Go to the Nature Institute’s home page

Organisms and Their Evolution
A book by
Stephen L. Talbott

Chapter 21

Inheritance, Genetics, and the Particulate View of Life

book home page RSS Feed: RSS Feed for this page   

This is a preliminary draft of one chapter of a book-in-progress entitled, “Organisms and Their Evolution — Agency and Meaning in the Drama of Life”. This material is part of the Biology Worthy of Life project of The Nature Institute. Copyright 2022 by Stephen L. Talbott. All rights reserved. You may freely download this chapter for noncommercial, personal use, including classroom use.

Tags: evolution/and development; evolution/inheritance; evolution/and mutations; explanation/part and whole; gene/mutation; gene/genocentrism; holism

This chapter about the gene-centered (“genocentric”) view of evolving organisms shouldn’t have needed to be written. Today genocentrism has been challenged from so many different sides, by so many leading biologists, and in such an ever more insistent manner, that it might easily seem a waste of time to raise the usual issues afresh here. So I won’t.1

But there are caveats. One is that, despite the criticism, the idea of the masterful, controlling gene remains as strongly entrenched as ever in the minds of most biologists. This is especially true of evolutionary theorists, for whom the word “genetic” has long been synonymous with “heritable”. In other words, for purposes of evolutionary theory genes substitute for the entire, one-celled living being that passes between generations. This means that, as participants in an inheritance-based evolutionary lineage, organisms themselves scarcely exist for the theorist.

The century-long habit of genocentrism is seemingly resistant to all criticism. As three Duke University biologists summarized the matter in 2017, “Everyone understands” that the idea of a definitive gene for this or that feature of an organism “is a distortion of the biological facts, yet, as a profession, we have yet to rid ourselves of this crutch” (Gawne, McKenna and Nijhout 2018).

Much of the criticism of genocentrism has arisen from the field of evolutionary-developmental biology (“evo-devo”). Yet even here, according to a leader in that discipline, “increasing gene centrism characterizes the field today”:

This reductionist attitude continues to be upheld, even though overwhelming evidence points to the fact that it is not gene expression and regulation that singularly define body structures but the systemic processes of interaction between genes, cells, and tissues as well as the physics and physiologies of the involved entities and their interactions with numerous factors of the environment (Müller 2019).

A second caveat, even more discouraging, is that the critics themselves leave the door wide open for the persistance of genetic reductionism. This is because few if any well-positioned, reputable biologists are willing, at the risk of reputation and career, to speak out against the reigning materialist dogma of their profession.2

What isn’t generally recognized is that this dogma invites, as an inevitable counter-movement, an almost worshipful regard for the all-determining, informational gene — a machine-like gene intelligently designed and engineered from outside by the “creative forces” of evolution. In this way the theorist employs the gene as a lifeless stand-in for the present and effective wisdom that moment by moment lives and expresses itself in every animal’s inner, qualitative, perceptual experience and in its thought-full response to that experience.

In what follows I will review the genocentrism of the received evolutionary theory, and then explore some fundamental problems with genocentrism — problems the present critics of genocentrism cannot allow themselves to recognize at risk of violating the materialist taboo. This exploration will continue into the next chapter, where we will look at the contrast between the evolutionarily stable (“potentially immortal”) genetic particles that Richard Dawkins celebrates so vehemently, and the evolutionarily stable yet dynamically transformative whole organism.

DNA as the essential
substance of evolution

As we saw in Chapter 16 (“Let’s Not Begin With Natural Selection”), evolution is said to be inevitable once three conditions are met: (1) There must be trait variation among individuals in a breeding population; (2) This variation must to some degree be inherited, so that offspring generally resemble their parents more than they resemble others; and (3) Individuals possessing different variants of a trait must, at least in some cases, exhibit differential fitness (or differential survival) — that is, they must produce, on average, different numbers of offspring, whether immediate offspring or later descendents. This is often referred to as survival of the fittest, or the principle of competition.

We also saw in that earlier chapter that these three conditions — insofar as they are abstracted from the life of the organism and its agency — become a hollow formula that tells us nothing about why a species evolves in the direction it does. That is, until we reckon with what organisms do as a function of their entire way of being, the so-called “core logic”, or “algorithm”, of evolution can tell us nothing at all about their evolution or about the diversity that may arise from them — or even about whether they will evolve at all. And when we do reckon with the active life of organisms, it is this reckoning itself, not some core logic of inheritance and selection, that elucidates the evolutionary trajectory of a species.

The chief excuse for ignoring what organisms do has been found in a distorted picture of genes and DNA. This picture gives to the core logic described above what little appearance of biological content it has:

So genes are the one unquestioned material foundation and efficient cause upon which, from the organism’s side, the modern edifice of evolutionary theory has been erected. As defined in a classic introductory text, the process of evolution “includes all mechanisms of genetic change that occur in organisms through time...” (Hartl 1988, p. 143). Evolution, in this view, looks very much like a matter of genes and their fate — and not much more.

The aggressive claims of population geneticists

The author of that last quotation was a population geneticist, and it was preeminently the population geneticists who framed the twentieth-century “Modern Synthesis” as the definitive formulation of evolutionary theory. It is likewise the population geneticists who still today speak most forcefully about the primacy of genes. And it is the population geneticists (regarding themselves as working at the pinnacle of evolutionary theory) who continue to dominate the evolutionary field today, if only by virtue of their long-established control over the institutional levers of power.

Their life’s work has always centered on calculations relating to the transmission of genes between parents and offspring, and the changing distribution of genes within populations. The more advanced, technically impressive results of this work often take the form of sophisticated equations that tend to be more or less opaque to working biologists outside the field of population genetics.

Here are two examples of comments from population geneticists:

Michael Lynch, who holds the Distinguished Professorship of Evolution, Population Genetics and Genomics at Indiana University and was formerly president of the Genetics Society of America, has remarked that “the litmus test for any evolutionary hypothesis must be its consistency with fundamental population-genetic principles”. He freely admits that organisms themselves, as “phenotypic products”, result from “more than a change in gene frequencies”. But the crucial conclusion remains, if only by brute assertion: “If we are concerned with the process of evolutionary change, then evolution is indeed a change in genotype [gene] frequencies” (Lynch 2007a; Lynch 2007b, p. 371).

And then there is Dan Graur, author of a textbook on Molecular and Genome Evolution, who proceeds in the same confidently dogmatic spirit (to which he adds his own unique brand of arrogance, in which — judging by the larger body of his work — he apparently revels):

Evolutionary biology is a mature science. It is a coherent discipline with a handful of logical principles, each of which repeatedly withstood rigorous empirical and observational testing. Evolution is not difficult to define. If one ignores the obfuscations of the creationists, the casuistry of the philosophers, and the ruminations of the “sophisticates,” evolution turns out to be merely the process of change in allele [gene] frequencies over time. The only mandatory attribute of the evolutionary process is a temporal change in allele frequencies (Graur 2015).

This is truly amazing — a stunning contraction of human understanding among the would-be elder statesmen of evolutionary theory, smugly satisfied that they do indeed sit at the pinnacle of their discipline. The organism’s entire way of being along with its needs, interests, and agency have been reduced, for purposes of evolutionary theory, to one material part — DNA. The underlying drive appears to be the reduction of mindedness to mindlessness, logos to meaninglessness, telos to chance — all reflecting a horror of interiority.

The habit of ignoring organisms is so stubbornly entrenched among population geneticists that their ability to recognize crucial evolutionary issues seems scarcely to exist. Think, for example, of the problem of the origin of innovation — that is, the problem of the “arrival of the fittest”, or the arrival of new, living performances (traits) for natural selection to act upon. (We looked at this in Chapter 16.) Armin Moczek, a prominent evolutionary developmental biologist at the University of Indiana, remarked that “fields such as population-genetics have long stopped asking the question how evolution innovates, not because it is not a foundational question in evolutionary biology, but because population genetics lacks the ability to even frame the question” (Moczek_2022).

They have stopped asking in part because the fact that genes can make a more or less stable difference in existing traits (Chapter 22) has convinced them that they need not consider all the other, less experimentally and mathematically tractable features of an organism’s life that also make a difference. Nor need they inquire into the principles of organization through which cellular and organismal identity are stably maintained, and through which alone a trait can come into existence as an integral and viable aspect of the larger whole. And they are least of all inclined to consider whether the organism’s most dramatically demonstrated capacity — its capacity for directed developmental change and metamorphosis consistent with its own way of being — might be relevant to evolution.

Who defines what counts as an evolutionary process?

The population geneticists, in the manner of those adhering to many an aging and rickety scientific viewpoint, have constructed for themselves an institutional and intellectual fortress whose final crumbling must, as the rather brutal saying has it, await the dying out of its last, well-known defenders.

In the meantime, one thing giving a sense of impregnability to the fortress is the fact that evolution has been aggressively defined in terms of genes, as we heard above. One is then bound to argue that whatever is not adequately gene-like cannot be important for evolution.

This argument by definition is why population geneticists routinely dismiss epigenetic factors (Chapters 7 and 14) as irrelevant to evolution: such factors, they tell us, often don’t have the long-term, transgenerational stability usually ascribed to genes, and therefore can’t contribute much to evolution.3

In other words, “Since epigenetics doesn’t give us the kind of genetic stability we want to see as the essence of a mindless sort of evolution, we refuse to consider what evolutionary potentials it does give us”. Those who think this way can scarcely imagine that epigenetics presents us with a revealing expression of the highly adaptive processes of continual, directive change we discover in every sort of whole-organism activity — for example, in the many differentiating cell lineages of our own developing bodies. These lineages result from the changing organization of whole cells, which includes their changing ways of employing their genes.

Such transformative processes — which one might think would be the first things looked for by evolutionists — are ruled out of evolutionary theory on the grounds (now known to be false) that nothing occurring in development affects the genetic inheritance of the next generation. But even if that were true, so what? It overlooks the entire nature of cellular inheritance during cell differentiation. This inheritance is not only dramatic, but undeniably a whole-cell phenomenon. If biologists dismiss the significance of developmental processes for evolution, it is only because they have planted their flag and staked their claim in advance: “We’re not interested in the potentials of the whole cells that contribute an inheritance to the next generation, but only in the genes those cells contain”.

As with so many perverse doctrines, there is a dim and distant reflection of the truth in the gene’s-eye view of evolution, although it is a truth lost on Lynch, Graur, and their kin. DNA is indeed caught up in, and informed by, the character of the whole organism, including its adaptive character. Therefore we can in one way or another expect to find the whole organism reflected in DNA. Such is the case with all the other major aspects of any organism, as we have already seen in our discussion of cell membranes and the cytoskeleton (Chapter 4).

Getting to the
bottom of things?

One of the most common strategies for honoring the materialist taboo in all sciences is to describe a microscopic level of meaningless and inherently inert, mindless things, or particles, possessing fixed, well-defined natures with causal powers. (The fact that causal powers are powers, not things — so where in the material world do they come from? — is conveniently ignored.) Then one claims that whatever really counts in the explanation of phenomena derives from various effective combinations of these particles.

In Richard Dawkins’ biological theorizing, the particles at issue are genetic elements with a wonderfully computational nature. “Digitalness”, he has said in what must have been one of his rashest statements, “is probably a necessary precondition for Darwinism itself to work” (Dawkins 2006, p. 163). “What is truly revolutionary about molecular biology”, he wrote, “is that it has become digital”. We know that genes “are long strings of pure digital information … The machine code of the genes is uncannily computerlike. Apart from differences in jargon, the pages of a molecular-biology journal might be interchanged with those of a computer-engineering journal” (Dawkins 1995, pp. 17-19).

The meaning of “digital” can be illustrated by the game of basketball. The game is so designed that the making of a basket is always definite and unquestioned. The ball either goes through the hoop, or it does not. A team either earns points for the basket or it does not. How graceful or awkward, skillful or random the shot may have been has no bearing on the matter. A player’s approach to the basket can be ugly as sin, but if the ball ends up going through the hoop, the points are counted. There’s a clean, yes-or-no, “ones-and-zeros” aspect to the proceedings. The number of points earned is always exact and countable. Everything is precise, and there is no ambiguity.

Dawkins’ claim accordingly is that, when it comes to the identity and role of genetic particles, “there are no half measures and no intermediates or compromises” — certainly nothing we might recognize as a power of self-transformation. “Our particles of inheritance … don’t blend, but remain discrete and separate as they shuffle and reshuffle their way down the generations” (Dawkins 2006, pp. 159-63).

So this is what Dawkins assumes to be the essential character of our genetic material, which he situates at the causal bottom of every organism: it consists of discrete and separate causal elements that do not interpenetrate, neither are they subject to half measures, intermediates, or compromises. His reference to genes as strings of “pure digital information” makes it clear that his “causal” factors come very close to being elements of computer logic.

The convergence of the two ideas of (1) causal force and (2) compelling, computational logic in Dawkins’ mind — a symptomatic conflation running through much biology today — is ironic in light of his materialist commitments. If in fact the pure conceptuality of logic finds revelatory application in science, it would only reveal a world causally governed or organized according to thought. The only problem is that, whatever we may say about the inanimate world, the very real mindfulness we encounter in our concrete and engaging interactions with organisms is far more profound — and far less one-dimensional — than a mere play of abstract logic.

It hardly needs saying that Dawkins’ genetic informational bits are, by definition, incapable of participating in living wholes. Because their identity is conceived as almost eternally fixed and quantifiably (digitally) specifiable, they cannot lend themselves to being qualitatively transformed or reimagined in harmony with the fluid, organizing ideas and intentions at work in every organism.

But Dawkins gets his digital, non-blending particles of inheritance only by defining them into existence. Reality gives us a very different story.

How the image of “particles” has distorted the biologist’s imagination

Digital, unblending genetic elements of pure, computational logic or information: where is one even to begin a critique of these impossible notions, upon which so much contemporary evolutionary theory is erected? Where in the living organism do we find the slightest justification for them? Can Dawkins show us even one DNA sequence that functions in a strictly digital fashion?

schematic sketch of a molecule

Figure 21.1. Schematic representation of a molecule, where concentric “circles” are intended to map electron distributions of individual atoms, and straight lines artificially represent bonds between the atoms.4

We might start thinking about this at a fairly remote distance from Dawkins’ immediate genetic meanings by looking at an illustration offered by twentieth-century cell biologist, Paul Weiss. He reminds us of the commonplace pictures and models of molecules, represented as conglomerates of colored, spherical “billiard balls”, each standing for an atom, and each seeming to be a stable, self-contained unit. He then contrasts that with the image in Figure 21.1, which shows how the constituent atoms of a molecule interpenetrate (“blend” into) each other as continuous fields or a complex system of mutually shaped forces. These “blur the former sharpness of the outer boundary of the molecule and let it melt into the surrounding molecular domains” (Weiss 1971a, p. 9).

Weiss also remarks of figures such as this that “one is reminded of the contour maps of mountain ranges. Domains of particles are no more truly isolated than are mountain peaks” (Weiss 1971b, p. 235). Moreover, the parts of such an interactive system are rather “like islands”, so that they must be “conceived as interconnected, though not so solidly as by a bed of rock, but loosely by the all-pervading mesh of forces and interactions” (Weiss 1971a, p. 11).

This may remind us of the discovery (mentioned in Chapter 5) of how water interacts with DNA. Lifting one paragraph from that earlier chapter:

Early efforts to develop a computer simulation of a DNA molecule failed; the molecule (in the simulation) almost immediately broke up. But when [the researchers] included water molecules in the simulation, it proved successful. “Subsequent simulations of DNA in water have revealed that water molecules are able to interact with nearly every part of DNA’s double helix, including the base pairs that constitute the genetic code”.

Needless to say, this interaction of DNA with water alters the entire landscape (or seascape) of DNA, very much in the manner of Figure 21.1 above. None of it suggests much of a defense “at the bottom” for the idea of independent, digital, unblending genetic particles. But this point will become more explicit and more directly aimed at genes as we move along.

The fundamental reality of the matter was already glimpsed in the nineteenth century by the great experimental physicist, Michael Faraday. He recorded his prescient, “final brooding impression” that “particles are only centres of force” and that, for instance, water does not consist of atomic particles “side by side”, but rather of “spheres of power mutually penetrated”. And in notes for a talk given to the Royal Institution, he suggested that “matter is not merely mutually penetrable, but each atom extends, so to say, throughout the whole of the solar system, yet always retaining its own centre of force” (quoted in Barfield 1971, pp. 201n11 and 244n17).

Facing up to the gestural reality of DNA

What, then, do we see when we take even a cursory look at DNA? Exactly what Faraday and Weiss would have expected.

We have already heard about the positive and negative supercoiling of the double helix (Chapter 3). Both reflect a dramatic re-organization of forces along the DNA molecule — forces from which neither genes nor single nucleotide bases (“letters”) of DNA can claim to be immune. These forces are not respecters of the boundaries between genes, and their re-shaping effects are known to help determine the functional role of genes within the organism.

But supercoiling is only one of many ways the genome is continually restructured. The DNA molecule, while inherently rather stiff and inert, lends itself, under cellular influences, to endless, plastic, structural change, both subtle and not so subtle. Consider, for example, the many protein transcription factors that come to bear upon gene expression. In the simplistic thought of an earlier day, they were assumed to be just about the sole factors implementing the rigid “control logic of the genetic program”. Moreover, they were said to bind DNA in a manner strictly determined by the abstract sequence of DNA “digital letters”, rather than by gesturally engaging in a mutual dance with the plastic matrix of physical forces constituting the reality of that sequence. The determining role of the abstract sequence was then supposed to justify the geneticists’ belief that genes were the real agents in control of their own expression. But this conclusion sounds farcical today.

Transcription factors are now well known to contribute their own forces to the infinite variations in the way genetic sequences “blend” together with each other and with innumerable regulatory molecules. Actually, even within the terms of the earlier view, there were all sorts of unanswered questions about transcription factors. For example, what determined when and how often a particular factor was bound to the supposedly controlling DNA sequences? Why did it bind more to some of those sequences and less to others? And how was its activity coordinated with that of the many other transcription factors participating in the complex task at hand, which might involve the need for balanced expression of hundreds of genes?

Beyond this, however, the old concept of transcription factor binding (or not binding) to DNA is now recognized as hopelessly one-dimensional. The reality of DNA is difficult to think of as anything other than a play of ongoing gestural form orchestrated by the cell as a whole. You don’t need to know the complex details of the terminology in order to get a sense for what it means when an article in the journal Nature reports how transcription factors interact with “shape features of the DNA sequences, such as minor-groove width, roll, propeller twist and helical twist” (Burgess 2015). These ever-changing shape features are subject to influences arising from the limitless cloud of regulative molecules (including water molecules, as we heard above, as well as the transcription factors themselves) that more or less transiently swirl around a cell’s DNA. None of the analog shape features looks very much like Dawkins’ concept of digital or computer-like elements of a genetic program.

The vast majority of our DNA is tightly and forcibly bent so as to wrap around millions of nucleosome core particles in the cell nucleus. The nucleosome, as we saw in Chapter 14 (“How Our Genes Come to Expression”), is perhaps the central integrator of signals bearing on gene expression and coming from all corners of the cell.

The wrapping of DNA around these protein core particles is probably never exactly the same in the case of any two of the thirty million or so nucleosomes in our genome, nor ever the same at any two different times in the case of a single nucleosome. The bending and shape changes in the DNA require a great variety of force interactions between DNA and the core particle, all in the presence of numerous associated molecules. So we hear (as before, in the case of transcription factors) that the functionally critical spacing and location of nucleosomes along a stretch of DNA can vary, depending on transient DNA features imposed by the cell “such as propeller twist, opening, electrostatic potential, minor groove width, rise, stagger, helix twist, and shear and roll ... and buckle” (Kurup 2019). Everything is subject to dynamic variation.

The gestural form we are looking at in the case of DNA and related molecules appears to have no limit in its extent or its significance for the cell. It cannot even be said that the double helix is always even close to being a standard double helix, or a double helix at all:

Alternative conformations (including left-handed DNA, three-stranded triplex DNA, four-armed cruciforms, slipped-strand DNA with two three-armed junctions, four-stranded G-quadruplex structures and stable, unpaired helical regions) can exist in the context of chromosomes. Rather than being a static helix, DNA possesses dynamic flexibility and variability, as evidenced by helix regions that can be curved, straight or flexible. Differences result from variations in base stacking and twist angles inherent in different DNA sequences. DNA supercoiling [induced, for example, by transcriptional enzymes], particularly unconstrained supercoiling,5 plays a major part in the dynamic flexibility and topological contortions of the DNA double helix (Sinden 2013).

Everything we have been hearing about is fundamentally qualitative and gestural rather than digital, and it counts not only toward whether a “basket is scored” (a gene is expressed), but also toward the “size of the score” — for example, will there be a large, rapid burst of expression, or a low-level, steady-state expression? There is also the question of the nature of the score: for example, which of several functional variants of a protein will be synthesized? So the qualitative, gestural features mentioned above bear hugely on the practical meaning of a gene for the organism. They help to define what any given gene is.

DNA, when caught up within the whole cell, is a phenomenon of movement. Of the endless variety of its movements, two are referred to as “DNA breathing”. One of these (which we heard about in Chapter 3) involves the rhythmic loosening of parts of the double helix from the nucleosome core particles around which they are wrapped. This has a great influence on the accessibility of portions of the DNA to gene-regulatory proteins. The other sort of breathing consists of local, transient separation of the two strands of the DNA double helix, which also affects the accessibility of the DNA.

Furthermore, the foregoing represents only a minuscule introduction to all the ways gene expression turns out to be a fluid, non-digital, and non-machinelike expression of dynamic, gestural form. I have not even mentioned what is widely considered to be the most prominent way DNA “letters”, or nucleotide bases, are modified and the play of forces re-sculpted — namely, by the attachment of methyl groups to nucleotide bases in a process called “DNA methylation”. This is the most common of a fair number of modifications to DNA, affecting many millions of nucleotide bases in our genomes. By this means the letters become different letters.

Neither have I mentioned the large range of factors affecting the structure of RNA, an essential molecular carrier of the “genetic code” whose liveliness of functional form is, if anything, even more obvious than what we see in DNA.

And, perhaps most important of all, I have not cited the massive research effort today dealing with the form and movement of chromosomes — for example, the critical looping movements that bring genes and regulatory elements of DNA into functional proximity within the three-dimensional space of the nucleus. We are looking here at a gestural performance that many investigators can hardly resist referring to as a “dance” or an elaborate “choreography”. We encountered some of this in Chapter 3 (“What Brings Our Genome Alive?”). How that dance occurs critically shapes how genes will be expressed. Genes are as far from being discrete, well-defined, independent causes as two ballet dancers engaged in a pas de deux.

A gene that, within the full contextual life of the organism, can be interpreted as a determinate cause or a bearer of strict digital logic does not exist.

To view from the bottom
or view from the top?

Evolutionary theorists, so it appears, have great difficulty recognizing as significant either the stable and highly distinctive character of the whole cell and whole organism, or the remarkable interior capacities through which that character is consistently expressed and sustained amid all the transformations of individual development. And so they find it easy to discount everything living. They discount, that is, the future-directed powers of self-realization, adaptation, and whole-cell (whole-organism) reproduction — the very powers that hold the most obvious relevance to inheritance and evolution.

Organisms as such simply don’t show up through the death-shroud that is the particulate view of life. As for the particles themselves — the supposedly unblending, unexpressive, qualitatively inert genes and nucleotide bases (“letters”) — they are, in reality, illusions. The upshot of the foregoing discussion is that the discrete, non-blending, genetic particles that Dawkins, for example, is so quick to idealize as controllers of the organism’s evolution do not actually exist as real, material entities. They exist only as logical or digital constructs abstracted away from the living cell.

The situation can also be summarized by saying that evolutionary biologists are currently blind to organisms as beings organized from within. To speak of interiority is anathema to them, and therefore any genuine recognition of organizing ideas is also anathema. Which is too bad, because either the material world is such that organizing ideas have real consequences, or else the terms biologists do so freely use — “organizers”, “organization”, “self-organization” — are symptoms of shamefully vacuous theorizing.

Think of it: take away the organizing ideas, and what would be left of any sort of “organizing” or “organization”?6 (Or, for that matter, what would be left of any science?) “Imposing a kind of meaningful coherence” is just what the word “organizing” means.

Consider the much-criticized but still endemic idea that there are genes for particular traits. If we believe that genes possess, in their own right, the essential, organizing or directing power to realize traits — traits that are in fact qualitative, non-discrete, interpenetrating, and expressive of a specific (species-related) “way of being” — then we are ascribing to genes a living power to organize almost unthinkably complex physiological processes requiring a kind of moment-by-moment active judgment governing a virtually infinite number of molecular interactions in a fluid, continually changing context. (I focused especially on such processes in Chapter 8).

Genes surely do participate in such a power, but it is a power of activity belonging to the whole organism and is not properly attributed to any collection of material elements, such as genes, that are caught up in it.

Those who want to adhere to materialist principles gain nothing by contradicting them. They gain nothing, that is, by transferring the interior, organizing ideas of the cell or organism to genes. If genes really possessed their own “informational” powers for sensing their wider context; if they really could encode messages tuned to moment-by-moment changes under infinitely varying circumstances; if they really had a way to direct the interactions among countless billions of molecules in a fluid medium, enabling those molecules to carry out indescribably intricate operations such as RNA splicing (Chapter 8); and if in general they really were able to inform and organize the life of the entire cell7 … well, once we have accepted this impressive play of wisdom through a material genome conceived in a particulate manner, what reason would remain for denying it to the vividly expressive cell or the whole organism, where in fact we observe it?

If “context matters”, as so many biologists are now telling us, it can only be because it really does substantively matter. It makes a difference to what happens. The context exercises, in other words, its own, over-arching sort of causal power. It’s not a kind of power that can be transferred to materialistically conceived particles — particles abstracted from cell and organism as if they they carried a decisive causal logic independent of the living matrix in which they find their real existence.

Where are we now?

Looking Beyond Particulate Inheritance

Biology today is governed by a taboo: The biologist must never acknowledge stepping outside the materialist framework — or at least must remain unaware of doing so. She must never grant that animals have an interior — that every animal is an integral, unified whole possessing, or possessed by, an active, wise agency.

The biologist’s materialist commitments are impossible to keep. They would render the organism, as a living being, invisible to scientific investigation. This is why biologists can hardly avoid preserving the organism’s agency by covertly transferring it to special molecules (DNA). But at the same time — in order to keep an illusion of observing the taboo — they speak of these wonderfully effective, “informational” molecules as if they were “things” uninformed and ungoverned by the agency of the whole.

This double-talk, which would have DNA possessing the creative powers of life while at the same time consisting of “mere chemical stuff”, is somehow easier to stomach in the case of molecules than in the case of whole organisms. Molecules, being non-phenomenal (invisible to sense perception), lend themselves more obligingly to the projection of our mechanistic/animistic fantasies.

It is population genetics, above all else, that has converted evolutionary theory into a theory about genes rather than organisms. As we heard from one population geneticist, “The only mandatory attribute of the evolutionary process is a temporal change in allele [gene] frequencies”.

There is little beyond quantifiable (digital or logical) entities in this picture — nothing material, plastic, and expressive, nothing qualitative, nothing through which the interior life of organisms can shine. We cannot connect anything in the particulate gene to our own conscious awareness, or to animal sentience, or to perceptual experience and cognition, or to the organizing ideas underlying animal form and behavior. If particulate genes account for these aspects of life, no one has a clue how it could be. Yet these genes are routinely posited as the evolutionary basis for understanding all life.

In the next chapter we will, finally, consider the fervent and influential defense of the “gene’s-eye view of evolution” offered by Richard Dawkins.

Notes

1. For the record, currently relevant criticisms of genocentrism go all the way back a century. There is, for example, the brilliant work by marine biologist E. S. Russell, especially his 1930 book, The Interpretation of Development and Heredity: A Study in Biological Method, which I discussed in the previous chapter. In the modern era, one could start with Exploding the Gene Myth by Harvard professor of biology Ruth Hubbard and Nobel Prize recipient Elijah Wald, or the essay, “Unraveling the DNA Myth”, by cell biologist Barry Commoner.

The physicist, biologist, and philosopher of science, Evelyn Fox Keller, has illuminated genocentrism from many sides, including in her book, The Century of the Gene, published in 2000, and her chapter on Genes as Difference Makers in 2013. The one-time molecular biologist and now philosopher of science, Lenny Moss, wrote an incisive and influential critique titled What Genes Can’t Do in 2003. And his book chapter, “Darwinism, Dualism, and Biological Agency” (2005), has perhaps never been exceeded for the succinctness and penetrating depth of its take-down of the controlling gene.

There are many other worthy commentaries on genocentrism, of which one (Gawne, McKenna, and Nijhout 2018) is cited below. See also Holdrege 1996, Rose 1998, Moczek 2012, Noble 2013, Walsh 2015, Noble 2018, and any number of other books and journal articles published over the past couple of decades.

Special mention also goes to University of Chicago microbiologist James Shapiro’s book, Evolution: A View from the 21st Century (second edition, 2022). Evolution can only occur if there is useful variation, or potentials for variation, in the evolving organisms. And Shapiro provides overwhelming evidence that, whatever whole-organism features may count as viable heritable variation, organisms certainly have the genetic aspect very well covered. That is, they possess a sophisticated and wide-ranging ability to revise their own DNA — and they put it to use in a huge variety of ways. One could prefer that Shapiro not rely so heavily on computer and program metaphors, but nevertheless he makes abundantly clear the organism’s effective exercise of a well-directed agency with respect to its DNA.

2. Disappointingly, those who continue giving support to genetic reductionism include proponents of the “extended evolutionary synthesis” and the “third way of evolution”. As near as I can tell, these movements remain as thoroughly materialist in their fundamental assumptions as the evolutionary mainstream. It truly does appear that any questioning of the materialist dogma in biology is likely to spell the immediate end of an otherwise promising career. The existence of such metaphysical dogma in science ought to be ringing alarm bells on all sides.

3. For a discussion of the question of stability, see Chapter 22, “A Curiously Absolute Demand for Stable Variation”.

4. Figure 21.1 credit: from Weiss 1971a.

5. Unconstrained supercoiling is supercoiling that creates tension. That is, in the unconstrained case the tension that results from supercoiling is not relieved by means of single-strand DNA breaks, the binding of proteins, or other means. I offer a very brief explanation of supercoiling in Chapter 3 (“What Brings Our Genome Alive?”).

6. To speak of organizing ideas at work in an animal’s life is not to imagine the animal thinking them. We will look at the distinction between an organism possessing its own intelligence and being possessed by it in our concluding chapter, “Some Principles of Biological Understanding”.

7. Everything becomes nonsensical if we overlook or deny the inner power of the whole when in fact we have quietly transferred it to a part while pretending not to believe in it. Already in 1930 E. S. Russell recognized the consequences of this transfer:

The germ-plasm, even in its modern genic form, is [thought to be] something which itself remains unaltered while acting as the cause of visible change in the organism. Aristotle would have recognized in this almost mystical conception something strangely like his “soul”! (Russell 1930, pp. 267-68).

I have no wish to belittle the idea of the soul. But anyone who believes in it should believe in it — not transfer it “under the table” to particular material particles, genetic or otherwise.

More recently the philosopher of biology Jason Scott Robert remarked on the “animistic (and otherwise problematic) idea of a genetic programme” (Robert 2004, p. 37). One of the most obvious ways DNA is treated as if it were by itself an animated, living being lies in the common, yet false, conviction that it exercises not only the powers of a human programmer in order to maintain and modify a single inherited “program” for the highly divergent purposes of hundreds of cell types (and trillions of cell contexts), but also the power to reproduce itself. In reality both the replication of DNA and its adaptation to the needs of different cell types are extraordinarily complex activities of whole living cells and organisms — activities of the sort we looked at throughout the first half of this book.

Tags: evolution/and development; evolution/inheritance; evolution/and mutations; explanation/part and whole; gene/mutation; gene/genocentrism; holism

Sources

Barfield, Owen (1971). What Coleridge Thought. Middletown CT: Wesleyan University Press.

Burgess, Darren J. (2015). “Shaping Up Transcription Factor Binding”, Nature Reviews Genetics vol. 16, no. 259. doi:10.1038/nrg3944

Commoner, Barry (2002). “Unraveling the DNA Myth: The Spurious Foundation of Genetic Engineering”, Harper’s Magazine (February), pp. 39-47.

Dawkins, Richard (1995). River Out of Eden: A Darwinian View of Life. New York: BasicBooks.

Dawkins, Richard (2006). The Blind Watchmaker, third edition. New York: W. W. Norton. First edition published in 1986.

Gawne, Richard, Kenneth Z. McKenna, and H. Frederik Nijhout (2018). “Unmodern Synthesis: Developmental Hierarchies and the Origin of Phenotypes”, Bioessays vol. 40, no. 1, article 1600265. doi:10.1002/bies.201600265

Graur, Dan (2015). “Introduction to Molecular and Genome Evolution”. https://judgestarling.tumblr.com/post/129931383826/introduction-to-molecular-and-genome-evolution

Hartl, Daniel L. (1988). A Primer of Population Genetics, 2nd edition. Sunderland MA: Sinauer Associates.

Holdrege, Craig (1996). Genetics and the Manipulation of Life: The Forgotten Factor of Context. Hudson NY: Lindisfarne.

Hubbard, Ruth and Elijah Wald (1993). Exploding the Gene Myth: How Genetic Information is Produced and Manipulated by Scientists, Physicians, Employers, Insurance Companies, Educators, and Law Enforcers. Beacon Press, 1993.

Keller, Evelyn Fox (2000). The Century of the Gene. Cambridge MA: Harvard University Press.

Keller, Evelyn Fox (2013). “Genes as Difference Makers”, in Genetic Explanations: Sense and Nonsense, edited by Sheldon Krimsky and Jeremy Gruber. Cambridge MA: Harvard University Press, pp. 34-42.

Kurup, Jiji T., Ion J. Campeanu and Benjamin L. Kidder (2019). “Contribution of H3K4 Demethylase KDM5B to Nucleosome Organization in Embryonic Stem Cells Revealed by Micrococcal Nuclease Sequencing”, Epigenetics & Chromatin vol. 12, no. 20. doi:10.1186/s13072-019-0266-9

Lynch, Michael (2007a). “The Frailty of Adaptive Hypotheses for the Origins of Organismal Complexity”, PNAS vol. 104, suppl. 1 (May 15), pp. 8597-604. doi:10.1073/pnas.0702207104

Lynch, Michael (2007b). The Origins of Genome Architecture. Sunderland MA: Sinauer Associates.

Moczek, Armin P. (2012). “The Nature of Nurture and the Future of Evodevo: Toward a Theory of Developmental Evolution”, Integrative and Comparative Biology vol. 52, no. 1, pp. 108-19. doi:10.1093/icb/ics048

Moczek, Armin P. (2022). “When the End Modifies Its Means: The Origins of Novelty and the Evolution of Innovation”, Biological Journal of the Linnean Society (July 6). https://doi.org/10.1093/biolinnean/blac061

Moss, Lenny (2003). What Genes Can’t Do. Cambridge MA: MIT Press.

Moss, Lenny (2005). “Darwinism, Dualism, and Biological Agency”, in Darwinism and Philosophy, edited by Vittorio Hösle and Christian Illies. Notre Dame IN: University of Notre Dame Press, pp. 349-63.

Müller, Gerd B. (2019). “An Evolutionary Biology for the 21st Century”, in Perspectives on Evolutionary and Developmental Biology, edited by Giuseppe Fusco. Padova, Italy: Padova University Press, pp. 29-39.

Noble, Denis (2013). “Physiology Is Rocking the Foundations of Evolutionary Biology”, Experimental Physiology vol. 98.8, pp. 1235-43. doi:10.1113/expphysiol.2012.071134

Noble, Denis (2018). Dance to the Tune of Life: Biological Relativity. Cambridge UK: Cambridge University Press.

Robert, Jason Scott (2004). Embryology, Epigenesis, and Evolution: Taking Development Seriously. Cambridge UK: Cambridge University Press.

Rose, Steven (1998). Lifelines: Life Beyond the Gene. Oxford: Oxford University Press.

Russell, E. S. (1930). The Interpretation of Development and Heredity: A Study in Biological Method. Reprinted in 1972. Freeport NY: Books for Libraries Press.

Sachs, Joe (1995). Aristotle’s Physics: A Guided Study. New Brunswick NJ: Rutgers University Press.

Sinden, Richard R. (2013). “The Helix Turns at 60: Writhing Free in Chromosomes”, Nature Structural and Molecular Biology vol. 20, no. 3 (March), pp. 251-53. doi:10.1038/nsmb.2533

Walsh, Denis (2015). Organisms, Agency, and Evolution. Cambridge: Cambridge University Press.

Weiss, Paul A. (1971a). “The Basic Concept of Hierarchic Systems”, in Hierarchically Organized Systems in Theory and Practice, edited by Paul A. Weiss, pp. 1-43. New York: Hafner.

Weiss, Paul (1971b). “One Plus One Does Not Equal Two”, in Within the Gates of Science and Beyond: Science in Its Cultural Commitments, by Paul Weiss, pp. 213-61. New York: Hafner. Reprinted from The Neurosciences: A Study Program, edited by Gardner C. Quarton, Theodore Melnechuk and Francis O. Schmitt. Rockefeller University Press, 1967.

This document: https://bwo.life/bk/inherit2.htm

Steve Talbott :: Inheritance, Genetics, and the Particulate View of Life