implies a revolution in the discipline, so be it.  It’s long overdue.

 

ENDNOTES

  1. Actually, it was W.D. Hamilton who started it. Hamilton had previously asserted that there were only three forms of social interaction — (1) altruism, (2) exploitative (“zero-sum”) selfishness, and (3) spite (E.O. Wilson 1975).
  1. Douglas H. Boucher (1985), in an edited volume on mutualism, pointed out that there is a long-standing debate among ecologists over the relative importance of competition and co-operation in nature, which can be traced back at least to the 1920s. He noted the remarkable fact that, despite a general bias over the years in favor of competition as the basic organizing principle of nature and a concomitant preference among theoretical ecologists for using the famed Lotka-Volterra competition model in their analyses, in fact a co-operative version of the model (involving a simple sign change) has been reinvented (evidently independently) at least 29 times since 1935. Boucher’s volume reflected yet another of the periodic renewals of interest in the co-operative aspect of ecology. Similarly, in an overview and analysis of co-operative behaviors, Jerram Brown (1983, p. 29) noted: “Natural selection is an ecological process and cannot be understood solely from genetic considerations. Relatedness to non-descendants does not determine the direction or product of natural selection; it only supplies an additional cost or benefit.” Also, Jon Seger (1991), echoing Darwin’s proposed explanation for human evolution in The Descent of Man, points out that the various hypothesized explanations for social life are not mutually exclusive and in many cases might reinforce one another.
  1. Another implication of this insight is that political theory can now be connected directly to evolutionary theory; insofar as politics and government are related to the problem of survival and reproduction, in any species, there is a common functional basis. Human political systems may thus be viewed as variations on an evolutionary theme. For an extended treatment of this paradigm, see Corning 1983, 1996b. Also, see the offerings in Somit 1976; Willhoite 1976; Wiegele 1979; Somit, et al., 1980; de Waal 1982; Schubert 1989; Masters 1989; Schubert and Masters 1991; Vanhanen 1992; Johnson 1995.
  1. Among the many recent publications that are relevant to this complex issue — including items in virtually every issue of the American Anthropologist and Current Anthropology — see especially Johnson and Earle 1987; Byrne and Whiten 1988; Mellars 1989; Mithen 1990; Durham 1991; Jones, et al., 1992; Maryanski and Turner 1992; Smith and Winterhalder 1992; Gibson and Ingold 1993; Quiatt and Reynolds 1993; Soltis, et al., 1995; Holloway 1996; Boehm 1996; Feldman and Laland 1996; Flinn 1997. Particular note should be made of anthropologist William Durham’s (1991) dualistic gene-culture “coevolution” paradigm, which directs our attention to the partially independent nature of human cultural evolution, even as it recognizes its interdependency with biological constraints and influences, and the interactions between the two evolutionary modes. See also Cavalli-Sforza and Feldman 1981; Corning 1983; Boyd and Richerson 1985. Ghiselin (1993), on the other hand, sharply criticizes Durham for using “information” as the basis for his definition of culture, an obvious analogy with the informational properties of genes. Ghiselin rightly observes that this confuses a means, or an instrumentality of transmission and replication, with the functional product: “the recipe is not the cake.” It should also be noted that Flinn, an evolutionary psychologist, stresses the influence of evolved, biologically-based psychological mechanisms in cultural selection and evolution, but this is not incompatible with the notion that there are multiple, interacting levels of causation involved in cultural processes.
  1. In support of this scenario, anthropologist Christopher Boehm (1996) has proposed that a suite of behavioral/cultural “factors” which are widely observed in contemporary egalitarian foraging bands might also have enhanced the influence of group selection among prehistoric human groups (whose life-styles are presumed to have been similar). These factors are (1) internal social “leveling” pressures, (2) moralistic policing of cheats and shirkers, (3) consensual decision-making and shared within-group adaptive strategies, and (4) marked differences between groups in adaptive strategies. Together, these behaviors could have had the effect of dampening within-group variation and selection pressures while augmenting between-group variation and selection. Once again, this implies a behavioral “pacemaker” for natural selection. Also consistent is the novel proposal of Wilson and Dugatkin (1997) that “assortative interactions” among individual organisms at the cognitive/ behavioral level may also play a role in determining differential reproductive success. (It could serve as an alternative mechanism of group selection.) As Wilson and Dugatkin note, this scenario is particularly relevant to human evolution. On the other hand, Soltis et al., (1995), have questioned what they call “cultural group selection” (the differential survival of different cultural traits) as a significant factor in human evolution. Utilizing data from New Guinea, they conclude that it would take 500-1000 years for a cultural trait to spread by a process of differential group “extinction”. However, as various commentators on their paper pointed out, much depends upon the assumptions used, how groups are defined, which data set is used, and, indeed, which trait is involved. A food taboo is one thing, but a weapon — say thrown spears or horse cavalry, or wheeled chariots, or Greek Fire, or the phalanx, or the cross-bow, or siege cannons, or tanks — can be used by one group to gain a military advantage that results in a rapid process of differential group selection. History is replete with examples of such military “breakthroughs”. Conversely, many cultural traits “diffuse” between groups without discernable (biological) selective consequences.
  1. This is not to deny either the partial autonomy of cultural processes/systems or our unique (evolved) biological needs and psychological capabilities — “human nature”. The “mechanism” of cultural evolution, which I refer to elsewhere as “teleonomic selection,” involves an “interaction” between these two sets of causal factors (and much more besides). Teleonomic selection obviously plays a unique role in shaping the course of humankind. But the rudiments of this “mechanism” can be found elsewhere in nature as well. (For a full discussion of this issue, see Corning 1983, 2003; also see Barkow, et al., 1992; Flinn 1997.)
  1. It is curious that both in economic theory and in classical neo-Darwinian evolutionary theory (both of which utilize game theory models), there has been little attention and even less theoretical work done on what I refer to as “corporate goods.” In the corporate goods model (which can include any number of players), the participants may contribute in many different ways to a joint product — say the capture of a large game animal or the manufacture and sale of an automobile.  (I like to call it a “combination of labor.”) However, unlike “collective goods,” or “public goods” that are indivisible and must be equally shared (even possibly with non-participants and cheaters), corporate goods can be divided in accordance with various principles, or “rules” or “contracts”.  The division of the spoils is thus not preordained, as is the case with the payoffs in most game theory models. In other words, the payoff matrix can be manipulated at will.  Indeed, the question of how the goods are divided up may be crucially important in determining if the “game” will be played at all.  If this sounds familiar, even commonplace, it is because corporate goods “games” are, in fact, ubiquitous in human societies, and are fairly common in nature as well.  It is the predominant form of economic organization in a complex human society.
  1. Orthogenetic (law-like, deterministic) theories of evolution, human evolution, and even political evolution have a lineage that can be traced back to Aristotle and include Lamarck, Herbert Spencer, Ilya Prigogine, Tielhard de Chardin and a host of others. In the political sphere, a recent notable case in point was Robert Wright’s popularization, Nonzero: The Logic of Human Destiny (2000), as well as the more scholarly theoretical work of George Modelski (1987) and Devezas and Modelski (2003). Holistic Darwinism and the Synergism Hypothesis contradict such orthogenetic theories. They are directly at odds with the fundamental Darwinian assumption of a contingent historical process that is “chaotic” in the sense that it is fully determined but inherently not predictable. For much more on this issue, see Corning 2003,2005.

 

REFERENCES

Adams, R. (1975). Energy and Structure. Austin: University of Texas Press.

Adams, R. (1988). The Eighth Day: Social Evolution as the Self-Organization of Energy. Austin: University of Texas Press.

Alexander, R.D. (1979). Darwinism and Human Affairs. Seattle and London:University of Washington Press.

Ardrey, R. (1966). The Territorial Imperative. New York:Atheneum.

Atsatt, P.R. (1988). “Are Vascular Plants Inside-Out Lichens?” Ecology, 69, 17-23.

Atsatt, P.R. (1991). “Fungi and the Origin of Land Plants.” In Symbiosis as a Source of Evolutionary Innovation, eds. L. Margulis and R. Fester, pp. 301-305. Cambridge, MA:The MIT Press.

Avital, E., and E. Jablonka.(1994) Social Learning and the Evolution of Behaviour. Animal Behaviour, 48:1195-1199.

Avital, E. and E. Jablonka (2000) Animal Traditions: Behavioural Inheritance in Evolution. Cambridge, U.K.: Cambridge University Press.

Axelrod, R. (1984)  The Evolution of Cooperation.  New York:  Basic Books.

Axelrod, R. (2001)  The Complexity of Cooperation: Agent-Based Models of Competition and Collaboration.  Princeton, NJ: Princeton University Press.

Axelrod, R., and W. Hamilton. (1981). “The Evolution of Cooperation.” Science, 211, 1390.

Barkow, J., et al. (1992). The Adapted Mind. Oxford:Oxford University Press.

Bateson, P.P.G. (1988) “The Active Role of Behavior in Evolution.” In Evolutionary Processes and Metaphors, eds., M.W. Ho and S.W. Fox, pp. 191-207.   New York: Wiley.

Bell, G. (1985). “Origin and Early Evolution of Germ Cells as Illustrated by the Volvocales.” In Origin and Evolution of Sex, eds. H.O. Halvorson and A. Monroy, pp. 221-256. New York:Alan R. Liss, Inc.

Bermudes, D., and L. Margulis. (1987). “Symbiont Acquisition as Neoseme: Origin of Species and Higher Taxa.” Symbiosis, 4, 185-98.

Bigelow, R. (1969). The Dawn Warriors: Man’s Evolution Towards Peace. Boston:Little  Brown.

Binmore, K. (1994) Game Theory and the Social Contract: Volume I: Playing Fair.    Cambridge, MA: The MIT Press.

Binmore, K. (1998) Game Theory and the Social Contract: Volume II: Just Playing.    Cambridge

Binmore, K. (2004) Reciprocity and the Social Contract. Politics, Philosophy & Economics, 3(1):5-35.

Bloom, H. (1997). “A History of the Global Brain: Creative Nets in the Pre-Cambrian Age.” ASCAP (Across Species Comparisons and Psychopathology Society) 10(3), 7-11.

Boehm, C. (1996). “Emergency Decisions, Cultural-Selection Mechanics, and Group Selection.” Current Anthropology, 37, 763-793.

Boehm, C. (1997) “Impact of the Human Egalitarian Syndrome on Darwinian Selection Mechanics.”  The American Naturalist, 150:5100-121.

Boehm, C. (1999) Hierarchy in the Forest: The Evolution of Egalitarian Behavior.

Cambridge, MA: Harvard University Press.

Boesch, C. and Tomasello, M. (1998) “Chimpanzee and Human Cultures.”  Current Anthropology, 39(5): 591-614.

Boucher, D.H.,ed. (1985). The Biology of Mutualism: Ecology and Evolution. New York:Oxford University Press.

Bowles, S. and H. Gintis. (2002) “Homo Reciprocans.”  Nature, 415: 125-128.

Bowles, S., et al. (2003) The Co-evolution of Individual Behaviors and Social Institutions Journal of Theoretical Biology, 223:135-147.

Boyd, R., and P.J. Richerson. (1985). Culture and the Evolutionary Process. Chicago:University of Chicago Press.

Boyd, R., and P.J. Richerson. (1992). “Punishment Allows the Evolution of Cooperation (or Anything Else) in Sizable Groups.” Ethology and Sociobiology, 13, 171-195.

Boyd, R., et al. (2003) The Evolution of Altruistic Punishment. Proceedings of the National Academy of Sciences,  100(6):3531-3535.

Brandon, R.N., and R.M. Burian, eds. (1984). Genes, Organisms, Populations: Controversies Over the Units of Selec­tion. Cambridge, MA:The MIT Press.

Brembs, B. (1996). “Chaos, Cheating and Cooperation: Potential Solutions to the Prisoner’s Dilemma.” Oikos, 76, 14-24.

Brown, J.L. (1983). “Cooperation–A Biologists’ Dilemma.” Advances in the Study of Behavior, 13, 1-37.

Buss, L.W. (1987). The Evolution of Individuality. Princeton, NJ:Princeton University Press.

Byrne, R.W. and A. Whiten, eds. (1988). Machiavellian Intelligence: Social Expertise and the Evolution of Intellect in Monkeys, Apes and Humans. Oxford: Clarendon Press.

Calderone, N.W., and R.E. Page. (1992). “Effects of Interactions Among Genotypically Diverse Nestmates on Task Specialization by Foraging Honey Bees (Apis mellifera).” Behavioral Ecology and Sociobiology, 30, 219-226.

Campbell, D.T. (1974). “Downward Causation in Hierarchically Organized Biological Systems.” In Studies in the Philosophy of Biology, eds. T. Dobzhansky and F.J. Ayala, pp. 85-90. Macmillan.

Cavalli-Sforza, L.L., and M.W. Feldman. (1981). Cultural Transmission And Evolution. Princeton, NJ:Princeton University Press.

Childe, V.G. (1951[1936]). Man Makes Himself. New York:New American Library.

Clutton-Brock, T.H., and G.A. Parker. (1995). “Punishment in Animal Societies.” Nature, 373, 209-216.

Conradt, L., and T.J. Roper (2003) Group Decision-Making in Animals. Nature. 421: 155-158.

Corning, P.A. (1982). “Durkheim and Spencer.” The British Journal of Sociology, 33, 359-382.

Corning, P.A. (1983). The Synergism Hypothesis: A Theory of Progressive Evolution. New York:McGraw-Hill.

Corning, P.A. (1995). “Synergy and Self-Organization in the Evolution of Complex Systems.” Systems Research, 12, 89-121.

Corning, P.A. (1996a). “The Co-operative Gene: On the Role of Synergy in Evolution.” Evolutionary Theory, 11, 183-207.

Corning, P.A. (1996b). “Synergy, Cybernetics and the Evolution of Politics.” International Political Science Review, 17, 91-119.

Corning, P. A. (2003a) Nature’s Magic: Synergy in Evolution and the Fate of Humankind  New York/Cambridge: Cambridge University Press.

Corning, P.A. (2003b) Fair Shares: Beyond Capitalism and Socialism, Or the Biological Basis of Social Justice. Politics and the Life Sciences, 22, 12-32.

Corning, P.A. (2004) The Evolution of Politics In Handbook of Evolution, Eds., F.M. Wuketits and C. Antweiler, pp 191-252. Weinheim:Wiley-VCH Verlag GmbH & Co.

Corning, P.A.  (2005)  Holistic Darwinism:  Synergy, Cybernetics and the Bioeconomics of Evolution. Chicago: University of Chicago Press.

Corning, P.A. (2007) “Synergy Goes to War: A Bioeconomic Theory of Collective Violence.” Journal of Bioeconomics, 9, 109-144.

Cottrell, F. (1953). Energy and Society. New York:McGraw Hill.

Darwin, C.R. (1874/1871). The Descent of Man, and Selection in Relation to Sex. New York:A.L. Burt.

Dawkins, R. (1987/1986). The Blind Watchmaker. New York:W.W. Norton.

Dawkins, R. (1989/1976). The Selfish Gene, 2nd ed. Oxford:Oxford University Press.

Deacon, T.W.(1997) The Symbolic Species:  The Co-Evolution of Language and the Brain. New York: W.W. Norton.

Devezas, T., and G. Modelski (2003)  “Power Law Behavior and World System Evolution: A Millenial Learning Process.”  Technological Forecasting and Social Change,70: 819-859.

de Waal, F. (1982) Chimpanzee Politics: Power and Sex Among Apes.    New York: Harper and Row.

de Waal, F.B.M. (1997) Bonobo:  The Forgotten Ape.    Berkeley: University of California Press

de Waal, F.B.M. (1996) Good Natured: The Origins of Right and Wrong in Humans and Other Animals. Cambridge, MA:Harvard University Press.

de Waal, F.B.M. (1999) Cultural Primatology comes of Age. Nature, 399:635-636.

de Waal, F.B.M., Ed. (2001) Tree of Origin: What Primate Behavior Can Tell Us about Human Social Evolution.    Cambridge, MA: Cambridge University Press.

Diamond, J. (2000) Collapse: How Societies Choose to Fail or Succeed. New York: Viking Press.

Dugatkin, L.A., and M. Mesterton-Gibbons. (1996). “Cooperation Among Unrelated Individuals: Reciprocal Altruism, By-product Mutualism and Group Selection in Fishes.”  BioSystems 37, 19-30.

Dugatkin, L.A., and H.K. Reeve. (1994). “Behavioral Ecology and Levels of Selection: Dissolving the Group Selection Controversy.” Advances in the Study of Behavior, 23, 101-133.

Dugatkin, L.A., et al. (1992). “Beyond the Prisoner’s Dilemma: Towards Models to Discriminate Among Mechanisms of Cooperation in Nature.”  Trends in Ecology and Evolution, 7, 202-205.

Dunbar, R. I. M. (1988) Primate Social Systems.    London: Croom Helm.

Durham, W.H. (1991). Coevolution: Genes, Culture and Human Diversity. Stanford, CA:Stanford University Press.

Eldredge, N. (1985). Unfinished Synthesis: Biological Hierarchies and Modern Evolutionary Thought. Oxford University Press.

Eldredge, N. (1995). Reinventing Darwin. New York:John Wiley & Sons.

Eldredge, N., and S.N. Salthe. (1984). “Hierarchy and Evolution.” In Oxford Surveys in Evolutionary Biology, Volume 1, eds. R. Dawkins and M. Ridley, pp. 184-208. Oxford:Oxford University Press.

Emlen, S.T. (1996). “Living With Relatives: Lessons from Avian Family Systems.” IBIS, 138, 87-100.

Falk, D., et al. (2001) Driving Forces of Informal Sanctions Working Paper Series ISSN 1424-0459 University of Zurich

Fehr, E. and  S. Gächter (2000a) Cooperation and Punishment in Public Goods Experiment. American Economic Review, 90: 980‑994.

Fehr, E. and  S. Gächter (2000b)  Fairness and Retaliation: The Economics of Reciprocity. Journal of Economic Perspectives, 14: 159‑181.

Fehr, E. and S. Gächter. (2002) “Altruistic Punishment in Humans.”  Nature, 415:137-140.

Feldman, M.W., and K.N. Laland. (1996). “Gene-Culture Coevolutionary Theory.” Trends in Ecology and Evolution, 11,  453-457.

Fewell, J.H., and M.L. Winston. (1992). “Colony State and Regulation of Pollen Foraging in the Honey Bee, Apis Mellifera L.” Behavioral Ecology and Sociobiology, 30, 387-393.

Flinn, M.V. (1997). “Culture and the Evolution of Social Learning.” Evolution and Human Behavior, 18, 23-67.

Frank, S.A. (1995). “Mutual Policing and Repression of Competition in the Evolution of Cooperative Groups.” Nature (London), 377, 520-522.

Frank, S.A. (1996). “Policing and Group Cohesion When Resources Vary.” Animal Behaviour, 52, 1163-1169.

Frank, S.A. (2003) Perspective: Repression of Competition and the Evolution of Cooperation. Evolution, 57(4):693-705.

Franks, N.R. (1989). “Army Ants: A Collective Intelligence.” American Scientist, 77, 139-45.

Franks, N.R., et al. (1991). “The Blind Leading the Blind in Army Ant Raid Patterns: Testing a Model of Self-organization (Hymenoptera: Formicidae).Journal of Insect Behavior, 4(5), 583-607.

Ghiselin, M.T. (1993). “Book Review: Coevolution: Genes, Culture, Human Diversity.” Politics and the Life Sciences, 12,     123-124.

Gibson, R., and T. Ingold, eds. (1993). Tools, Language, and Cognition in Human Evolution. Cambridge, MA:Cambridge University Press.

Gintis, H. (2000a) Game Theory Evoloving: A Problem-Centered Introduction to Modeling Strategic Behavior. Princeton, NJ: Princeton University Press

Gintis, H. (2000b) “Strong Reciprocity and Human Sociality.” Journal of Theoretical Biology, 206: 169-179.

Gintis, H. et al. (2003) Explaining Altruistic Behavior in Humans. Evolution and Human Behavior,  24: 153‑172.

Goodnight, C., and L. Stevens. (1997). “Experimental Studies of Group Selection: What They Tell Us About Group Selection in Nature.”  Bulletin of the Ecological Society of America, 77 (3 Suppl. Part 2), 168.

Gould, S.J. (1996). Full House: The Spread of Excellence from Plato to Darwin. New York, NY:Harmony Books.

Gould, S.J. (2002) The Structure of Evolutionary Theory. Cambridge, MA: Belknap (Harvard University Press).

Grady, M.F. and M.T. McGuire (1999) The Nature of Constitutions. Journal of Bioeconomics, 1:227-240.

Grene, M. (1987). “Hierarchies in Biology.” American Scientist, 75, 504-610.

Guzmán-Novoa, E., et al. (1994). “Behavioral and Life-History Components of Division of Labor in Honey Bees, (Apis mellifera L.)” Behavioral Ecology and Sociobiology, 34, 409-417.

Gygax, L. (2002) Evolution of Group Size in the Dolphins and Porpoises: Interspecific Consistency of Intraspecific Patterns. Behavioral Ecology, 13(5):583-590.

Hamilton, W.D. (1964a). “The Genetical Evolution of Social Behavior, I.” Journal of Theoretical Biology, 7, 1-16.

Hamilton, W.D. (1964b). “The Genetical Evolution of Social Behavior, II.” Journal of Theoretical Biology, 7, 17-52.

Hammerstein, P., ed. (2003) Genetic and Cultural Evolution of Cooperation. Cambridge, MA: MIT Press.

Harcourt, A.H., and F.B.M. de Waal, eds. (1992) Coalitions and Alliances in Humans and Other Animals.    Oxford: Oxford University Press.

Hasegawa, E. (1993). “Nest Defense and Early Production of the Major Workers in the Dimorphic Ant Colobopsis nipponicus (Wheeler) (Hymenoptera: Formicidae).” Behavioral Ecology and Sociobiology, 33, 73-77.

Henrich, J. and R. Boyd (2001) AWhy People Punish Defectors.@ Journal of Theoretical Biology,  208:79-89.

Hölldobler, B., and E.O. Wilson. (1990). The Ants. Cambridge, MA:Harvard University Press.

Holloway, R.L. (1995). “Toward a Synthetic Theory of Brain Evolution.” In Origins of the Human Brain, eds. J.P. Changeux and J. Chavaillon.  Oxford:Clarendon Press.

Holloway, R.L. (1996). “Evolution of the Human Brain.” In Handbook of Human Symbolic Evolution, eds. A. Lock and C.R. Peters.  Oxford:Oxford Science Publications.

Hoogland, J.L., and P.W. Sherman. (1976). “Advantages and Disadvantages of Bank Swallow (Riparia riparia) Coloniality.” Ecological Monographs, 46, 33-58.

Hull, D.L. (1980). “Individuality and Selection.” Annual Review of Ecology and Systematics, 11, 311-332.

James, S.R. (1989). “Hominid Use of Fire in the Lower and Middle Pleistocene.” Current Anthropology, 30, 1-26.

Jeon, K.W. (1972). “Development of Cellular Dependence in Infective Organisms: Micurgical Studies in Amoebas.” Science, 176, 1122-23.

Jeon, K.W. (1983). “Integration of Bacterial Endosymbionts in Amoebae.” International Review of Cytology, Supplement 14, 29-47.

Jeon, K.W. (1992). “Macromolecules Involved in the Amoeba-Bacteria Symbiosis.” Journal of Protozoology, 39, 199-204.

Johnson, A.W., and T. Earle. (1987). The Evolution of Human Societies: From Foraging Group to Agrarian State. Stanford, CA:Stanford University Press.

Johnson, G.R. (1995). “The Evolutionary Origins of Government and Politics.” In Human Nature and Politics, eds. A. Smot and J. Losco, pp. 243-305.  Greenwich, CT:JAI Press.

Jones, S., et al. (1992). The Cambridge Encyclopedia of Human Evolution. Cambridge, MA:Cambridge University Press.

Keith, A. (1947). A New Theory of Human Evolution. Gloucester, MA:Peter Smith.

Kendrick, B. (1991). “Fungal Symbioses and Evolutionary Innovations.” In Symbiosis as a Source of Evolutionary Innovation, eds. L. Margulis and R. Fester, pp. 249-261. Cambridge, MA:MIT Press.

Knauft, B.M. (1991) “Violence and Sociality in Human Evolution.” Current Anthropology, 32(4): 391-409.

Koestler, A. (1967). The Ghost in the Machine. New York:Macmillan.

Kummer, H. (1968) Social Organization of Hamadryas Baboons: A Field Study.    Chicago: University of Chicago.

Kummer, H. (1971) Primate Societies: Group Techniques of Ecological Adaptation.    Chicago: Aldine-Atherton.

Laland, K.N., et al. (2001) Cultural Niche Construction and Human Evolution. Journal of Evolutionary Biology, 14:22-33.

Lamarck, J.B. de (1963[1809]) Zoological Philosophy. Translated by H. Elliot.    New York: Hafner

Le Boeuf, B.J. (1985). “Elephant Seals.” In The Natural History of Año Neuvo, eds. B.J. Le Boeuf and S. Kaza. Pacific Grove, CA:Boxwood Press.

Le Boeuf, B.J., and R.M. Laws. (1994). Elephant Seals: Population Ecology, Behavior and Physiology. Berkeley, CA:University of California Press.

Leakey, R. (1994). The Origin of Human Kind. New York, NY:Basic Books.

Leigh, E. (1971). Adaptation and Diversity. San Francisco, CA:Freeman, Cooper & Co.

Leigh, E. (1977). “How Does Selection Reconcile Individual Advantage with the Good of the Group.” Proceedings of the National Academy of Sciences (USA), 74, 4542-4546.

Leigh, E. (1983). “When Does the Good of the Group Override the Advantage of the Individual?” Proceedings of the National Academy of Sciences (USA), 80, 2985-2989.

Leigh, E. (1991). “Genes, Bees and Ecosystems: The Evolution of a Common Interest Among Individuals.” Tree, 6, 257-262.

Leopold, C.A., and R. Ardrey. (1972). “Toxic Substances in Plants and the Food Habits of Early Man.” Science, 176, 512-14.

Lewis, D.H. (1991). “Mutualistic Symbioses in the Origin and Evolution of Land Plants.” In Symbiosis as a Source of Evolutionary Innovation, eds. L. Margulis and R. Fester, pp. 288-300. Cambridge, MA:MIT Press.

Little, P. (1995). “The Genome Directory: Navigational Progress.” Nature, 377, 288.

Lopez, B. (1978) Of Wolves and Men.    New York: Schribner.

Lorenz, L. (1966). On Aggression. (M.K. Wilson, trans.) New York:Harcourt, Brace, World.

Mann, J., et al., eds (2000) Cetacean Societies: Field Studies of Dolphins and Whales. Chicago: Chicago University Press.

Margulis, L. (1993). Symbiosis in Cell Evolution. 2nd ed. New York:W.H. Freeman.

Margulis, L., and K.V. Schwartz. (1982). Five Kingdoms: An Illustrated Guide to the Phyla of Life on Earth. San Francisco:W.H. Freeman.

Margulis, L., and R. Fester, eds. (1991). Symbiosis as a Source of Evolutionary Innovation: Speciation and Morphogenesis. Cambridge, MA:MIT Press.

Maryanski, A., and J.H. Turner. (1992). The Social Cage: Human Nature and the Evolution of Society. Stanford,CA:Stanford University Press.

Masters, R.D. (1989). The Nature of Politics. New Haven:Yale University Press.

Matessi, C., and S.D. Jayakar. (1976). Conditions for the Evolution of Altruism Under Darwinian Selection.  Theoretical Population Biology, 9, 360-387.

Maynard Smith, J. (1982). “The Evolution of Social Behavior – A Classification of Models.” In Current Problems in Sociobiology, ed. King’s College Sociobiology Group, pp. 28-44. Cambridge, MA:Cambridge University Press.

Maynard Smith, J. (1983). “Models of Evolution.” Proceedings of the Royal Society of London (B), 219, 315-325.

Maynard Smith, J. (1984). “Game Theory and the Evolution of Behaviour.” The Behavioral and Brain Sciences, 7, 95-125.

Maynard Smith, J. (1989). Evolutionary Genetics. Oxford:Oxford University Press.

Maynard Smith, J., and E. Szathmáry. (1993). “The Origin of Chromosomes I. Selection for Linkage.” Journal of Theoretical Biology, 164, 437-446.

Maynard Smith, J., and E. Szathmáry. (1995). The Major Transitions in Evolution. Oxford:Freeman Press.

Maynard Smith, J., and E. Szathmáry. (1999) The Origins of Life:  From the Birth of Life to the Origin of Language.    Oxford: Oxford University Press.

Mayr, E. (1960). “The Emergence of Evolutionary Novelties.” In Evolution After Darwin, Vol. I ed. S. Tax. Chicago: University of Chicago Press.

Mayr, E. (1963). Animal Species and Evolution. Cambridge, MA:Harvard University Press.

Mayr, E. (1976). Evolution and the Diversity of Life: Selected Essays. Cambridge, MA:Harvard University Press.

Mellars, P. (1989). “Major Issues in the Emergence of Modern Humans.”  Current Anthropology, 30, 349-385.

Michod, R.E. (1996). “Cooperation and Conflict in the Evolution of Individuality. II. Conflict Mediation.” Proceedings of the Royal Society of London (B), 263, 813-822.

Michod, R.E. (1997) Cooperation and Conflict in the Evolution of Individuality I. Multilevel Selection of the organism.  American  Naturalist, 149:607-45.

Michod, R.E. (1999)  Darwinian Dynamics, Evolutionary Transitions in Fitness and Individuality. Princeton, NJ: Princeton University Press.

Mithen, S.J. (1990). Thoughtful Foragers: A Study of Prehistoric Decision Making. Cambridge, MA:Cambridge University Press.

Modelski, G. (1987)  Long Cycles in World Politics. London: Macmillan.

Moritz, R.F.A., and E.E. Southwick. (1992). Bees as Superorganisms: An Evolution and Reality. New York, NY:Springer-Verlag.

Nowak, M., and K. Sigmund. (1993). “A Strategy of Win-Stay, Lose-Shift that Outperforms Tit-for-Tat in the Prisoner’s Dilemma Game.” Nature, 364, 56-58.

Odling-Smee, F.J. et al. (2003) Niche Construction: The Neglected Process in Evolution.  Princeton, NJ: Princeton University Press.

Odum, H.T. (1971). Environment, Power and Society. London:Wiley.

Oldroyd, B.P., et al. (1992a). “Intra-Colonial Foraging Specialism by Honey Bees (Apix Mellifera) (Hymenoptera: Apidae).” Behavioral Ecology and Sociobiology, 30, 291-295.

Oldroyd, B.P., et al. (1992b). “Effects of Intracolonial Genetic Diversity on Honey Bee (Hymenoptera: Apidae) Colony Performance.”  Annals of the Entomological Society of American, 85(3), 335-343.

Olendorf, R., et al. (2004) ACooperative Nest Defence in Red-Winged Blackbirds: Reciprocal Altruism, Kinship or By-product

Oyama, S. (2000) Evolution’s Eye: A Systems View of the Biology-Culture Divide. Durham, NC: Duke University Press.

Packer, C., and L. Ruttan. (1988). “The Evolution of Cooperative Hunting.” The American Naturalist, 132(2), 159-198.

Packer, C., et al. (1990). “Reproductive Success of Lions.” In Reproductive Success, ed. T.H. Clutton-Brock. Chicago:University of Chicago Press.

Page, R.E., and G.E. Robinson. (1991). “The Genetics of Division of Labour in Honey Bee Colonies.” Advances in Insect Physiology, 23, 118-169.

Peck, J.R. (1993). “Friendship and the Evolution of Co-operation in Honey Bee Colonies.” Journal of Theoretical Biology, 162, 195-228.

Pigliucci, M. (2001) Phenotypic Plasticity: Beyond Nature and Nurture. Baltimore, MD: Johns Hopkins University Press.

Pirozynski, K.A., and D.W. Malloch. (1975). “The Origins of Land Plants: A Matter of Mycotrophism.” Biosystems, 6,     153-64.

Plotkin, H.C., ed. (1988) The Role of Behavior in Evolution.    Cambridge, MA: MIT Press.

Polanyi, K. (2001/1944) The Great Transformation: The Political and Economic Origins of Our Time. Boston, MA: Beacon Press.

Price, P.W. (1991). “The Web of Life: Development Over 3.8 Billion Years of Trophic Relationships.” In Symbiosis as a Source of Evolutionary Innovation, eds. L. Margulis and R. Fester, pp. 262-272. Cambridge, MA:The MIT Press.

Quiatt, D.D., and V. Reynolds. (1993). Primate Behaviour: Information, Social Knowledge, and the Evolution of Culture. Cambridge, MA:Cambridge University Press.

Ratnieks, F.L.W., and P.K. Visscher. (1989). “Worker Policing in the Honey Bee.” Nature, 342, 796-797.

Rainey, P.B. and K. Rainey (2003) Evolution of Cooperation and Conflict in Experimental Bacterial Populations Nature, 425:72-74.

Richerson, P.J. and R. Boyd (2004) The Nature of Cultures. Chicago, IL: University of Chicago Press.

Rissing, S.W., and G.B. Pollack. (1991). “An Experimental Analysis of Pleometrotic Advantage in the Desert Seed-Harvester Ant Messor Pergandei (Hymenoptera: Formicidae).”  Ins. Soc., 38, 205-211.

Roe, A., and G.G. Simpson, eds. (1958) Behavior and Evolution.    New Haven, CT: Yale University Press.

Rollo, C.D. (1995) Phenotypes: Their Epigenetics, Ecology and Evolution. New York, NY: Chapman & Hall.

Rubin, Paul H. (2002) Darwinian Politics: The Evolutionary Origins of Freedom. New Brunswick, NJ: Rutgers University Press.

Salthe, S.N. (1985). Evolving Hierarchical Systems. Columbia University Press.

Scheel, D., and C. Packer. (1991). “Group Hunting Behaviour of Lions: A Search for Cooperation.” Animal Behaviour, 41, 697-710.

Schubert, G. (1989). Evolutionary Politics. Carbondale, IL:Southern Illinois University Press.

Schubert, G., and R. Masters, eds. (1991). Primate Politics. Carbondale, IL:Southern Illinois University Press.

Seeley, T.D. (1989). “The Honey Bee Colony as a Super-Organism.” American Scientist, 77, 546-553.

Seger, J. (1991). “Cooperation and Conflict in Social Insects.” In Behavioural Ecology: An Evolutionary Approach (2nd ed.), eds. J.R. Krebs and N.B. Davis, pp. 338-373. Oxford:Blackwell Scientific Publications.

Service, E.R. (1971). Cultural Evolutionism: Theory in Practice. New York:Holt, Rinehart, & Winston.

Shapiro, J.A. (1988). “Bacteria as Multicellular Organisms.” Scientific American, 258(6), 82-89.

Shapiro, J.A., and M. Dworkin, eds. (1997). Bacteria as Multicellular Organisms.  New York, NY:Oxford University Press.

Sherman, P.W., et al., eds. (1991). The Biology of the Naked Mole-Rat. Princeton, NJ: Princeton University Press.

Sherman, P.W., et al. (1992). “Naked Mole Rats.” Scientific American, 267(2), 72-78.

Sigmund, K. (1995) Games of Life: Explorations in Ecology, Evolution, and Behaviour.     Oxford: Oxford        University Press.

Sigmund, K. et al. (2002) “The Economics of Fair Play.”  Scientific American, 286(1): 83-87.

Simpson, G.G. (1953) The Baldwin Effect. Evolution, 2:110-117.

Smith, A. (1964/1776). The Wealth of Nations. 2 vols. London:Dent.

Smith, A. (1976/1759). The Theory of Moral Sentiments. Oxford:Clarendon Press.

Smith, D.C., and A.E. Douglas. (1987). The Biology of Symbiosis. Baltimore, MD:Edward Arnold.

Smith, D.C. (1992). “The Symbiotic Condition.” Symbiosis, 14, 3-15.

Smith, E.A., and B. Winterhalder, eds. (1992). Evolutionary Ecology and Human Behavior. New York, NY:Aldine Gruyter.

Sober, E. and D.S. Wilson. (1998) Unto Others:  The Evolution and Psychology of Unselfish Behavior.    Cambridge, MA: Harvard University Press.

Soltis, J., et al. (1995). “Can Group-Functional Behaviors Evolve by Cultural Group Selection?” Current               Anthropology, 36,  473-494.

Somit, A., ed. (1976). Biology and Politics: Recent Explorations. The Hague/Paris:Mouton.

Somit, A., et al. (1980). The Literature of Biopolitics. (Revised) Dekalb, IL:The Center for Biopolitical Research.

Sonea, S. (1991). “Bacterial Evolution Without Speciation.” In Symbiosis as a Source of Evolutionary Innovation, eds. L. Margulis and R. Fester, pp. 95-105. Cambridge, MA:The MIT Press.

Spencer, H. (1897/1874-1882). The Principles of Sociology, 3 vols., 3rd ed. New York:D. Appleton.

Sperry, R.W. (1969). “Modified Concept of Consciousness.” Psychological Review, 76, 532-36.

Sperry, R.W. (1991). “In Defense of Mentalism and Emergent Interaction.” The Journal of Mind and Behavior, 12(2), 221-46.

Stahl, A.B. (1984). “Hominid Dietary Selection Before Fire.” Current Anthropology, 25, 151-168.

Stander, P.E. (1992). “Cooperative Hunting in Lions: The Role of the Individual.” Behavioral Ecology and Sociobiology, 29,  45-454.

Stephens, D.W., et al. (2002) Discounting and Reciprocity in an Iterated Prisoner’s Dilemma. Science, 298:2216-2218.

Strum, S.C. (1987) Almost Human:  A Journey into the World of Baboons.    New York: Random House.

Tattersall, I. (1995). The Fossil Trail: How We Know What We Think We Know About Human Evolution. Oxford:Oxford University Press.

Tiger, L., and R. Fox. (1971) The Imperial Animal. New York: Holt, Rinehart & Winston.

Trivers, R.L. (1971). “The Evolution of Reciprocal Altruism.” Quarterly Review of Biology, 46, 35-57.

Trivers, R.L. (1985). Social Evolution.  Menlo Park, CA:Benjamin/Cummings.

Vanhanen, T. (1992). On the Evolutionary Roots of Politics. New Delhi:Sterling.

Van Horn, R., et al. (2004) Behavioral Structuring of Relatedness in the Spotted Hyena (Crocuta crocuta) Suggests Direct Fitness Benefits of Clan-level Cooperation. Molecular Ecology, 13:449-458.

Van Schaik, C.P. et al. (2003) Orangutan Cultures and the Evolution of Material Culture. Science, 299:102-105.

Vetter, R.D. (1991). “Symbiosis and the Evolution of Novel Trophic Strategies: Thiotrophic Organisms at Hydrothermal Vents.” In Symbiosis as a Source of Evolutionary Innovation, eds. L. Margulis and R. Fester, pp. 219-245. Cambridge, MA:MIT Press.

de Waal, F.B.M. (1982). Chimpanzee Politics: Power and Sex Among Apes.  New York:Harper & Row.

de Waal, F.B.M. (1996). Good Natured: The Origins of Right and Wrong in Humans and Other Animals. Cambridge, MA:Harvard University Press.

Wade, M.J. (1977). “An Experimental Study of Group Selection.” Evolution, 31, 134-153.

Wade, M.J. (1985). “Soft Selection, Hard Selection, Kin Selection, and Group Selection.” The American Naturalist, 125(1),  61-73.

Weber, B.H., and D.J. Depew, eds. (2003) Evolution and Learning: The Baldwin Effect Reconsidered. Cambridge, MA: MIT Press.  West-Eberhard, M J. (2003) Developmental Plasticity and Evolution. Oxford and New York: Oxford University Press.

West-Eberhard, M.J. (2003). Developmental Plasticity and Evolution. Oxford: Oxford University Press.

Wheeler, W. M. (1927a) “Emergent Evolution of the Social.” In Proceedings of the Sixth International Congress of Philosophy, ed. E.S. Brightman.   New York: Longmans, Green and Co.

Wheeler, W. M. (1927b) Emergent Evolution and the Social. London: Kegan Paul, Trench, Trubner.

Wheeler, W.M. (1928). The Social Insects: Their Origin and Evolution. London:Kegan, Paul, Trench, Trubner & Co.

White, L.A. (1943). “Energy and the Evolution of Culture.” American Anthropologist, 45, 335-356.

White, L.A. (1949). The Science of Culture: A Study of Man and Civilization. New York:Grove Press.

Whiten, A., et al. (1999)    Cultures in Chimpanzees. Nature, 399: 682-685.

Whitehead, H. and  L. Rendell (2004) Movements, Habitat Use and Feeding Success of Cultural Clans of South Pacific Sperm Whales. Journal of Animal Ecology, 73:190-196.

Wiegele, T.C. (1979). Biopolitics: Search for a More Human Political Science. Boulder, CO:Westview Press.

Wilkinson, G.S. (1984). “Reciprocal Food Sharing in the Vampire Bat.” Nature, 308, 181-184.

Wilkinson, G.S. (1988). “Reciprocal Altruism in Bats and Other Mammals.” Ethology and Sociobiology, 9, 85-100.

Wilkinson, G.S. (1990). “Food Sharing in Vampire Bats.” Scientific American, 262(2), 76-82.

Willhoite, F.H., Sr. (1976). “Primates and Political Authority: A Biobehavioral Perspective.”  The American Political Science Review, 70, 1110-1126.

Williams, G.C. (1966). Adaptation and Natural Selection: A Critique of Some Current Evolutionary Thought. Princeton, NJ:Princeton University Press.

Williams, G.C. (1992). Natural Selection: Domains, Levels, and Challenges.  New York:Oxford University Press.

Wilson, D.S. (1975). “A General Theory of Group Selection.” Proceedings of the National Academy of Sciences (USA), 72, 143-146.

Wilson, D.S. (1980). The Natural Selection of Populations and Communities. Menlo Park, CA:Benjamin/Cummings.

Wilson, D.S., and L.A. Dugatkin. (1997). “Group Selection and Assortative Interactions.”  The American Naturalist, 149,   336-351.

Wilson, D.S., and E. Sober. (1989). “Reviving the Superorganism.” Journal of Theoretical Biology, 136, 337-56.

Wilson, D.S., and E. Sober. (1994). “Reintroducing Group Selection to the Human Behavioral Sciences.” Behavioral and Brain Sciences, 17, 585-608.

Wilson, E.O. (1975). Sociobiology: The New Synthesis. Cambridge, MA:Harvard University Press.

Wilson, E.O. (1985). “The Sociogenesis of Insect Colonies.” Science, 228, 1489-1495.

Wrangham, R. W., et al., eds. (1994) Chimpanzee Cultures.    Cambridge, MA: Harvard University Press.

Wrangham, R. W. (2001) “Out of the Pan, Into the Fire: How Our Ancestors’ Evolution Depended on What They Ate.” In Tree of Origin: What Primate Behavior Can Tell Us about Human Social Evolution, ed., F.B.M. de Waal, pp.121-143.   Cambridge, MA: Harvard University Press.

Wrangham, R.W., et al. (1999) “The Raw and the Stolen:  Cooking and the Ecology of Human Origins.”  Current Anthropology, 40(5):567-594.

Wright, R. (2000)  Nonzero: The Logic of Human Destiny. New York: Pantheon Books.

Wright, S. (1968-1978). Evolution and the Genetics of Populations: A Treatise.  4 vols. Chicago:University of Chicago Press.

Würsig, B. (1988). “The Behavior of Baleen Whales.” Scientific American, 258(4), 102-107.

Würsig, B. (1989). “Cetaceans.” Science, 244, 1550-1557.

Wynne-Edwards, V.C. (1962). Animal Dispersion in Relation to Social Behaviour. New York:Hafner.

Wynne-Edwards, V.C. (1963). “Intergroup Selection in the Evolution of Social Systems.” Nature, 200, 623.

Yurk, H., et al. (2002) Cultural Transmission Within Maternal Lineages: Vocal Clans in Resident Killer Whales in Southern Alaska. Animal Behaviour, 63:1103-1119.

 

 

 

Category: Publications