In Defence of Physiognomy

Edward Dutton, How to Judge People by What they Look Like (Wrocław: Thomas Edward Press, 2018) 

Never judge a book by its cover’ – or so a famous proverb advises. 

However, given that Edward Dutton’s ‘How to Judge People by What they Look Like’, represents, from its provocative title onward, a spirited polemic against this received wisdom, one is tempted, in the name of irony, to review his book entirely on the basis of its cover. 

I will resist this temptation. However, it is perhaps worth pointing out that two initial points are apparent, if not from the book’s cover alone, then at least from its external appearance. These are: 

1) It is rather cheaply produced and apparently self-published; and

2) It is very short – a pamphlet rather than a book.[1]

Both these facts are probably excusable by reference to the controversial and politically-incorrect nature of the book’s title, theme and content.

Thus, on the one hand, the notion that we can, with some degree of accuracy, judge people by appearances alone is a very politically-incorrect idea and hence one that many publishers would be reluctant to associate themselves with or put their name to.

On the other hand, the fact that the topic is so controversial may also explain why the book is so short. After all, relatively little research has been conducted on this topic for precisely this reason.

Moreover, even such research as has been conducted is often difficult to track down. 

After all, physiognomy, the field of research which Dutton purports to review, is no longer a recognized science. On the contrary, most people today dismiss it as a discredited pseudoscience.

Therefore, there is no ‘International Journal of Physiognomy’ available at the click of a mouse on ScienceDirect. 

Neither are there any Departments of Physiognomy or Professors of Physiognomy at major universities, or a recent undergraduate, or graduate-level textbook on physiognomy collating all important research on the subject. Indeed, the closest thing we have to such a textbook is Dutton’s own thin, meagre pamphlet. 

Therefore, not only has relatively little research has been conducted in this area, at least in recent years, but also such research as has been conducted is spread across different fields, different journals and different researchers, and hence not always easy to track down. 

Moreover, such research rarely actually refers to itself as ‘physiognomy’, in part precisely because physiognomy is widely regarded as a pseudoscience and hence something to which researchers, even those directly researching correlations between morphology and behaviors, are reluctant to associate themselves.[2]

Therefore, conducting a key word search for the term ‘physiognomy’ in one or more of the many available databases of scientific papers would not assist the reader much, if at all, in tracking down relevant research.[3]

It is therefore not surprising that Dutton’s book is quite short. 

For this same reason, it is perhaps also excusable that Dutton has evidently failed to track down some interesting studies relevant to his theme. 

For example, a couple of interesting studies not cited by Dutton purported to uncover an association between behavioural inhibition and iris pigmentation in young children (Rosenberg & Kagan 1987; Rosenberg & Kagan 1989). 

Another interesting study not mentioned by Dutton presents data apparently showing that subjects are able to distinguish criminals from non-criminals at better than chance levels merely from looking at photographs of their faces (Valla, Ceci & Williams 2011).[4]

Such omissions are inevitable and excusable. More problematically however, Dutton also seems to have omitted at least one entire area of research relevant to his subject-matter – namely research on so-called minor physical anomalies or MPAs

These are certain physiological traits, interpreted as minor abnormalities, probably reflecting developmental instability and mutational load, which have been found in several studies to be associated with various psychiatric and developmental conditions, as well as being a correlate of criminal behaviour (see below).

Defining the Field 

Yet Dutton not only misses out on several studies relevant to the subject-matter of his book, he also is not entirely consistent in identifying just what the precise subject-matter of his book actually is. 

It is true that, at many points in his book, he talks about physiognomy

This term is usually defined as the science (or, according to many people, the pseudoscience) of using a person’s morphology in order to determine their character, personality and likely behaviour. 

However, the title of Dutton’s book, ‘How to Judge People by What They Look Like’, is potentially much broader. 

After all, what people look like includes, not just our morphology, but also, for example, how we dress and what clothes we wear.

For example, we might assess a person’s job from their uniform, or, more generally, their socioeconomic status and income level from the style and quality of their clothing, or the designer labels and brand names adorning it. 

More specifically, we might even determine their gang allegiance from the color of their bandana, and their sexuality and fetishes from the colour and positioning of their handkerchief

We also make assessments of character from clothing style. For example, a person who is sloppily dressed and is hence perceived not take care in his or her appearance (e.g. whose shirt is unironed or unclean) might be interpreted as lacking in self-worth and likely to produce similarly sloppy work in whatever job s/he is employed at. On the other hand, a person always kitted out in the latest designer fashions might be thought shallow and materialistic. 

In addition, certain styles of dress are associated with specific youth subcultures, which are often connected, not only to taste in music, but also with lifestyle (e.g. criminality, drug-use, political views).[5]

Dutton does not discuss the significance of clothing choice in assessments of character. However, consistent with this broader interpretation of his book’s title, Dutton does indeed sometimes venture beyond physiognomy in the strict sense. 

For example, he discusses tattoos (p46-8) and beards (p60-1). 

I suppose the decision to get tattooed or grow a beard reflects both genetic predispositions and environmental influence, just as all aspects of phenotype, including morphology, reflect the interaction between genes and environment. 

However, this is also true of clothing choice, which, as I have already mentioned, Dutton does not discuss.  

On the other hand, both tattoos and, given that they take time to grow, even beards are relatively more permanent than whatever clothes we are wearing at any given time. 

However, Dutton also discusses the significance of what he terms a “blank look” or “glassy eyes” (p57-9). But this is a mere facial expression, and hence even more transitory than clothing. 

Yet Dutton omits discussion of other facial expressions which, unlike his wholly anecdotal discussion of “glassy eyes”, have been researched by ethologists at least since Charles Darwin’s seminal The Expression of the Emotions in Man and Animals was published in 1872. 

Thus, Paul Ekman famously demonstrated that the meanings associated with at least some facial expressions are cross-culturally universal (e.g. smiling being associated with happiness). 

Indeed, some human facial expressions even appear to be homologues of behaviour patterns among non-human primates. For example, it has been suggested that the human smile is homologous with an appeasement gesture, namely the baring of clenched teeth (aka a ‘fear grin’), among chimpanzees. 

Of particular relevance to the question posed in Dutton’s book title, namely ‘How to Judge People by What They Look Like’, it is suggested some facial expressions lie partly outside of conscious control – e.g. blushing when embarrassed, going pale when shocked or fearful.  

Indeed, even a fake smile is said to be distinguishable from a Duchenne smile

This then explains the importance of reading facial expressions when playing poker or interrogating suspects, as people often inadvertently give away their true feelings through their facial expressions, behaviour and other mannerisms (e.g. so-called microexpressions). 

Somatotypes and Physique 

Dutton begins his book with a remarkable attempt to resurrect William Sheldon’s theory that certain types of physiques (or, as Sheldon called them, somatotypes) are associated with particular types of personality (or as Sheldon called them, constitutions). 

Although the three dimensions by which Sheldon classified physiques – endomorphy, ectomorphy and mesomorphy – have proven useful as dimensions for classifying body-type, Sheldon’s attempt to equate these ideal types with personality is now widely dismissed as pseudoscience. 

Dutton, however, argues that physique is indeed associated with character, and moreover provides what was conspicuously lacking in Sheldon’s own exposition – namely, compelling theoretical reasons for the postulated associations. 

Yet, interestingly, the associations suggested by Dutton do indeed to some extent mirror those first posited by William Shelton over half a century previously.

Whereas, elsewhere, Dutton draws on previously published research, here, Dutton’s reasoning is, to my knowledge, largely original to himself, though, as I show below, psychometric studies do support the existence of at least some of the associations he postulates. 

This part of Dutton’s book represents, in my view, the most important and convincing original contribution in the book. 

Endomorphy/Obesity, Self-Control and Conscientiousness

First, he discusses what Sheldon called endomorphy – namely, a body-type that can roughly be equated with what we would today call fatness or obesity

Dutton points out that, at least in contemporary Western societies, where there is a superabundance of food, and starvation is all but unknown even among the relatively less well-off, obesity tends to correlate with personality. 

In short, people who lack self-control and willpower will likely also lack the self-control and willpower to diet effectively. 

Endomorphy (i.e. obesity) is therefore a reliable correlate of the personality factor known to psychometricians as conscientiousness (p31-2).  

Although Dutton himself cites no data or published studies in support of this conclusion, nevertheless several published studies confirm an association between BMI and conscientiousness (Bagenjuk et al 2019; Jokela et al 2012; Sutin et al 2011). 

Obesity is also, Dutton claims, inversely correlated with intelligence

This is, first, because IQ is, according to Dutton, correlated with time-preference – i.e. a person’s willingness to defer gratification by making sacrifices in the short-term in return for a greater long-term pay-off. 

Therefore, low-IQ people, Dutton claims: 

Are less able to forego the immediate pleasure of ice cream for the future positive of not being overweight and diabetic” (p31). 

However, far from being associated with a short-time preference, some evidence, not discussed by Dutton, suggests that intelligence is actually inversely correlated with conscientiousness, such that more intelligent people are actually on average less conscientious (e.g. Rammstedt et al 2016; cf. Murray et al 2014). 

This would suggest that low IQ people might, all else being equal, actually be more successful at dieting than their high IQ counterparts. 

However, according to Dutton, there is a second reason that low-IQ people are more likely to be fat, namely: 

They are likely to understand less about healthy eating and simply possess less knowledge of what constitutes healthy food or a reasonable portion” (p31). 

This may be true. 

However, while there are some borderline cases (e.g. foods misleadingly marketed by advertisers as healthy), I suspect that virtually everyone knows that, say, eating lots of cake is unhealthy. Yet resisting the temptation to eat another slice is often easier said than done. 

I therefore suspect conscientiousness is a better predictor of weight than is intelligence

Interestingly, a few studies have investigated the association between IQ and the prevalence of obesity. However, curiously, most seem to be premised on the notion that, rather than low intelligence causing obesity, obesity somehow contributes to cognitive decline, especially in children (e.g. Martin et al 2015) and the elderly (e.g. Elias et al 2012). 

In fact, however, longitudinal studies confirm that, as contended by Dutton, it is low IQ that causes obesity rather than the other way around (Kanazawa 2014). 

At any rate, people lacking in intelligence and self-control also likely lack the intelligence and self-discipline to excel in school and gain promotions into high-income jobs, since both earnings and socioeconomic status correlate with both intelligence and conscientiousness.[6]

One can also, then, make better than chance assessments of a person’s socioeconomic status  and income from their physique. 

In other words, whereas in the past (and perhaps still in the developing world) the poor were more likely to starve or suffer from malnutrition and only the rich could afford to be fat, in the affluent west today it is the relatively less well-off who are, if anything, more likely to suffer from obesity and diseases of affluence such as diabetes and heart disease

This, then, all rather confirms the contemporary stereotype of the fat, lazy slob. 

However, Dutton also provides a let-off clause for offended fatties. Obesity is associated, not only with conscientiousness, but also with the factor of personality known as extraversion. This refers to the tendency to be outgoing, friendly and talkative, traits that are generally viewed positively. 

Several studies, again not cited by Dutton, do indeed suggest an association between extraversion and BMI (Bagenjuk et al 2019; Sutin et al 2011). Dutton, for his part, explains it this way: 

Extraverts simply enjoy everything positive more, and this includes tasty (and thus unhealthy) food” (p32). 

Dutton therefore provides theoretical support to the familiar stereotype of, not only the fat, lazy slob, but also the jolly and gregarious fat man, and the ‘bubbly’ fat woman.[7]

Mesomorphy/Muscularity and Testosterone

Mesomorphs were another of Sheldon’s supposed body-types. Mesomorphy can roughly be equated with muscularity. 

Here, Dutton concludes that: 

Sheldon’s theory… actually fits quite well with what we know about testosterone” (p33). 

Thus, mesomorphy is associated with muscularity, and muscularity with testosterone

Yet testosterone, as well as masculinizing the body, also masculinizes brain and behaviour. 

This is why anabolic steroids, not only increase muscularity, but are also said to be associated with roid rage.[8]

Testosterone, at least during development, may also be associated, not only with muscularity, but also with certain aspects of facial morphology, such as a wide and well-defined jawline, prominent brow ridges, deep-set eyes and facial width.  

I therefore wonder if this might go some way towards explain the finding, not mentioned by Dutton (but clearly relevant to his subject-matter), that observers are apparently able to identify convicted criminals at better than chance levels from a facial photograph alone (Valla, Ceci & Williams 2011).[9]

Testosterone and Autism 

Further exploring the effects of testosterone on both psychology and morphology, Dutton also proposes: 

We would also expect the more masculine-looking person to have higher levels of autism traits” (p34). 

This idea seems to be based on Simon Baron-Cohen’s extreme male brain theory of autism

However, the relationship between, on the one hand, levels of androgens such as testosterone and, on the other, degree of masculinization in respect of a given sexually-dimorphic trait may be neither one-dimensional nor linear

Thus, interestingly, Kingsley Browne in his excellent Biology at Work: Rethinking Sexual Equality (which I have reviewed here) reports: 

The relationship between spatial ability and [circulating] testosterone levels is described by an inverted U-shaped curve… Spatial ability is lowest in those with the very lowest and the very highest testosterone levels, with the optimal testosterone level lying in the lower end of the normal male range. Thus, males with testosterone in the low-normal range have the highest spatial ability” (Biology at Work: p115; Gouchie & Kimura 1991). 

Similarly, leading intelligence researcher Arthur Jensen reports, in The g Factor: The Science of Mental Ability, that:

Within each sex there is a nonlinear (inverted-U) relationship between an individual’s position on the estrogen/testosterone continuum and the individual’s level of spatial ability, with the optimal level of testosterone above the female mean and below the male mean. Generally, females with markedly above-average testosterone levels (for females) and males with below-average levels of testosterone (for males) tend to have higher levels of spatial ability, relative to the average spatial ability for their own sex” (The g Factor: p534).

In contrast, however, Dutton claims: 

There is evidence that testosterone level in healthy males is positively associated with spatial ability” (p36). 

However, the only study he cites in support of this assertion was, according to its methodology section and indeed its very title, conducted among “older males”, reported as having been between the ages of 60 and 75 years of age (Janowsky et al 1994). 

Therefore, since testosterone levels are known to decline with age, this finding is not necessarily inconsistent with the relationship between testosterone and spatial ability described by Browne (see Moffat & Hampson 1996). 

This, of course, accords with the anecdotal observation that math nerds and autistic males are rarely athletic, square-jawed ‘alpha male’-types.[10]

Testosterone and Baldness 

Another trait associated with testosterone levels, according to Dutton, is male pattern baldness. Thus, Dutton contends: 

Baldness is yet another reflection of high testosterone… [B]aldness in males known as androgenic apolecia, is positively associated with levels of testosterone” (p55). 

As evidence, he cites a study both a review (Batrinos 2014) and some indirect anecdotal evidence: 

It is widely known among doctors – I base this on my own discussions with doctors – that males who come to them in their 60s complaining of impotence tend to have full heads of fair or only very limited hair loss” (p55).[11]

If male pattern baldness is indeed associated with testosterone levels then this is somewhat surprising, because our perceptions regarding men suffering from male pattern baldness seem to be that they are, if anything, less masculine than other males. 

Thus, Nancy Etcoff, in Survival of the Prettiest (which I have reviewed here), reports that one study  found that: 

Both sexes assumed that balding men were weaker and found them less attractive” (Survival of the Prettiest: p121; Cash 1990).[12]

Yet, if the main message of Dutton’s book is that individual differences in morphology and appearance do indeed predict individual differences in behaviour, psychology and personality, then a second implicit theme seems also to be that our intuitions and stereotypes regarding the association between appearance and behaviors are often correct.  

True, it is likely that few people notice, say, digit ratios, or make judgements about people based on them either consciously or unconsciously. However, elsewhere, Dutton cites studies showing that subjects are able to estimate the IQ of male students at better than chance levels simply by viewing a photograph of their faces (Kleisner et al 2014; discussed at p50); and identify homosexuals and heterosexual men at better than chance levels from a facial photograph alone (Kosinski & Wang 2017; discussed at p66). 

Yet, according to Etcoff and Cash, perceptions regarding the personalities of balding men are almost the opposite of what would be expected if male pattern balding were indeed a reflection of high testosterone levels, as suggested by Dutton. 

In fact, however, although a certain level of testosterone is indeed a necessary condition for male pattern hair loss (this is why neither women nor castrated eunuchs experience the condition, though their hair does thin with age), this seems to be a threshold effect, and among non-castrated males with testosterone levels within the normal range levels of circulating testosterone do not seem to significantly predict either the occurrence, or severity, of male pattern baldness

Thus, healthline reports: 

It’s not the amount of testosterone or DHT that causes baldness; it’s the sensitivity of your hair follicles. That sensitivity is determined by genetics. The AR gene makes the receptor on hair follicles that interact with testosterone and DHT. If your receptors are particularly sensitive, they are more easily triggered by even small amounts of DHT, and hair loss occurs more easily as a result. 

In other words, male pattern baldness is yet another trait that is indeed related to testosterone, but does not evince a simple linear relationship

2D:4D Ratio

Another presumed correlate of prenatal androgens is 2D:4D ratio (aka digit ratio). 

Over the last two decades, a huge body of research has reported correlations between 2D:4D ratio and a variety of psychiatric conditions and behavioural propensities, including autism (Manning et al 2001), ADHD (Martel et al 2008; Buru 2020; Işık 2020), psychopathy (Blanchard & Lyons 2010), aggressive behaviours (Bailey & Hurd 2005; Benderlioglu & Nelson 2005), sports and athletic performance (Manning & Taylor 2001Hönekopp & Urban 2010; Griffin et al 2012; Keshavarz et al 2017), criminal behaviour (Ellis & Hoskin 2015; Hoskin & Ellis 2014) and homosexuality (Williams et al 2000; Lippa 2003; Kangassalo et al 2011; Li et al 2016; Xu & Zheng 2016). 
 
Unfortunately, and slightly embarrassingly, Dutton apparently misunderstands what 2D:4D ratio actually measures. Thus, he writes: 

If the profile of someone’s fingers is smoother, more like a shovel, then it implies high testosterone. If, by contrast, the little finger is significantly smaller than the middle finger, which is highly prevalent among women, then it implies lower testosterone exposure” (p69). 

Actually, however, both the little finger and middle finger are irrelevant to 2D:4D ratio.

Indeed, for virtually everyone, “the little finger is significantly smaller than the middle finger”. This is, of course, why the latter is called “the little finger”.

Actually, 2D:4D ratio concerns the ratio between index finger and the ring finger – i.e. the two fingers on either side of the middle finger

These fingers are, of course, the second and fourth digit, respectively, if you begin counting from your thumb outwards, hence the name ‘2D:4D ratio’. 

In evidently misnumbering his digits, I can only conclude that Dutton began counting at the correct end, but missed out his thumb. 

At any rate, the evidence for any association between digit ratios and measures of behavior and psychology is, at best, mixed

Skimming the literature on the subject, one finds many conflicting findings – for example, sometimes significant effects are found only for one sex, while other studies find the same correlations limited to the other sex (e.g. Bailey & Hurd 2005; Benderlioglu & Nelson 2005; see also Hilgard et al 2019), and also many failures to replicate earlier reported associations (e.g. Voracek et al 2011; Fossen et al 2022; Kyselicová et al 2021). 

Likewise, meta-analyses of published studies have generally found, at best, only small and inconsistent associations (e.g Voracek et al 2011 ; Pratt et al 2016). Thus, 2D:4D ratio has been a major victim of the recent so-called replication crisis in psychology

Indeed, it is not entirely clear that 2D:4D ratio represents a useful measure of prenatal androgens in the first place (Hollier et al 2015), and even the universality of the sex difference that originally led researchers to posit such a link is has been called into question (Apicella 2015; Lolli et al 2017).  

In short, the usefulness of digit ratio as a measure of exposure to prenatal androgens, let alone an important correlate of behaviour, psychology, personality or athletic performance, is questionable. 

Testosterone and Height 

The examples of male pattern baldness and spatial ability demonstrate that the effect of testosterone on some sexually-dimorphic traits is not necessarily always linear. Instead, it can be quite complex. 

Therefore, just because men are, on average, higher for a given trait than are women, which is ultimately a consequence of androgens such as testosterone, this does not necessarily mean that men with relatively higher levels of testosterone are necessarily higher for this trait than are men with relatively lower levels of testosterone. 

Indeed, Dutton himself provides another example of such a trait – namely height

Thus, although men, in general, are taller than women, nevertheless, according to Dutton: 

Men who are high in testosterone… tend to be of shorter stature than those who are low in it. High levels of testosterone at a relatively early age have been shown to reduce stature” (p34).[13]

In evolutionary terms, Dutton explains this in terms of the controversial Life History Theory of Philippe Rushton, of whom Dutton seems to be, with some reservations, something of a disciple (p22-4). 

If true, this might explain why eunuchs who were castrated before entering puberty are said to grow taller, on average, than other men. 

Further corroboration is provided by the fact that, in the Netherlands, whose population is among the tallest in the world, excessively tall boys are sometimes treated with testosterone in order to prevent them growing any taller (de Waal et al 1995).[14]

This is said to occur because additional testosterone speeds up puberty, and produces a growth spurt, but it also brings this to an end when height stabilizes and we cease to grow any taller. This is discussed in Carole Hooven’s book Testosterone: The Story of the Hormone that Dominates and Divides Us.

Short Man Syndrome’?

Interestingly, although Dutton does not explore the idea, the association between testosterone levels and height among males may even explain the supposed phenomenon of short man syndrome (also referred to, by reference to the supposed diminutive stature of the French emperor Napoleon, as a Napoleon complex), whereby short men are said to be especially aggressive and domineering. 

This is something that is usually attributed to a psychological need among shorter men to compensate for their diminutive stature. However, if Dutton is right, then the supposed aggressive predilections of short men might simply reflect differences between short and taller man in testosterone levels during adolescence. 

Actually, however, so-called short man syndrome is likely a myth – and yet another way society in general demeans and belittles short men. Certainly, it is very much a folk-psychiatric diagnosis with no empirical or real evidential basis, besides the merely anecdotal.  

Indeed, far from short men being, on average, more aggressive and domineering than taller men, one study commissioned by the BBC actually found that short men were less likely to respond aggressively when provoked

Given that tall men have an advantage in combat, it would actually make sense for relatively shorter men to avoid potentially violent confrontations with other men where possible, since, all else being equal, they would be more likely to come off worse in any such altercation.  

Consistent with this, some studies have found a link between increased stature and anti-social personality disorder, which is associated with aggressive behaviours (e.g. Ishikawa et al 2001; Salas-Wright & Vaughn 2016), while another study found a positive association between height and dominance, especially among males (Malamed 1992).[15]

Height and Intelligence 

Height is also, Dutton reports, correlated with intelligence, with taller people having, on average, slightly higher IQs than shorter people.  

The association between height and IQ is, like most if not all of those discussed by Dutton in this book, modest in magnitude or effect size.[16]

However, unlike many other associations reported by Dutton, many of which are based on just a single published study, or sometimes by purely theoretical arguments, the association between height and intelligence is robust and well-established.[17] Indeed, there is even wikipedia page on the topic

Dutton’s explanation for this phenomenon is that intelligence and height “have been sexually selected for as a kind of bundle” (p46). 

Females have sexually selected for intelligent men (because intelligence predicts social status and they have been specifically selected for this) but they have also selected for taller men, realising that taller men will be better able to protect them. This predilection for tall but intelligent men has led to the two characteristics being associated with one another” (p46). 

Actually, as I see it, this explanation would only work, or at least work much better, if both men and women had a preference for partners who are both tall and intelligent

This is indeed Arthur Jensen’s explanation for the association between height and IQ

Probably represents a simple genetic correlation resulting from cross-assortative mating for the two traits. Both height and ‘intelligence’ are highly valued in western culture. There is also evidence for cross-assortative mating for height and IQ. There is some trade-off between them in mate selection. When short and tall women are matched on IQ, educational level and social class of origin, for example, it is found that taller women tend to marry men of higher socioeconomic status… than do shorter women” (The G Factor: The Science of Mental Ability: p146). 

An alternative explanation might be that both height and intelligence reflect developmental stability and a lack of deleterious mutations. On this view, both height and intelligence might represent indices of genetic quality and lack of mutational load. 

However, this alternative explanation is inconsistent with the finding that there is no ‘within-family’ correlation between height and intelligence. In other words, when one looks at, say, full-siblings from the same family, there is no tendency for the taller sibling to have a higher IQ (Mackintosh, IQ and Human Intelligence: p6). 

This suggests that the genes that cause greater height are different from those that cause greater intelligence, but that they have come to be found in the same individuals through assortative mating, as suggested by Jensen and Dutton.[18]

Height and Earnings 

Although not discussed by Dutton, there is also a correlation between height and earnings. Thus, economist Steven Landsburg reports that: 

In general, an extra inch of height adds roughly an extra $1,000 a year in wages, after controlling for education and experience. That makes height as important as race or gender as a determinant of wages” (More Sex is Safer Sex: p53). 

This correlation could be mediated by the association between height and intelligence, since intelligence is known to be correlated with earnings (Case & Paxson 2009). 

However, one interesting study found that it was actually height during adolescence that accounted for the association, and that, once this was controlled for, adult height had little or no effect on earnings (Persico, Postlewaite & Silverman 2004). 

Controlling for teen height essentially eliminates the effect of adult height on wages for white males. The teen height premium is not explained by differences in resources or endowments” (Persico, Postlewaite & Silverman 2004). 

Thus, Landsburg reports: 

Tall men who were short in high school earn like short men, while short men who were tall (for their age) in high school” (More Sex is Safer Sex: p54). 

This suggests that it is height during a key formative period (a critical period’) in adolescence that increases self-confidence, which self-confidence continues into adulthood and ultimately contributes to higher adult earnings of men who were relatively taller as adolescents. 

On the other hand, however, Case and Paxon report that, in addition to being associated with adult height, intelligence is also associated with an earlier growth spurt. This leads them to conclude that adolescent height might be a better marker for cognitive ability than adult height, thereby providing an alternative explanation for Persico et al’s finding (Case & Paxson 2009). 

Head Size and Intelligence 

Dutton also discusses the finding that there is an association between intelligence and head-size. This is indeed true and is a topic I have written about elsewhere

However, Dutton’s illustration of this phenomenon seems to me rather unhelpful. Thus, he writes: 

Intelligent people have big heads in comparison to the size of their bodies. This association is obvious at the extremes. People who suffer from a variety of conditions that reduce their intelligence, including fetal alcohol syndrome or the zika virus, have noticeably very small heads” (p56). 

However, to me, this seems to be the wrong way to think about it. 

While it is indeed true that microcephaly (i.e. a smaller than usual head size) is usually associated with lower than normal intelligence levels, the reverse is not true. Thus, although head-size is indeed correlated with IQ, people suffering from macrocephaly (i.e. abnormally large heads) do not generally have exceptionally high IQs. On the contrary, macrocephaly is often associated with impaired cognitive function, probably because, like microcephaly, it reflects a malfunction in brain development.

Neither do people afflicted with forms of disproportionate dwarfism, such as achondroplasia, have higher than average IQs even though their heads are larger relative to their body-size than are those of ordinary-sized people.  

In short, rather than being, as Dutton puts it “obvious at the extremes”, the association between head-size and intelligence is obvious at only one of the extremes and not at all apparent at the other extreme. 

In general, species, individuals and races with larger brains have higher intelligence because, because brain-size is highly metabolically expensive and therefore unlikely to evolve without some compensating advantage (i.e. higher intelligence). 

However, conditions such achondroplasia and macrocephaly did not evolve through positive selection. On the contrary, they are pathological and maladaptive. Therefore, in these cases, the additional brain tissue may indeed be wasted and hence confer no cognitive advantage. 

Mate Choice 

In evolutionary psychology, there is a large literature on human mate-choice and beauty/attractiveness standards. Much of this depends on the assumption that the physical characteristics favoured as mate-choice criteria represent fitness-indicators, or otherwise correlate with traits desirable in a mate. 

For example, a low waist-to-hip ratio (or ‘WHR’) is said to be perceived as attractive among females because it is supposedly a correlate of both health and fertility. Similarly, low levels of fluctuating asymmetry are thought to be perceived as attractive by members of the opposite sex in both humans and other animals, supposedly because it is indicative of developmental stability and hence indirectly of genetic quality

Dutton reviews some of this literature. However, an introductory textbook on evolutionary psychology (e.g. David Buss’s Evolutionary Psychology: The New Science of the Mind), or on the evolutionary psychology of mating behaviour in particular (e.g. David Buss’s The Evolution of Desire), would provide a more comprehensive review. 

Also, some of Dutton’s speculations are rather unconvincing. He claims: 

Hipsters with their Old Testament beards are showcasing their genetic quality… Beards are a clear advertisement of male health and status. They are a breeding ground for parasites” (p61). 

However, if this is so, then it merely raises the question as to why have beards come back into fashion very recently? Indeed, until the last few years, beards had not been in fashion for men in the west to my knowledge since the 1970s.[19]

Moreover, it is not at all clear that beards do increase attractiveness (e.g. Dixson & Vasey 2012). Rather, it seems that beards increase perceptions of male age, dominance, social status and aggressiveness, but not their attractiveness.[20]

This suggests that beards are more likely to have evolved through intrasexual selection (i.e. dominance competition or fighting between males) than by intersexual selection (i.e. female choice). 

This is actually consistent with a recently-emerging consensus among evolutionary psychologists that human male physiology (and behaviour) has been shaped more by intrasexual selection than by intersexual selection (Puts 2010; Kordsmeyer et al 2018). 

Consistent with this, Dutton notes: 

“[Beards] have been found to make men look more aggressive, of higher status, and older… in a context in which females tend to be attracted to slightly older men, with age tending to be associated with status in men” (p61). 

However, this raises the question as to why, today, most men prefer to look younger.[21]

Are Feminine Faces More Prone to Infidelity?

Another interesting idea discussed by Dutton is that mate-choice criteria may vary depending on the sort of relationship sought. For example, he suggests: 

A highly feminine face is attractive, in particular in terms of a short term relationship… [where] a healthy and fertile partner is all that is needed” (p43). 

In contrast, however, he concludes that for a long-term relationship a less feminine face may be desirable, since he contends “being extremely feminine in terms of secondary sexual characteristics is associated with an r-strategy” and hence supposedly with a greater risk of infidelity (p43).[22]

However, Dutton presents no evidence in favour of the claim that less feminine women are less prone to sexual infidelity. 

Actually, on theoretical grounds, I would contend that the precise opposite relationship is more likely to exist. 

After all, less feminine and more masculine females, having been subjected to higher levels of androgens, would presumably also have a more male-typical sexuality, including a high sex drive and preference for promiscuous sex with multiple partners

Indeed, there is data in support of this conclusion, from studies of women afflicted with a rare condition, congenital adrenal hyperplasia, which results in their having been exposed to abnormally high levels of masculinizing androgens such as testosterone both in the womb and sometimes in later life as compared to other females, and who, as a consequence, exhibit a more male-typical psychology and sexuality than other females. 

Thus, Donald Symons in his seminal The Evolution of Human Sexuality (which I have reviewed here) reports:  

There is evidence that certain aspects of adult male sexuality result from the effects of prenatal and postpubertal androgens: before the discovery of cortisone therapy women with andrenogenital syndrome [AGS] were exposed to abnormally high levels of androgens throughout their lives, and clinical data on late-treated AGS women indicate clear-cut tendencies toward a male pattern of sexuality” (The Evolution of Human Sexuality: p290). 

Thus, citing the work of, among others the much-demonized John Money, Symons reports that women suffering from andrenogenital syndrome

Tended to exhibit clitoral hypersensitivity and an autonomous, initiatory, appetitive sexuality which investigators have characterized as evidencing a high sex drive or libido” (The Evolution of Human Sexuality: p290). 

This suggests that females with a relatively more masculine appearance, having been subject, on average, to higher levels of masculinizing androgens, will also evidence a more male-typical sexuality, including greater promiscuity and hence presumably a greater proclivity towards infidelity, rather than a lesser tendency as theorized by Dutton. 

Good Looks, Politics and Religion 

Dutton also cites studies showing that conservative politicians, and voters, are more attractive than liberals (Peterson & Palmer 2017; Berggren et al 2017). 

By way of explanation for these findings, Dutton speculates that in ancestral environments: 

Populations… so low in ethnocentrism as to espouse Multiculturalism and reject religion would simply have died out… Therefore… the espousal of leftist dogmas would partly reflect mutant genes, just as the espousal of atheism does. This elevated mutational load… would be reflected in their bodies as well as their brains” (p76). 

However, this seems unlikely, since atheism and possibly socially liberal political views as well have usually been associated with higher intelligence, which is probably a marker for good genes.[23]

Moreover, although mutations might result in suboptimal levels of both ethnocentrism and religiosity, these suboptimal levels would presumably also manifest in the form of excessive levels of religiosity and ethnocentrism

This would suggest that religious fundamentalists and extreme xenophobes and racial supremacists would be just as mutated, and hence just as ugly, as atheists and extreme leftists supposedly are. 

Yet Dutton instead insists that religious fundamentalists, especially Mormons, tend to be highly attractive (Dutton et al 2017). However, he and his co-authors cite little evidence for this claim beyond the merely anecdotal.[24]

The authors of the original paper, Dutton reports, themselves suggested an alternative explanation for the greater attractiveness of conservative politicians, namely: 

Beautiful people earn more, which makes them less inclined to support redistribution” (p75). 

This, to me seems, both simpler more plausible. However, in response, Dutton observes: 

There is far more to being… right-wing… than not supporting redistribution” (p75). 

Here, he is right. The correlation between socioeconomic status/income and political ideology and voting is actually quite modest (see What’s Your Bias). 

However, earnings do still correlate with voting patterns, and this correlation is perhaps enough to explain the modest association between physical attractiveness and political opinions. 

Nevertheless, other factors may also play a role. For example, a couple of studies have found, among men, an association between grip strength and support for policies that benefit oneself economically (Peterson et al 2013; Peterson & Laustsen 2018). 

Grip strength is associated with muscularity, which is generally considered attractive in males

Since most leading politicians mostly come from middle-class, well-to-do, if not elite backgrounds, this would suggest that conservative male politicians are likely to be, on average, more attractive than liberal or leftist politicians.

Indeed, Noah Carl has even purported to observe, and presents evidence suggesting, a general, and widening, masculinity gap between the political left and right, and some studies have found evidence that more physically formidable males have more conservative and less egalitarian political views (Price et al 2017; Kerry & Murray 2018). 

Since masculinity in general (e.g. not just muscularity, but also square jaws etc.) is associated with attractiveness in males (see discussion here), this might explain at least part of the association between political views and physical attractiveness. 

On the other hand, among females, an opposite process may be at work. 

Among women, leftist politics seem to be strongly associated with feminist views

Since feminists reject traditional female sex roles, it is likely they would be relatively less ‘feminine’ than other women, perhaps having been, on average, subjected to relatively higher levels of androgens in the womb, masculinizing both their behaviour and appearance. 

Yet it is relatively more feminine women, with feminine, sexually-dimorphic traits such as large breasts, low waist to hip ratios, and neotenous facial features, who are perceived by men as more attractive.

It is therefore unsurprising that feminist women in particular tend to be less attractive than women who are attracted to traditional sex roles.[25]

Developmental Disorders and MPAs

One study cited by Dutton found that observers are able to estimate a male’s IQ from a facial photograph alone at better than chance level (Kleisner 2014). To explain this, Dutton speculates: 

Having a small nose is associated with Downs [sic] Syndrome and Foetal Alcohol Syndrome and this would have contributed to our assuming that those with smaller noses were less intelligent” (p51). 

Thus, he explains: 

“[Whereas] Downs [sic] Syndrome and Foetal Alcohol Syndrome are major disruptions of developmental pathways and they lead to very low intelligence and a very small nose… even minor disruptions would lead to slightly reduced intelligence and a slightly smaller nose” (p51-2). 

Indeed, foetal alcohol syndrome itself seems to exist on a continuum and is hence a matter of degree. 
 
Indeed, going further than Dutton, I would agree with publisher/blogger Chip Smith, who observes in his blog

Dutton only mention[s] trisomy 21 (Down syndrome) in passing, but I think that’s a pretty solid place to start if you want to establish the baseline premise that at least some mental traits can be accurately inferred from external appearances.” 

Thus, the specific ‘look associated with Down Syndrome is a useful counterexample to cite to anyone who dismisses the idea of physiognomy, and the existence of any association between looks and ability or behaviour, a priori

Indeed, other developmental disorders and chromosomal abnormalities, not mentioned by Dutton, are also associated with a specific specific ‘look’ – for example, Williams Syndrome, the distinctive appearance, and personality, associated with which has even been posited as the basis for the elf figure in folklore.[26]

Less obviously, it has even been suggested that there are also subtle facial features that distinguish autistic children from neurotypical children, and which also distinguish boys with relatively more severe forms of autism from those who are likely to be diagnosed as higher functioning (Aldridge et al 2011; Ozgen et al 2011). 

However, Dutton neglects to mention that there is in fact a sizable literature regarding the association between so-called minor physical anomalies (aka MPAs) and several psychiatric conditions including autism (Ozgen et al 2008), schizophrenia (Weinberg et al 2007; Xu et al 2011) and paedophilia (Dyshniku et al 2015). 

MPAs have also been identified in several studies as a correlate of criminal behaviour (Kandel et al 1989; see also Criminology: A Global Perspective: p70-1). 

Yet these MPAs are often the very same traits – the single transverse palmar crease; sandal toe gap; fissured tongue – that are also used to diagnose Down Syndrome in nenates.

The Morality of Making Judgements

But is it not superficial to judge a book by its cover? And, likewise, by extension, isn’t it morally wrong to judge people by their appearance? 

Indeed, it is not only morally wrong to judge people by their appearance, but also, worse still, isn’t it racist

After all, skin colour is obviously a part of our appearance, and did not our Lord and Saviour, Dr Martin Luther King, himself advocate for a world in which people would be judged “not be judged by the color of their skin but by the content of their character.” 

Here, Dutton turns from science to morality, and convincingly contends that, at least in certain circumstances, it is indeed morally acceptable to judge people by appearances. 

It is true, he acknowledges, that most of the correlations that he has uncovered or reported are modest in magnitude. However, he is at pains to emphasize, the same is true of almost all correlations that are found throughout psychology and the social sciences. Thus, he exhorts: 

Let us be consistent. It is very common in psychology to find a correlation between, for example, a certain behaviour and accidents (or health) of 0.15 or 0.2 and thus argue that action should be taken based on the results. These sizes are considered large enough to be meaningful and even for policy to be changed” (p82). 

However, Dutton also includes a few sensible precautions and caveats to be borne in mind by those readers who might be tempted overenthusiastically apply some of his ideas. 

First, he warns against regarding making inferences regarding “people from a racial group with which you have relatively limited contact”, where the same cues used with respect to your own group may be inapplicable, or must be applied relative to the group averages for the other group, something we may not be adept at doing (p82-3). 

Thus, to give an obvious example, among Caucasians, epicanthic folds (i.e. so-called ‘slanted’ eyes) may be indicative of a developmental disorder such as Down syndrome. However, among East Asians, Southeast Asians and some other racial groups (notably the Khoisan of Southern Africa), such folds are entirely normal and not indicative of any pathology. 

He also cautions regarding people’s ability to disguise their appearance, both by makeup and plastic surgery. However, also notes that the tendency to wear excessive makeup, or undergo cosmetic surgery, is itself indicative of a certain personality type, and indeed often, Dutton asserts, of psychopathology (p84-5). 

Using physical appearance to make assessments is particularly useful, Dutton observes, “in extreme situations when a quick decision must be made” (p80). 

Thus, to take a deliberately extreme reductio ad absurdum, if we see someone stabbing another person, and this first person then approaches us in an aggressive manner brandishing the knife, then, if we take evasive action, we are, strictly speaking, judging by appearances. The person appears as if they are going to stab us, so we assume they are and act accordingly. However, no one would judge us morally wrong for so doing. 

However, in circumstances where we have access to greater individualizing information, the importance of appearances becomes correspondingly smaller. Here, a Bayesian approach is useful. 

In 2013, evolutionary psychologist Geoffrey Miller caused predictable outrage and hysteria when he tweeted

Dear obese PhD applicants: if you didn’t have the willpower to stop eating carbs, you won’t have the willpower to do a dissertation #truth.” 

According to Dutton, as we have seen above, willpower is indeed likely correlated with obesity, because, as Miller argues, people lacking in willpower also likely lack the willpower to diet. 

However, a PhD supervisor surely has access to far more reliable information regarding a person’s personality and intelligence, including their conscientiousness and willpower, in the form of their application and CV, than is obtainable from their physique alone. 

Thus, the outrage that this tweet provoked, though indeed excessive and a reflection of the intolerant climate of so-called cancel culture’ and public shaming in the contemporary west, was not entirely unwarranted. 

Similarly, if geneticist James Watson did indeed say, as he was rather hilariously reported as having said, that “Whenever you interview fat people, you feel bad, because you know you’re not going to hire them”, he was indeed being prejudiced, because, again, an employer has access to more reliable information regarding applicants than their physique, namely, again, their application and CV. 

Obesity may often—perhaps even usually—be indicative of low levels of conscientiousness, willpower and intelligence. But, it is not always indicative of low levels of conscientiousness, willpower and intelligence. Instead, it may instead, as Dutton himself points out, reflect only high extraversion, or indeed an unusual medical condition. 

However, even at job interviews, employers do still, in practice, judge people partly by their appearance. Moreover, we often regard them as well within their rights to do so. 

This is, of course, why we advise applicants to dress smartly for their interviews.

Endnotes

[1] If ‘How to Judge People by What They Look Like’ is indeed a very short book, then, it must be conceded that this is, by comparison, a rather long and detailed book review. While, as will become clear in the remainder of this review, I have many points of disagreement with Dutton (as well as many points of agreement) and there are many areas where I feel he is mistaken, nevertheless the length of this book review is, in itself, testament to the amount of thinking that Dutton’s short pamphlet has inspired in this reader. 

[2] In addition, I suspect few of the researchers whose work Dutton cites ever even regarded themselves as working within, or somehow reviving, the field of physiognomy. On the contrary, despite researching and indeed demonstrating robust associations between morphology and behavior, this idea may never even have occurred to them.
Thus, for example, I was already familiar with some of this literature even before reading Dutton’s book, but it never occurred to me that what I was reading was a burgeoning literature in a revived science of physiognomy. Indeed, despite being familiar with much of this literature, I suspect that, if questioned directly on the matter, I may well have agreed with the general consensus that physiognomy was a discredited pseudoscience.
Thus, one of the chief accomplishments of Dutton’s book is simply to establish that this body of research does indeed represent a revived science of physiognomy, and should be recognized and described as such, even if the researchers themselves rarely if ever use the term.

[3] Instead, it would surely uncover mostly papers in the field of ‘history of science’, documenting the history of physiognomy as a supposedly discredited pseudoscience, along with such other real and supposed pseudosciences as phrenology and eugenics.

[4] The studies mentioned in the two paragraphs that precede this endnote are simply a few that I happen to have stumbled across that are relevant to Dutton’s theme and which I happen to have been able to recall. No doubt, any list of relevant studies that I could compile would be just as inexhaustive as that of Dutton and my own list would be longer than Dutton’s only because I have the advantage of having read Dutton’s book beforehand.

[5] Thus, a young person dressed as a hippy in the 60s and 70s was more likely to ascribe to certain (usually rather silly and half-baked) political beliefs, and also more likely to engage in recreational drug-use and live on a commune, while a young man dressed as a teddy boy in Britain in the 1950s, a skinhead in the 1970s and 80s, a football casual in the 1990s, or indeed a chav today, may be perceived as more likely to be involved in violent crime and thuggery. The goth subculture also seems to be associated with a certain personality type, and also with self-harm and suicide.

[6] The association between IQ and socioeconomic status is reviewed in The Bell Curve: Intelligence and Class Structure in American Life (which I have reviewed here). The association between conscientiousness and socioeconomic status is weaker, probably because personality tests are a less reliable measure of conscientiousness than IQ tests are of IQ, since the former rely on self-report. This is the equivalent of an IQ test that, instead of asking test-takers to solve logical puzzles, simply asked them how good they perceived themselves to be at solving logical puzzles. Nevertheless, conscientiousness, as measured in personality tests, does indeed correlate with earnings and career advancement, albeit less strongly than does IQ (Spurk & Abele 2011Wiersma & Kappe 2016).

[7] If some fat people are low in conscientiousness and intelligence, and others merely high in extraversion, there may, I suspect, also be a third category of people who do have self-control and self-discipline, but simply do not much care about whether they are fat or thin. However, given both the social stigma and health implications of obesity, this group is, I suspect, small. It is also likely young, since health dangers of obesity increase with age, and male, since both the social stigma of fatness, and especially its negative impact on mate value and attractiveness, seems to be greater for females. 

[8] Actually, whether roid rage is a real thing is a matter of some dispute. Although users of anabolic steroids do indeed have higher rates of violent crime, it has been suggested that this may be at least in part because the type of people who choose to use steroids are precisely those already prone to violence. In other words, there is a problem of self-selection bias.
Moreover, the association between testosterone and aggressive behaviours is more complex than this simple analysis assumes. One leading researcher in the field, Allan Mazur, argues that testosterone is not associated with aggression or violence per se, but only with dominance behaviours, which only sometimes manifest themselves through violent aggression. Thus, for example, a leading politician, business tycoon or chief executive of a large company may have high testosterone and be able to exercise dominance without resort to violence. However, a prisoner, being of low status in the legitimate world, is likely only able to assert dominance through violence (see Mazur & Booth 1998; Mazur 2009).

[9] Here, however, it is important to distinguish between the so-called organizing and ‘activating’ effects of testosterone. The latter can be equated with levels of circulating testosterone at any given time. The former, however, involves androgen levels at certain key points during development, especially in utero (i.e. in the womb) and during puberty, which thenceforth have long-term effects on both morphology and behaviour (and a person’s degree of susceptibility to circulating androgens).
Facial bone structure is presumable largely an effect of the ‘organizing’ effects of testosterone during development, though jaw shape is also affected by the size of the jaw muscles, which can be increased, it has been claimed, by regularly chewing gum. Bodily muscularity, on the other hand, is affected by both levels of circulating testosterone (hence the effects of anabolic steroids on muscle growth) but also levels of testosterone during development, not least because high levels of androgens during development increases the number and sensitivity of androgen receptors, which affect the potential for muscular growth.

[10] In this section, I have somewhat conflated spatial ability, mathematical ability and autism traits. However, these are themselves, of course, not the same, though each is probably associated with the others, albeit again not necessarily in a linear relationship.

[11] I have been unable to discover any evidence for this supposed association between lack of balding and impotence in men. On the contrary, googling the terms ‘male pattern baldness’ and ‘impotence’ finds only a results, mostly people speculating whether there is a positive correlation between balding and impotence in males, if only on the very unpersuasive ground that the two conditions tend to have a similar age of onset (i.e. around middle-age).

[12] In contrast, the shaven-head skinhead-look, or close-cropped military-style induction cut, buzz cut or high and tight is, of course, perceived as a quintessentially masculine, and even thuggish, hairstyle. This is perhaps because, in addition to contrasting with the long hair typically favoured by females, it also, by reducing the size of the upper part of the head, makes the lower part of the face e.g. the jaw and body, appear comparatively larger, and large jaws are a masculine trait, Thus, Nancy Etcoff observes:

The absense of hair on the head serves to exaggerate signals of strength. The smaller the head the bigger the look of the neck and body. Bodybuilders often shave or crop their hair, the size contrast between the head and neck and shoulders emphasizing the massiveness of the chest” (Survival of the Prettiest: p126).

[13] The source that Dutton cites for this claim is (Nieschlag & Behr 2013).

[14] In America, it has been suggested, especially tall boys are not treated with testosterone to prevent their growing any taller. Instead, they are encouraged to attempt to make a successful career in professional basketball

[15] On the other hand, one Swedish study investigating the association between height and violent crime found that the shortest men in Sweden had almost double convictions for violent crimes as compared to the tallest men in Sweden. However, after controlling for potential confounds (e.g. socioeconomic status and intelligence, both of which positively correlate with height), the association was reversed, with taller man having a somewhat higher likelihood of being convicted of a violent crime (Beckley et al 2014). 

[16] According to Dutton, the correlation between height and IQ is only about r = 0.1. This is a modest correlation even by psychology and social science standards.

[17] In other words, although modest in magnitude, the association between height and IQ has been replicated in so many studies with sufficiently large and representative sample sizes that we can be certain that it represents a real association in the population at large, not an artifact of small, unrepresentative or biased sampling in just one or a few studies. 

[18] An alternative explanation for the absence of a within-family correlation between height and intelligence is that some factor that differs as between families causes both increased height and increased intelligence. An obvious candidate would be malnutrition. However, in modern western economies where there is an superabundance of food, starvation is almost unknown and obesity is far more common than undernourishment even among the ostensible poor (indeed, as noted by Dutton, especially among the ostensible poor), it is doubtful that undernourishment is a significant factor in explaining either small stature or low IQs, especially since height is mostly heritable, at least by the time a person reaches adulthood.

[19] The conventional wisdom is that beards went out of fashion during the twentieth century precisely because their role as in spreading germs came to be more widely known. Thus, Nancy Etcoffwrites:

Facial hair has been less abundant in this century than in centuries past (except in the 1960s) partly because medical opinion turned against them. As people became increasingly aware of the role of germs in spreading diseases, beards came to be seen as repositories of germs. Previously, they had been advised by doctors as a means to protect the throat and filter air to the lungs” (Survival of the Prettiest: p156-7). 

Of course, this is not at all inconsistent with the notion that beards are perceived as attractive by women precisely because they represent a potential vector of infection and hence advertise the health and robustness of the male whom they adorn, as contended by Dutton. On the contrary, the fact that beards are indeed associated with infection, is consistent with and supportive of Dutton’s theory. 

[20] It would be interesting to discover whether these findings generalize to other, non-western cultures, especially those where beards are universal or the norm (e.g. among Muslims in the Middle East). It would also be discover whether women’s perceptions regarding the attractiveness of men with beards have changed as beards have gone in and out of fashion. 

[21] Perhaps this is because, although age is still associated with status, it is no longer as socially acceptable for older men to marry, or enter sexual relationships with, much younger women or girls as it was in the past, and such relationships are now less common. Indeed, in the last few years, this has become especially socially unacceptable. Therefore, given that most men are maximally attracted to females in this age category, they prefer to be thought of as younger so that it is more acceptable for them to seek relationships with younger, more attractive females.
Actually, while older men tend to have higher status on average, I suspect that, after controlling for status, it is younger men who would be perceived as more attractive. Certainly, a young multi-millionaire would surely be considered a more eligible bachelor than an older homeless man. Therefore, age per se is not attractive; only high status is attractive, which happens to correlate with age.

[22] This idea is again based on Philippe Rushton’s Differential K theory, which I have reviewed here and here.

[23] Dutton is apparently aware of this objection. He acknowledges, albeit in a different book, that “Intelligence, in general, is associated with health” (Why Islam Makes You Stupid: p174). However, in this same book, he also claims that: 

Intelligence has been shown to be only weakly associated with mutational load” (Why Islam Makes You Stupid: p169). 

Interestingly, Dutton also claims in this book: 

Very high intelligence predicts autism” (Why Islam Makes You Stupid: p175). 

This claim, namely that exceptionally high intelligence is associated with autism, seems anecdotally plausible. Certainly, autism seems to have a complex and interesting relationship with intelligence
Unfortunately, however, Dutton does not cite a source for the claim the claim that exceptionally high intelligence is associated with autism. Nevertheless, according to data cited here, there is indeed a greater variance in the IQs of autistic people, with greater proportions of autistic people at both tail-ends of the bell curve, the author even referring to an inverted bell curve for intelligence among autistic people, though, even according to her own cited data, this appears to be an exaggeration. However, this is not a scholarly source, but rather appears to be the website of a not entirely disinterested advocacy group, and it is not entirely clear from where this data derives, the piece referring only to data from the Netherlands collected by the Dutch Autism Register (NAR). 

[24] Admittedly, Dutton does cite one study showing that subjects can identify Mormons from facial photographs alone, and that the two groups differed in skin quality (Rule et al 2010). However, this might reflect merely the health advantages resulting from the religiously imposed abstention from the consumption of alcohol, tobacco, tea and coffee.
For what it’s worth, my own subjective and entirely anecdotal impression is almost the opposite of Dutton’s, at least here in secular modern Britain, where anyone who identifies as Christian, let alone a fundamentalist, unless perhaps s/he is elderly, tends to be regarded as a bit odd.
An interesting four-part critique of this theory, along very different lines from my own, is provided by Scott A McGreal at the Psychology Today website, see here, here, here, and here. Dutton responds with a two-part rejoinder here and here.

[25] However, when it comes to actual politicians, I suspect this difference may be attenuated, or even nonexistent, since pursuing a career in politics is, by its very nature, a very untraditional, and unfeminine, career choice, most likely because, in Darwinian terms, political power has a greater reproductive payoff for men than for women. Thus, it is hardly surprising that leading female politicians, even those who theoretically champion traditional sex roles, tend themselves to be quite butch and masculine in appearance and often as unattractive as their leftist opponents (e.g. Ann Widdecombe). Indeed, even Ann Coulter, a relatively attractive woman, at least by the standards of female political figures, has been mocked for her supposedly mannish appearance and pronounced Adam’s apple.
Moreover, most leading politicians are at least middle-aged, and female attractiveness peaks very young, in mid- to late-teens into early-twenties

[26] Another medical condition associated with a specific look, as well as with mental disability, is cretinism, though due to medical advances, most people with the condition in western societies, develop normally and no longer manifest either the distinctive appearance or the mental disability. 

References 

Aldridge et al (2011) Facial phenotypes in subgroups of prepubertal boys with autism spectrum disorders are correlated with clinical phenotypes. Molecular Autism 14;2(1):15. 
Apicella et al (2015) Hadza Hunter-Gatherer Men do not Have More Masculine Digit Ratios (2D:4D) American Journal of Physical Anthropology 159(2):223-32. 
Bagenjuk et al (2019) Personality Traits and Obesity, International Journal of Environmental Research and Public Health 16(15): 2675. 
Bailey & Hurd (2005) Finger length ratio (2D:4D) correlates with physical aggression in men but not in women. Biological Psychology 68(3):215-22. 
Batrinos (2014) The endocrinology of baldness. Hormones 13(2): 197–212. 
Beckley et al (2014) Association of height and violent criminality: results from a Swedish total population study. International Journal of Epidemiology 43(3):835-42 
Benderlioglu & Nelson (2005) Digit length ratios predict reactive aggression in women, but not in men Hormones and Behavior 46(5):558-64. 
Berggren et al (2017) The right look: Conservative politicians look better and voters reward it Journal of Public Economics 146:  79-86. 
Blanchard & Lyons (2010) An investigation into the relationship between digit length ratio (2D: 4D) and psychopathy, British Journal of Forensic Practice 12(2):23-31. 
Buru et al (2017) Evaluation of the hand anthropometric measurement in ADHD children and the possible clinical significance of the 2D:4D ratioEastern Journal of Medicine 22(4):137-142. 
Case & Paxson (2008) Stature and status: Height, ability, and labor market outcomes, Journal of Political Economy 116(3): 499–532. 
Cash (1990) Losing Hair, Losing Points?: The Effects of Male Pattern Baldness on Social Impression Formation. Journal of Applied Social Psychology 20(2):154-167. 
De Waal et al (1995) High dose testosterone therapy for reduction of final height in constitutionally tall boys: Does it influence testicular function in adulthood? Clinical Endocrinology 43(1):87-95. 
Dixson & Vasey (2012) Beards augment perceptions of men’s age, social status, and aggressiveness, but not attractiveness, Behavioral Ecology 23(3): 481–490. 
Dutton et al (2017) The Mutant Says in His Heart, “There Is No God”: the Rejection of Collective Religiosity Centred Around the Worship of Moral Gods Is Associated with High Mutational Load Evolutionary Psychological Science 4:233–244. 
Dysniku et al (2015) Minor Physical Anomalies as a Window into the Prenatal Origins of Pedophilia, Archives of Sexual Behavior 44:2151–2159. 
Elias et al (2012) Obesity, Cognitive Functioning and Dementia: Back to the Future, Journal of Alzheimer’s Disease 30(s2): S113-S125. 
Ellis & Hoskin (2015) Criminality and the 2D:4D Ratio: Testing the Prenatal Androgen Hypothesis, International Journal of Offender Therapy and Comparative Criminology 59(3):295-312 
Fossen et al (2022) 2D:4D and Self-Employment: A Preregistered Replication Study in a Large General Population Sample Entrepreneurship Theory and Practice 46(1):21-43. 
Gouchie & Kimura (1991) The relationship between testosterone levels and cognitive ability patterns Psychoneuroendocrinology 16(4): 323-334. 
Griffin et al (2012) Varsity athletes have lower 2D:4D ratios than other university students, Journal of Sports Sciences 30(2):135-8. 
Hilgard et al (2019) Null Effects of Game Violence, Game Difficulty, and 2D:4D Digit Ratio on Aggressive Behavior, Psychological Science 30(1):095679761982968 
Hollier et al (2015) Adult digit ratio (2D:4D) is not related to umbilical cord androgen or estrogen concentrations, their ratios or net bioactivity, Early Human Development 91(2):111-7 
Hönekopp & Urban (2010) A meta-analysis on 2D:4D and athletic prowess: Substantial relationships but neither hand out-predicts the other, Personality and Individual Differences 48(1):4-10. 
Hoskin & Ellis (2014) Fetal testosterone and criminality: Test of evolutionary neuroandrogenic theory, Criminology 53(1):54-73. 
Ishikawa et al (2001) Increased height and bulk in antisocial personality disorder and its subtypes. Psychiatry Research 105(3):211-219. 
Işık et al (2020) The Relationship between Second-to-Fourth Digit Ratios, Attention-Deficit/Hyperactivity Disorder Symptoms, Aggression, and Intelligence Levels in Boys with Attention-Deficit/Hyperactivity Disorder, Psychiatry Investigation 17(6):596–602. 
Janowski et al (1994) Testosterone influences spatial cognition in older men. Behavioral Neuroscience 108(2):325-32. 
Jokela et al (2012) Association of personality with the development and persistence of obesity: a meta-analysis based on individual–participant data, Etiology and Pathophysiology 14(4): 315-323. 
Kanazawa (2014) Intelligence and obesity: Which way does the causal direction go? Current Opinion in Endocrinology, Diabetes and Obesity (5):339-44. 
Kandel et al (1989) Minor physical anomalies and recidivistic adult violent criminal behavior, Acta Psychiatrica Scandinavica 79(1) 103-107. 
Kangassalo et al (2011) Prenatal Influences on Sexual Orientation: Digit Ratio (2D:4D) and Number of Older Siblings, Evolutionary Psychology 9(4):496-508 
Kerry & Murray (2019) Is Formidability Associated with Political Conservatism?  Evolutionary Psychological Science 5(2): 220–230. 
Keshavarz et al (2017) The Second to Fourth Digit Ratio in Elite and Non-Elite Greco-Roman Wrestlers, Journal of Human Kinetics 60: 145–151. 
Kleisner et al (2014) Perceived Intelligence Is Associated with Measured Intelligence in Men but Not Women. PLoS ONE 9(3): e81237. 
Kordsmeyer et al (2018) The relative importance of intra- and intersexual selection on human male sexually dimorphic traits, Evolution and Human Behavior 39(4): 424-436. 
Kosinski & Wang (2018) Deep neural networks are more accurate than humans at detecting sexual orientation from facial images. Journal of Personality and Social Psychology 114(2):246-257. 
Kyselicová et al (2021) Autism spectrum disorder and new perspectives on the reliability of second to fourth digit ratio Developmental Pyschobiology 63(6). 
Li et al (2016) The relationship between digit ratio and sexual orientation in a Chinese Yunnan Han population, Personality and Individual Differences 101:26-29. 
Lippa (2003) Are 2D:4D finger-length ratios related to sexual orientation? Yes for men, no for women, Journal of Personality &Social Psychology 85(1):179-8 
Lolli et al (2017) A comprehensive allometric analysis of 2nd digit length to 4th digit length in humans, Proceedings of the Royal Society B: Biological Sciences 284(1857):20170356 
Malamed (1992) Personality correlates of physical height. Personality and Individual Differences 13(12):1349-1350. 
Manning & Taylor (2001) Second to fourth digit ratio and male ability in sport: implications for sexual selection in humans, Evolution & Human Behavior 22(1):61-69. 
Manning et al (2001) The 2nd to 4th digit ratio and autism, Developmental Medicine & Child Neurology 43(3):160-164. 
Martel et al (2008) Masculinized Finger-Length Ratios of Boys, but Not Girls, Are Associated With Attention-Deficit/Hyperactivity Disorder, Behavioral Neuroscience 122(2):273-81. 
Martin et al (2015) Associations between obesity and cognition in the pre-school years, Obesity 24(1) 207-214 
Mazur & Booth (1998) Testosterone and dominance in men. Behavioral and Brain Sciences, 21(3), 353–397. 
Mazur (2009) Testosterone and violence among young men. In Walsh & Beaver (eds) Biosocial Criminology: New Directions in theory and Research. New York: Routledge. 
Moffat & Hampson (1996) A curvilinear relationship between testosterone and spatial cognition in humans: Possible influence of hand preference. Psychoneuroendocrinology. 21(3):323-37. 
Murray et al (2014) How are conscientiousness and cognitive ability related to one another? A re-examination of the intelligence compensation hypothesis, Personality and Individual Differences, 70, 17–22. 
Nieshclag & Behr (2013) Testosterone Therapy. In Nieschlag & Behr (eds) Andrology: Male Reproductive Health and Dysfunction. New York: Springer. 
Ozgen et al (2010) Minor physical anomalies in autism: a meta-analysis. Molecular Psychiatry 15(3):300–7. 
Ozgen et al (2011) Morphological features in children with autism spectrum disorders: a matched case-control study. Journal of Autism and Developmental Disorders 41(1):23-31. 
Peterson & Palmer (2017) Effects of physical attractiveness on political beliefs. Politics and the Life Sciences 36(02):3-16 
Persico et al (2004) The Effect of Adolescent Experience on Labor Market Outcomes: The Case of Height, Journal of Political Economy 112(5): 1019-1053. 
Pratt et al (2016) Revisiting the criminological consequences of exposure to fetal testosterone: a meta-analysis of the 2d:4d digit ratio, Criminology 54(4):587-620. 
Price et al (2017). Is sociopolitical egalitarianism related to bodily and facial formidability in men? Evolution and Human Behavior, 38, 626-634. 
Puts (2010) Beauty and the beast: Mechanisms of sexual selection in humans, Evolution and Human Behavior 31(3):157-175. 
Rammstedt et al (2016) The association between personality and cognitive ability: Going beyond simple effects, Journal of Research in Personality 62: 39-44. 
Rosenberg & Kagan (1987) Iris pigmentation and behavioral inhibition Developmental Psychobiology 20(4):377-92. 
Rosenberg & Kagan (1989) Physical and physiological correlates of behavioral inhibition Developmental Psychobiology 22(8):753-70. 
Rule et al (2010) On the perception of religious group membership from faces. PLoS ONE 5(12):e14241. 
Salas-Wright & Vaughn (2016) Size Matters: Are Physically Large People More Likely to be Violent? Journal of Interpersonal Violence 31(7):1274-92. 
Spurk & Abele (2011) Who Earns More and Why? A Multiple Mediation Model from Personality to Salary, Journal of Business and Psychology 26: 87–103. 
Sutin et al (2011) Personality and Obesity across the Adult Lifespan Journal of Personality and Social Psychology 101(3): 579–592. 
Valla et al (2011). The accuracy of inferences about criminality based on facial appearance. Journal of Social, Evolutionary, and Cultural Psychology, 5(1), 66-91. 
Voracek et al (2011) Digit ratio (2D:4D) and sex-role orientation: Further evidence and meta-analysis, Personality and Individual Differences 51(4): 417-422. 
Weinberg et al (2007) Minor physical anomalies in schizophrenia: A meta-analysis, Schizophrenia Research 89: 72–85. 
Wiersma & Kappe 2015 Selecting for extroversion but rewarding for conscientiousness, European Journal of Work and Organizational Psychology 26(2): 314-323. 
Williams et al (2000) Finger-Length Ratios and Sexual Orientation, Nature 404(6777):455-456. 
Xu et al (2011) Minor physical anomalies in patients with schizophrenia, unaffected first-degree relatives, and healthy controls: a meta-analysis, PLoS One 6(9):e24129. 
Xu & Zheng (2016) The Relationship Between Digit Ratio (2D:4D) and Sexual Orientation in Men from China, Archives of Sexual Behavior 45(3):735-41. 

Pierre van den Berghe’s ‘The Ethnic Phenomenon’: Ethnocentrism and Racism as Nepotism Among Extended Kin

Pierre van den Berghe, The Ethnic Phenomenon (Westport: Praeger 1987) 

Ethnocentrism is a pan-human universal. Thus, a tendency to prefer one’s own ethnic group over and above other ethnic groups is, ironically, one thing that all ethnic groups share in common. 

In ‘The Ethnic Phenomenon’, pioneering sociologist-turned-sociobiologist Pierre van den Berghe attempts to explain this universal phenomenon. 

In the process, he not only provides a persuasive ultimate evolutionary explanation for the universality of ethnocentrism, but also produces a remarkable synthesis of scholarship that succeeds in incorporating virtually every aspect of ethnic relations as they have manifested themselves throughout history and across the world, from colonialism, caste and slavery to integration and assimilation, within this theoretical and explanatory framework. 

Ethnocentrism as Nepotism? 

At the core of Pierre van den Berghe’s theory of ethnocentrism and ethnic conflict is the sociobiological theory of kin selection. According to van den Berghe, racism, xenophobia, nationalism and other forms of ethnocentrism can ultimately be understood as kin-selected nepotism, in accordance with biologist William D Hamilton’s theory of inclusive fitness (Hamilton 1964a; 1964b). 

According to inclusive fitness theory (also known as kin selection), organisms evolved to behave altruistically towards their close biological kin, even at a cost to themselves, because close biological kin share genes in common with one another by virtue of their kinship, and altruism towards close biological kin therefore promotes the survival and spread of these genes. 

Van den Berghe extends this idea, arguing that humans have evolved to sometimes behave altruistically towards, not only their close biological relatives, but also sometimes their distant biological relatives as well – namely, members of the same ethnic group as themselves. 

Thus, van den Berghe contends: 

Racial and ethnic sentiments are an extension of kinship sentiments [and] ethnocentrism and racism are… extended forms of nepotism” (p18). 

Thus, while social scientists, and social psychologists in particular, rightly emphasize the ubiquity, if not universality, of in-group preference, namely a preference for and favouring of individuals of the same social group as oneself, they also, in my view, rather underplay the extent to which the group identities which lead to the most conflict, animosity, division and discrimination, not only in the contemporary west, but throughout history and across the world, and are also most apparently impervious to resolution, are ethnic identities.

Thus, divisions such as those between social classes, or the sexes, different generations, or between members of different political factions, or youth subcultures (e.g. between mods’ and ‘rockers), or supporters of different sports teams, may indeed lead to substantial conflict, at least in the short-term, and are often cited as quintessential examplars of ‘tribal’ identity and conflict.

However, the most violent and intransient of group conflicts seem to me to be those between ethnic groups, namely a form of group identity that is passed down in families, from parent to offspring, in a quasi-biological fashion, based on a perception of shared kinship, and in respect of which people are usually expected to marry endogamously.

In contrast, aspects of group identity that vary even between individuals within a single family, including those that are freely chosen by individuals, tend to be somewhat muted in intensity, perhaps precisely because most people share bonds with close family members of a different group identity.

Thus, there has never, to my knowledge, been a civil war arising from conflict between the sexes, or between supporters of one or another football team.[1]

Ethnic Groups as Kin Groups?

Before reading van den Berghe’s book, I was skeptical regarding whether the degree of kinship shared among co-ethnics would ever be sufficient to satisfy Hamilton’s rule, whereby, for altruism to evolve, the cost of the altruistic act to the altruist, measured in terms of reproductive success, must be outweighed by the benefit to the recipient, also measured in terms of reproductive success, multiplied by the degree of relatedness of the two parties (Brigandt 2001; cf. Salter 2008; see also On Genetic Interests). 

Thus, Brigandt (2001) takes van den Berghe to task for his formulation of what the latter catchily christens “the biological golden rule”, namely: 

Give unto others as they are related unto you” (p20).[2]

However, contrary to both critics of his theory (e.g. Brigandt 2001) and others developing similar ideas (e.g. Rushton 2005; Salter 2000), van den Berghe is actually agnostic on the question of whether ethnocentrism is ever actually adaptive in modern societies, where the shared kinship of large nations or ethnic groups is, as van den Berghe himself readily acknowledges, “extremely tenuous at best” (p243). Thus, he concedes: 

Clearly, for 50 million Frenchmen or 100 million Japanese, any common kinship that they may share is highly diluted … [and] when 25 million African-Americans call each other ‘brothers’ and ‘sisters’, they know that they are greatly extending the meaning of these terms” (p27).[3]

Instead, van den Berghe suggests that nationalism and racism may reflect the misfiring of a mechanism that evolved when our ancestors still still lived in small kin-based groups of hunter-gatherers that represented little more than extended families (p35; see also Tooby and Cosmides 1989; Johnson 1986). 

Thus, van den Berghe explains: 

Until the last few thousand years, hominids interacted in relatively small groups of a few score to a couple of hundred individuals who tended to mate with each other and, therefore, to form rather tightly knit groups of close and distant kin” (p35). 

Therefore, in what evolutionary psychologists now call the environment of evolutionary adaptedness or EEA:

The natural ethny [i.e. ethnic group] in which hominids evolved for several thousand millennia probably did not exceed a couple of hundred individuals at most” (p24) 

Thus, van den Berghe concludes: 

The primordial ethny is thus an extended family: indeed, the ethny represents the outer limits of that inbred group of near or distant kinsmen whom one knows as intimates and whom therefore one can trust” (p25). 

On this view, ethnocentrism was adaptive when we still resided in such groups, where members of our own clan or tribe were indeed closely biologically related to us, but is often maladaptive in contemporary environments, where our ethnic group may include literally millions of people. 

Another not dissimilar theory has it that racism in particular might reflect the misfiring of an adaptation that uses phenotype matching, in particular physical resemblance, as a form of kin recognition

Thus, Richard Dawkins in his seminal The Selfish Gene (which I have reviewed here), cautiously and tentatively speculates: 

Conceivably, racial prejudice could be interpreted as an irrational generalization of a kin-selected tendency to identify with individuals physically resembling oneself, and to be nasty to individuals different in appearance” (The Selfish Gene: p100). 

Certainly, van den Berghe takes pains to emphasize that ethnic sentiments are vulnerable to manipulation – not least by exploitative elites who co-opt kinship terms such as ‘motherland’, fatherland and ‘brothers-in-arms‘ to encourage self-sacrifice, especially during wartime (p35; see also Johnson 1987; Johnson et al 1987; Salmon 1998). 

However, van den Berghe cautions, “Kinship can be manipulated but not manufactured [emphasis in original]” (p27). Thus, he observes how: 

Queen Victoria could cut a motherly figure in England; she even managed to proclaim her son the Prince of Wales; but she could never hope to become anything except a foreign ruler of India; [while] the fiction that the Emperor of Japan is the head of the most senior lineage descended from the common ancestor of all Japanese might convince the Japanese peasant that the Emperor is an exalted cousin of his, but the myth lacks credibility in Korea or Taiwan” (p62-3). 

This suggests that the European Union, while it may prove successful as customs union, single market and even an economic union, and while integration in other non-economic spheres may also prove a success, will likely never command the sort of loyalty and allegiance that a nation-state holds over its people, including, sometimes, the willingness of men to fight and lay down their lives for its sake. This is because its members come from many different cultures and ethnicities, and indeed speak many different languages. 

For van den Berghe, national identity cannot be rooted in anything other than a perception of shared ancestry or kinship. Thus, he observes: 

Many attempts to adopt universalistic criteria of ethnicity based on legal citizenship or acquisition of educational qualifications… failed. Such was the French assimilation policy in her colonies. No amount of proclamation of Algérie française could make it so” (p27). 

Thus, so-called civic nationalism, whereby national identity is based, not on ethnicity, but rather, supposedly, on a shared commitment to certain common values and ideals (democracy, the ‘rule of law’ etc.), as encapsulated by the notion of America as a proposition nation’, is, for van den Berghe, a complete non-starter. 

Yet this is today regarded as the sole basis for national identity and patriotic feeling that is recognised as legitimate, not only in the USA, but also all other contemporary western polities, where any assertion of racial nationalism or a racially-based or ethnically-based national identity is, at least for white people, anathema and beyond the pale. 

Moreover, due to the immigration policies of previous generations of western political leaders, policies that largely continue today, all contemporary western polities are now heavily multi-ethnic and multi-racial, such that any sense of national identity that was based on race or ethnicity is arguably untenable as it would necessarily exclude a large proportion of their populations.

On the other hand, however, van den Berghe’s reasoning also suggests that the efforts of some white nationalists to construct a pan-white, or pan-European, ethnic identity is also, like the earlier efforts of Japanese imperialist propagandists to create a pan-Asian identity, and of Marcus Garvey’s UNIA to construct a pan-African identity, likely to end in failure.[4]

Racism vs Ethnocentrism 

Whereas ethnocentrism is therefore universal, adaptive and natural, van den Berghe denies that the same can be said for racism

There is no evidence that racism is inborn, but there is considerable evidence that ethnocentrism is” (p240). 

Thus, van den Berge concludes: 

The genetic propensity is to favor kin, not those who look alike” (p240).[5]

As evidence, he cites:

The ease with which parental feelings take precedence over racial feeling in cases of racial admixture” (p240). 

In other words, fathers who sire mixed-race offspring with women of other races, and the women of other races with whom they father such offspring, often seemingly love and care for the resulting offspring just as intensely as do parents whose offspring is of the same race as themselves.[6]

Thus, cultural, rather than racial, markers are typically adopted to distinguish ethnic groups (p35). These include: 

  • Clothing (e.g. hijabs, turbans, skullcaps);
  • Bodily modification (e.g. tattoos, circumcision); and 
  • Behavioural criteria, especially language and dialect (p33).

Bodily modification and language represent particularly useful markers because they are difficult to fake, bodily modification because it is permanent and hence represents a costly commitment to the group (in accordance with Zahavi’s handicap principle), and language/dialect, because this is usually acquirable only during a critical period during childhood, after which it is generally not possible to achieve fluency in a second language without retaining a noticeable accent. 

In contrast, racial criteria, as a basis for group affiliation, is, van den Berghe reports, actually quite rare: 

Racism is the exception rather than the rule in intergroup relations” (p33). 

Racism is also a decidedly modern phenomenon. 

This is because, prior to recent technological advances in transportation (e.g. ocean-going ships, aeroplanes), members of different races (i.e. groups distinguishable on the basis of biologically inherited physiological traits such as skin colour, nose shape, hair texture etc.) were largely separated from one another by the very geographic barriers (e.g. deserts, oceans, mountain ranges) that reproductively isolated them from one another and hence permitted their evolution into distinguishable races in the first place. 

Moreover, when different races did make contact, then, in the absence of strict barriers to exogamy and miscegenation (e.g. the Indian caste system), racial groups typically interbred with one another and hence become phenotypically indistinguishable from one another within just a few generations. 

This, van den Berghe explains, is because: 

Even the strongest social barriers between social groups cannot block a specieswide [sic] sexual attraction. The biology of reproduction triumphs in the end over the artificial barriers of social prejudice” (p109). 

Therefore, in the ancestral environment for which our psychological adaptations are designed (i.e. before the development of ships, aeroplanes and other methods of long-distance intercontinental transportation), different races did not generally coexist in the same locale. As a result, van den Berghe concludes: 

We have not been genetically selected to use phenotype as an ethnic marker, because, until quite recently, such a test would have been an extremely inaccurate one” (p 240). 

Humans, then, have simply not had sufficient time to have evolved a domain-specificracism module’ as suggested by some researchers.[7]

Racism is therefore, unlike ethnocentrism, not an innate instinct, but rather “a cultural invention” (p240). 

However, van den Berghe rejects the fashionable, politically correct notion that racism is “a western, much less a capitalist monopoly” (p32). 

On the contrary, racism, while not innate, is, not a unique western invention, but rather a recurrent reinvention, which almost invariably arises where phenotypically distinguishable groups come into contact with one another, if only because: 

Genetically inherited phenotypes are the easiest, most visible and most reliable predictors of group membership” (p32).

For example, van den Berghe describes the relations between the Tutsi, Hutu and Pygmy Twa of Rwanda and neighbouring regions as “a genuine brand of indigenous racism” which, according to van den Berghe, developed quite independently of any western colonial influence (p73).[8]

Moreover, where racial differences are the basis for ethnic identity, the result is, van den Berghe claims, ethnic hierarchies that are particularly rigid, intransient and impermeable.

For van den Berghe, this then explains the failure of African-Americans to wholly assimilate into the US melting pot in stark contrast to successive waves of more recently-arrived European immigrants. 

Thus, van den Berghe observes: 

Blacks who have been English-speaking for several generations have been much less readily assimilated in both England… and the United States than European immigrants who spoke no English on arrival” (p219). 

Thus, language barriers often break down within a generation. 

As Judith Harris emphasizes in support of peer group socialization theory, the children of immigrants whose parents are not at all conversant in the language of their host culture nevertheless typically grow up to speak the language of their host culture rather better than they do the first language of their parents, even though the latter was the cradle tongue to which they were first exposed, and first learnt to speak, inside the family home (see The Nurture Assumption: which I have reviewed here). 

As van den Berghe observes: 

It has been the distressing experience of millions of immigrant parents that, as soon as their children enter school in the host country, the children begin to resist speaking their mother tongue” (p258). 

While displeasing to those parents who wish to pass on their language, culture and traditions to their offspring, this response is wholly adaptive from the perspective of the offspring themselves:  

Children quickly discover that their home language is a restricted medium that not useable in most situations outside the family home. When they discover that their parents are bilingual they conclude – rightly for their purposes – that the home language is entirely redundant… Mastery of the new language entails success at school, at work and in ‘the world’… [against which] the smiling approval of a grandmother is but slender counterweight” (p258).[9]

However, whereas one can learn a new language, it is not usually possible to change one’s race – the efforts of Rachel Dolezal, Elizabeth Warren, Jessica Krug and Michael Jackson notwithstanding. However, due to the one-drop rule and the history of miscegenation in America, passing is sometimes possible (see below). 

Instead, phenotypic (i.e. racial) differences can only be eradicated after many generations of miscegenation, and sometimes, as in the cases of countries like the USA and Brazil, not even then. 

Meanwhile, van den Berghe observes, often the last aspect of immigrant culture to resist assimilation is culinary differences. However, he observes, increasingly even this becomes only a ‘ceremonial’ difference reserved for family gatherings (p260). 

Thus, van den Berghe surmises, Italian-Americans probably eat hamburgers as often as Americans of any other ethnic background, but at family gatherings they still revert to pasta and other traditional Italian cuisine

Yet even culinary differences eventually disappear. Thus, in both Britain and America, sausage has almost completely ceased to be thought of as a distinctively German dish (as have hamburgers, originally thought to have been named in reference to the city of Hamburg) and now pizza is perhaps on the verge of losing any residual association with Italians. 

Is Racism Always Worse than Ethnocentrism? 

Yet if raciallybased ethnic hierarchies are particularly intransigent and impermeable, they are also, van den Berghe claims, “peculiarly conflict-ridden and unstable” (p33). 

Thus, van den Berghe seems to believe that racial prejudice and animosity tends to be more extreme and malevolent in nature than mere ethnocentrism as exists between different ethnic groups of the same race (i.e. not distinguishable from one another on the basis of inherited phenotypic traits such as skin colour). 

For example, van den Berghe claims that, during World War Two: 

There was a blatant difference in the level of ferociousness of American soldiers in the Pacific and European theaters… The Germans were misguided relatives (however distant), while the ‘Japs’ or the ‘Nips’ were an entirely different breed of inscrutable, treacherous, ‘yellow little bastards.’ This was reflected in differential behavior in such things as the taking (versus killing) of prisoners, the rhetoric of war propaganda (President Roosevelt in his wartime speeches repeatedly referred to his enemies as ‘the Nazis, the Fascists, and the Japanese’), the internment in ‘relocation camps’ of American citizens of Japanese extraction, and in the use of atomic weapons” (p57).[10]

Similarly, in his chapter on ‘Colonial Empires’, by which he means “imperialism over distant peoples who usually live in noncontiguous territories and who therefore look quite different from their conquerors, speak unrelated languages, and are so culturally alien to their colonial masters as to provide little basis for mutual understanding”, van den Berghe writes: 

Colonialism is… imperialism without the restraints of common bonds of history, culture, religion, marriage and blood that often exist when conquest takes place between neighbors” (p85). 

Thus, he claims: 

What makes for the special character of the colonial situation is the perception by the conqueror that he is dealing with totally unrelated, alien and, therefore, inferior people. Colonials are treated as people totally beyond the pale of kin selection” (p85). 

However, I am unpersuaded by van den Berghe’s claim that conflict between more distantly related ethnic groups is always, or even typically, more brutal than that among biologically and culturally more closely related groups. 

After all, even conquests of neighbouring peoples, identical in race, if not always in culture, to the conquering group, are often highly brutal, for example the British in Ireland or the Japanese in Korea and China in the first half of the twentieth century. 

Indeed, many of the most intense and intractable ethnic conflicts are those between neighbours and ethnic kin, who are racially (and culturally) very similar to one another. 

Thus, for example, Catholics and Protestants in Northern Ireland, Greeks and Turks in Cyprus, and Bosnians, Croats, Serbs and Albanians in the Balkans, and even Jews and Palestinians in the Middle East, are all racially and genetically quite similar to one another, and also share many aspects of their culture with one another too. (The same is true, to give a topical example at the time of writing, of Ukrainians and Russians.) However, this has not noticeably ameliorated the nasty, intransient and bloody conflicts that have been, and continue to be, waged among them.  

Of course, the main reason that most ethnic conflict occurs between close neighbours is because neighbouring groups are much more likely to come into contact, and hence into conflict, with one another, especially over competing claims to land.[11]

Yet these same neighbouring groups are also likely to be related to one another, both culturally and genetically, because of both shared origins and the inevitable history of illicit intermarriage or miscegenation, and cultural borrowings, that inevitably occur even among the most hostile of neighbours.[12]

Nevertheless, the continuation of intense ethnic animosity between ethnic groups who are genetically, close to one another seems to pose a theoretical problem, not only for van den Berghe’s theory, but also, to an even greater degree, for Philippe Rushton’s so-called genetic similarity theory (which I have written about here), which argues that conflict between different ethnic groups is related to their relative degree of genetic differentiation from one another (Rushton 1998a; 1998b; 2005). 

It also poses a problem for the argument of political scientist Frank K Salter, who argues that populations should resist immigration by alien immigrants proportionally to the degree to which the alien immigrants are genetically distant from themselves (On Genetic Interests; see also Salter 2002). 

Assimilation, Acculturation and the American Melting Pot 

Since racially-based hierarchies result in ethnic boundaries that are both “peculiarly conflict-ridden and unstable” and also peculiarly rigid and impermeable, Van den Berghe controversially concludes: 

There has never been a successful multiracial democracy” (p189).[13]

Of course, in assessing this claim, we must recognize that ‘success’ is not only a matter of degree, but also can also be measured on several different dimensions. 

Thus, many people would regard the USA as the quintessential “successful… democracy”, even though the US has been multiracial, to some degree, for the entirety of its existence as a nation. 

Certainly, the USA has been successful economically, and indeed militarily.

However, the US has also long been plagued by interethnic conflict, and, although successful economically and militarily, it has yet to be successful in finding a way to manage its continued interethnic conflict, especially that between blacks and whites.

The USA is also afflicted with a relatively high rate of homicide and gun crime as compared to other developed economies, as well as low levels of literacy and numeracy and educational attainment. Although it is politically incorrect to acknowledge as much, these problems also likely reflect the USA’s ethnic diversity, in particular its large black underclass.

Indeed, as van den Berghe acknowledges, even societies divided by mere ethnicity rather than race seem highly conflict-prone (p186). 

Thus, assimilation, when it does occur, occurs only gradually, and only under certain conditions, namely when the group which is to be assimilated is “similar in physical appearance and culture to the group to which it assimilates, small in proportion to the total population, of low status and territorially dispersed” (p219). 

Thus, van den Berghe observes: 

People tend to assimilate and acculturate when their ethny [i.e. ethnic group] is geographically dispersed (often through migration), when they constitute a numerical minority living among strangers, when they are in a subordinate position and when they are allowed to assimilate by the dominant group” (p185). 

Moreover, van den Berghe is careful distinguish what he calls assimilation from mere acculturation.  

The latter, acculturation, involves a subordinate group gradually adopting the norms, values, language, cultural traditions and folkways of the dominant culture into whom they aspire to assimilate. It is therefore largely a unilateral process.[14]

In contrast, however, assimilation goes beyond this and involves members of the dominant host culture also actually welcoming, or at least accepting, the acculturated newcomers as a part of their own community.  

Thus, van den Berghe argues that host populations sometimes resist the assimilation of even wholly acculturated and hence culturally indistinguishable out-groups. Examples of groups excluded in this way include, according to van den Berghe, pariah castes, such as the untouchable dalits of the Indian subcontinent, the Burakumin of Japan and blacks in the USA.[15]

In other words, assimilation, unlike acculturation, is very much a two-way street. Thus, just as it ‘takes two to tango’, so assimilation is very much a bilateral process: 

It takes two to assimilate” (p217).  

On the one hand, minority groups may sometimes themselves resist assimilation, or even acculturation, if they perceive themselves as better off maintaining their distinct identify. This is especially true of groups who perceive themselves as being, in some respects, better-off than the host outgroup into whom they refuse to be absorbed. 

Thus, middleman minorities, or market-dominant minorities, such as Jews in the West, the overseas Chinese in contemporary South-East Asia, the Lebanese in West Africa and South Asians in East Africa, being, on average, much wealthier than the bulk of the host populations among whom them live, often perceive no social or economic advantage to either assimilation or acculturation and hence resist the process, instead stubbornly maintaining their own language and traditions and marrying only among themselves. 

The same is also true, more obviously, of alien ruling elites, such as the colonial administrators, and settlers, in European colonial empires in Africa, India and elsewhere, for whom assimilation into native populations would have been anathema.

Passing’, ‘Pretendians’ and ‘Blackfishing’ 

Interestingly, just as market-dominant minorities, middleman minorities, and European colonial rulers usually felt no need to assimilate into the host society in whose midst they lived, because to do so would have endangered their privileged position within this host society, so recent immigrants to America may no longer perceive any advantage to assimilation. 

On the contrary, there may now be an economic disincentive operating against assimilation, at least if assimilation means forgoing from the right to benefit from affirmative action in employment and college admissions

Thus, in the nineteenth and early twentieth centuries, the phenomenon of passing, at least in America, typically involved non-whites, especially light-skinned mixed-race African-Americans, attempting to pass as white or, if this were not realistic, sometimes as Native American.  

Some non-whites, such as Bhagat Singh Thind and Takao Ozawa, even brought legal actions in order to be racially reclassified as ‘white’ in order to benefit from America’s then overtly racialist naturalization law.

Contemporary cases of passing, however, though rarely referred to by this term, typically involve whites themselves attempting to somehow pass themselves off as some variety of non-white (see Hannam 2021). 

Recent high-profile recent examples have included Rachel Dolezal, Elizabeth Warren and Jessica Krug

Interestingly, all three of these women were both employed in academia and involved in leftist politics – two spheres in which adopting a non-white identity is likely to be especially advantageous, given the widespread adoption of affirmative action in college admissions and appointments, and the rampant anti-white animus that infuses so much of academia and the cultural Marxist left.[16]

Indeed, the phenomenon is now so common that it even has its own associated set of neologisms, such as Pretendian, ‘blackfishing’ and, in Australia, box-ticker.[17]

Indeed, one remarkable recent survey purported to uncover that fully 34% of white college applicants in the United States admitted to lying about their ethnicity on their applications, in most cases either to improve their chances of admission or to qualify for financial aid

Although Rachel Dolezal, Elizabeth Warren and Jessica Krug were all women, this survey found that white male applicants were even more likely to lie about their ethnicity than were white female applicants, with only 16% of white female applicants admitting to lying, as compared to nearly half (48%) of white males.[18]

This is, of course, consistent with the fact that it is white males who are the primary victims of affirmative action and other forms of discrimination.  

This strongly suggests that, whereas there were formerly social (and legal) benefits that were associated with identifying as white, today the advantages accrue to instead to those able to assume a non-white identity.  

For all the talk of so-called ‘white privilege’, when whites and mixed-race people, together with others of ambiguous racial identity, preferentially choose to pose as non-white in order to take advantage of the perceived benefits of assuming such an identity, they are voting with their feet and thereby demonstrating what economists call revealed preferences

This, of course, means that recent immigrants to America, such as Hispanics, will have rather less incentive in integrate into the American mainstream than did earlier waves of European immigrants, such as Irish, Poles, Jews and Italians, the latter having been, primarily, the victims of discrimination rather than its beneficiaries

After all, who would want to be another, boring unhyphenated American when to do so would presumably mean relinquishing any right to benefit from affirmative action in job recruitment or college admissions, not to mention becoming a part of the hated white ‘oppressor’ class. 

In short, ‘white privilege’ isn’t all it’s cracked up to be. 

This perverse incentive against assimilation obviously ought to be worrying to anyone concerned with the future of American as a stable unified polity. 

Ethnostates – or Consociationalism

Given the ubiquity of ethnic conflict, and the fact that assimilation occurs, if at all, only gradually and, even then, only under certain conditions, a pessimist (or indeed a racial separatist) might conclude that the only way to prevent ethnic conflict is for different ethnic groups to be given separate territories with complete independence and territorial sovereignty. 

This would involve the partition of the world into separate ethnically homogenous ethnostates, as advocated by racial separatists and many in the alt-right. 

Yet, quite apart from the practical difficulties such an arrangement would entail, not least the need for large-scale forcible displacements of populations, this ‘universal nationalism’, as championed by political scientist Frank K Salter among others, would arguably only shift the locus of ethnic conflict from within the borders of a single multi-ethnic state to between those of separate ethnostates – and conflict between states can be just as destructive as conflict within states, as countless wars between states throughout history have amply proven.  

In the absence of assimilation, then, perhaps fairest and least conflictual solution is what van den Berghe terms consociationalism. This term refers to a form of ethnic power-sharing, whereby elites from both groups agree to share power, each usually retaining a veto power regarding major decisions, and there is proportionate representation for each group in all important positions of power. 

This seems to be roughly the basis of the power sharing agreement imposed on Northern Ireland in the Good Friday Agreement, which was largely successful in bringing an end to the ethnic conflict known as ‘the Troubles.[19]

On the other hand, however, power-sharing was explicitly rejected by both the ANC and the international anti-apartheid movement as a solution in another ethnically-divided polity, namely South Africa, in favour of majority rule, even though the result has been a situation very similar to the situation in Northern Ireland which led to the Troubles, namely an effective one-party state, with a single party in power for successive decades and institutionalized discrimination against minorities.[20]

Consociationalism or ethnic power-sharing also arguably the model towards which the USA and other western polities are increasingly moving, with quotas and so-called ‘affirmative action increasingly replacing the earlier ideals of appointment by merit, color blindness or freedom of association, and multiculturalism and cultural pluralism replacing the earlier ideal of assimilation

Perhaps the model consociationalist democracy is van den Berghe’s own native Belgium, where, he reports: 

All the linguistic, class, religious and party-political quarrels and street demonstrations have yet to produce a single fatality” (p199).[21]

Belgium is, however, very much the exception rather than the rule, and, at any rate, though peaceful, remains very much a divided society

Indeed, power-sharing institutions, in giving official, institutional recognition to the existing ethnic divide, function only to institutionalize and hence reinforce and ossify the existing ethnic divide, making successful integration and assimilation almost impossible – and certainly even less likely to occur than it had been in the absence of such institutional arrangements. 

Moreover, consociationalism can be maintained, van den Berghe emphasizes, only in a limited range of circumstances, the key criterion being that the groups in question are equal, or almost equal, to one another in status, and not organized into an ethnic hierarchy. 

However, even when the necessary conditions are met, it invariably involves a precarious balancing act. 

Just how precarious is illustrated by the fate of other formerly stable consociationalist states. Thus, van den Bergh notes the irony that earlier writers on the topic had cited Lebanon as “a model [consociationalist democracy] in the Third World” just a few years before the Lebanese Civil War broke out in the 1970s (p191). 

His point is, ironically, only strengthened by the fact that, in the three decades since his book was first published, two of his own examples of consociationalism, namely the USSR and Yugoslavia, have themselves since descended into civil war and fragmented along ethnic lines. 

Slavery and Other Recurrent Situations  

In the central section of the book, van den Berghe discusses such historically recurrent racial relationships as “slavery”, middleman minorities, “caste” and “colonialism”. 

In large part, his analyses of these institutions and phenomena do not depend on his sociobiological theory of ethnocentrism, and are worth reading even for readers unconvinced by this theory – or even by readers skeptical of sociobiology and evolutionary psychology altogether. 

Nevertheless, the sociobiological model continues to guide his analysis. 

Take, for example, his chapter on slavery. 

Although the overtly racial slavery of the New World was quite unique, slavery often has an ethnic dimension, since slaves are often captured during warfare from among enemy groups. 

Indeed, the very word slave is derived from the ethnonym, Slav, due to the frequency with which the latter were captured as slaves, both by Christians and Muslims.[22]

In particular, van den Berghe argues that: 

An essential feature of slave status is being torn out of one’s network of kin selection. This condition generally results from forcible removal of the slave from his home group by capture and purchase” (p120).

This then partly explains, for example, why European settlers were far less successful in enslaving the native inhabitants of the Americas than they were in exploiting the slave labour of African slaves who had been shipped across the Atlantic, far from their original kin groups, precisely for this purpose.[23]

Thus, for van den Berghe, the quintessential slave is: 

Not only involuntarily among ethnic strangers in a strange land: he is there alone, without his support group of kinsmen and fellow ethnics” (p115)

Here van den Berghe seemingly anticipates the key insight of Jamaican sociologist Orlando Peterson in his comparative study of slavery, Slavery and Social Death, who terms this key characteristic of slavery natal alienation.[24]

This, however, is likely to be only a temporary condition, since, at least if allowed to reproduce, then, gradually over time, slaves would put down roots, produce new families, and indeed whole communities of slaves.[25]

When this occurs, however, slaves gradually, over generations, cease to be true slaves. The result is that: 

Slavery can long endure as an institution in a given society, but the slave status of individuals is typically only semipermanent and nonhereditary… Unless a constantly renewed supply of slaves enters a society, slavery, as an institution, tends to disappear and transform itself into something else” (p120). 

This then explains the gradual transformation of slavery during the medieval period into serfdom in much of Europe, and perhaps also the emergence of some pariah castes such as the untouchables of India. 

Paradoxically, van den Berghe argues that racism became particularly virulent in the West precisely because of Western societies’ ostensible commitment to notions of liberty and the rights of man, notions obviously incompatible with slavery. 

Thus, whereas most civilizations simply took the institution of slavery for granted, feeling no especial need to justify its existence, western civilization, given its ostensible commitment to such lofty notions as individual liberty and the equality of man, was always on the defensive, feeling a constant need to justify and defend slavery. 

The main justification hit upon was racialism and theories of racial superiority

If it was immoral to enslave people, but if at the same time it was vastly profitable to do so, then a simple solution to the dilemma presented itself: slavery became acceptable if slaves could somehow be defined as somewhat less than fully human” (p115).  

This then explains much of the virulence of western racialism in the much of the eighteenth, nineteenth and even early-twentieth centuries.[26]

Another important, and related, ideological justification for slavery was what van den Berghe refers to as ‘paternalism’. Thus, Van den Berghe observes that: 

All chattel slave regimes developed a legitimating ideology of paternalism” (p131). 

Thus, in the American South, the “benevolent master” was portrayed a protective “father figure”, while slaves were portrayed as childlike and incapable of living an independent existence and hence as benefiting from their own enslavement (p131). 

This, of course, was a nonsense. As van den Berghe cynically observes: 

Where the parentage was fictive, so, we may assume, is the benevolence” (p131). 

Thus, exploitation was, in sociobiological terms, disguised as kin-selected parental benevolence

However, despite the dehumanization of slaves, the imbalance of power between slave and master, together with the men’s innate and evolved desire for promiscuity, made the sexual exploitation of female slaves by male masters all but inevitable.[27]

As van den Berghe observes: 

Even the strongest social barriers between social groups cannot block a specieswide [sic] sexual attraction. The biology of reproduction triumphs in the end over the artificial barriers of social prejudice” (p109). 

Thus, he notes the hypocrisy whereby: 

Dominant group men, whether racist or not, are seldom reluctant to maximize their fitness with subordinate-group women” (p33). 

The result was that the fictive ideology of ‘paternalism’ that served to justify slavery often gave way to literal paternity of the next generation of the slave population. 

This created two problems. First, it made the racial justification for slavery, namely the ostensible inferiority of black people, ring increasingly hollow, as ostensibly ‘black slaves acquired greater European ancestry, lighter skins and more Caucasoid features with each successive generation of miscegenation. 

Second, and more important, it also meant that the exploitation of this next generation of slaves by their owners potentially violated the logic of kin selection, because: 

If slaves become kinsmen, you cannot exploit them without indirectly exploiting yourself” (p134).[28]

This, van den Berghe surmises, led many slave owners to free those among the offspring of slave women whom they themselves, or their male relatives, had fathered. As evidence, he observes:  

In all [European colonial] slave regimes, there was a close association between manumission and European ancestry. In 1850 in the United States, for example, an estimated 37% of free ‘negroes’ had white ancestry, compared to about 10% of the slave population” (p132). 

This leads van den Bergh to conclude that many such free people of color – who were referred to as people of color precisely because their substantial degree of white ancestry precluded any simple identification as black or negro – had been freed by their owner precisely because their owner was now also their kinsmen. Indeed, many may have been freed by the very slave-master who had been responsible for fathering them. 

Thus, to give a famous example, Thomas Jefferson is thought to have fathered six offspring, four of whom survived to adulthood, with his slave, Sally Hemings – who was herself already three-quarters white, and indeed Jefferson’s wife’s own half-sister, on account of miscegenation in previous generations. 

Of these four surviving offspring, two were allowed to escape, probably with Jefferson’s tacit permission or at least acquiescence, while the remaining two were freed upon his death in his will.[29]

This seems to have been a common pattern. Thus, van den Berghe reports: 

Only about one tenth of the ‘negro’ population of the United States was free in 1860. A greatly disproportionate number of them were mulattoes, and, thus, presumably often blood relatives of the master who emancipated them or their ancestors. The only other slaves who were regularly were old people past productive and reproductive age, so as to avoid the cost of feeding the aged and infirm” (p129). 

Yet this made the continuance of slavery almost impossible, because each new generation more and more slaves would be freed.  

Other slave systems got around this problem by continually capturing or importing new slaves in order to replenish the slave population. However, this option was denied to American slaveholders by the abolition of the slave trade in 1807

Instead, the Americans were unique in attempting to ‘breed’ slaves. This leads van den Berghe to conclude that: 

By making the slave woman widely available to her master…Western slavery thus literally contained the genetic seeds of its own destruction” (p134).[30]

Synthesising Marxism and Sociobiology 

Given the potential appeal of his theory to nationalists, and even to racialists, it is perhaps surprising that van den Berghe draws heavily on Marxist theory. Although Marxists were almost unanimously hostile to sociobiology, sociobiologists frequently emphasized the potential compatibility of Marxist theory and sociobiology (e.g. The Evolution of Human Sociality). 

However, van den Berghe remains, to my knowledge, the only figure (except myself) to actually successfully synthesize sociobiology and Marxism in order to produce novel theory.  

Thus, for example, he argues that, in almost every society in existence, class exploitation is disguised by an ideology (in the Marxist sense) that disguises exploitation as either: 

1) Kin-selected nepotistic altruism – e.g. the king or dictator is portrayed as benevolent ‘father’ of the nation; or
2) Mutually beneficial reciprocity – i.e. social contract theory or democracy (p60). 

However, contrary to orthodox Marxist theory, van den Berghe regards ethnic sentiments as more fundamental than class loyalty since, whereas the latter is “dependent on a commonality of interests”, the former is often “irrational” (p243). 

Nationalist conflicts are among the most intractable and unamenable to reason and compromise… It seems a great many people care passionately whether they are ruled and exploited by members of their own ethny or foreigners” (p62). 

In short, van den Berghe concludes: 

Blood runs thicker than money” (p243). 

Another difference is that, whereas Marxists view control over the so-called means of production (i.e. the means necessary to produce goods for sale) as the ultimate factor determining exploitation and conflict in human societies, Darwinians instead focus on conflict over access to what I have termed the means of reproduction – in other words, the means necessary to produce offspring (i.e. fertile females, their wombs and vaginas etc.). 

This is because, from a Darwinian perspective: 

The ultimate measure of human success is not production but reproduction. Economic productivity and profit are means to reproductive ends, not ends in themselves” (p165). 

Thus, unlike his contemporary Darwin, Karl Marx was, for all his ostensible radicalism, in his emphasis on economics rather than sex, just another Victorian sexual prude.[31]

Mating, Miscegenation and Intermarriage 

Given that reproduction, not production, is the ultimate focus of individual and societal conflict and competition, van den Berghe argues that ultimately questions of equality, inequality and assimilation must be also determined by reproductive, not economic, criteria. 

Thus, he concludes, intermarriage, especially if it occurs, not only frequently, but also in both directions (i.e. involves both males and females of both ethnicities, rather than always involving males of one ethnic group, usually the dominant ethnic group, taking females of the other ethnic group, usually the subordinate group, as wives), is the ultimate measure of racial equality and assimilation: 

Marriage, especially if it happens in both directions, that is with both men and women of both groups marrying out, is probably the best measure of assimilation” (p218). 

In contrast, however, he also emphasizes that mere “concubinage is frequent [even] in the absence of assimilation” (p218). 

Moreover, such concubinage invariably involves males of the dominant-group taking females from the subordinate-group as concubines, whereas dominant-group females are invariably off-limits as sexual partners for subordinate group males. 

Thus, van den Berghe observes, although “dominant group men, whether racist or not, are seldom reluctant to maximize their fitness with subordinate-group women”, they nevertheless are jealously protective of their own women and enforce strict double-standards (p33). 

For example, historian Wynn Craig Wade, in his history of the Ku Klux Klan (which I have reviewed here), writes: 

In [antebellum] Southern white culture, the female was placed on a pedestal where she was inaccessible to blacks and a guarantee of purity of the white race. The black race, however, was completely vulnerable to miscegenation.” (The Fiery Cross: p20). 

The result, van den Berghe reports, is that: 

The subordinate group in an ethnic hierarchy invariably ‘loses’ more women to males of the dominant group than vice versa” (p75). 

Indeed, this same pattern is even apparent in the DNA of contemporary populations. Thus, geneticist James Watson reports that, whereas the mitochondrial DNA of contemporary Columbians, which is passed down the female line, shows a “range of Amerindian MtDNA types”, the Y-chromosomes of these same Colombians, are 94% European. This leads him to conclude: 

The virtual absence of Amerindian Y chromosome types, reveals the tragic story of colonial genocide: indigenous men were eliminated while local women were sexually ‘assimilated’ by the conquistadors” (DNA: The Secret of Life: p257). 

As van den Berghe himself observes: 

It is no accident that military conquest is so often accompanied by the killing, enslavement and castration of males, and the raping and capturing of females” (p75). 

This, of course, reflects the fact that, in Darwinian terms, the ultimate purpose of power is to maximize reproductive success

However, while the ethnic group as a whole inevitably suffers a diminution in its fitness, there is a decided gender imbalance in who bears the brunt of this loss. 

The men of the subordinate group are always the losers and therefore always have a reproductive interest in overthrowing the system. The women of the subordinate group, however frequently have the option of being reproductively successful with dominant-group males” (p27). 

Indeed, subordinate-group females are not only able, and sometimes forced, to mate with dominant-group males, but, in purely fitness terms, they may even benefit from such an arrangement.  

Hypergamy (mating upward for women) is a fitness enhancing strategy for women, and, therefore, subordinate-group women do not always resist being ‘taken over’ by dominant-group men” (p75). 

This is because, by so doing, they thereby obtain access to both the greater resources that dominant group males are able to provide in return for sexual access or as provisioning for their offspring, as well as the superior’ genes which facilitated the conquest in the first place. 

Thus, throughout history, women and girls have been altogether too willing to consort and intermarry with their conquerors. 

The result of this gender imbalance in the consequences of conquest and subjugation, is, a lack of solidarity as between men and women of the subjugated group. 

This sex asymmetry in fitness strategies in ethnically stratified societies often creates tension between the sexes within subordinate groups. The female option of fitness maximization through hypergamy is deeply resented by subordinate group males” (p76). 

Indeed, even captured females who were enslaved by their conquerers sometimes did surprisingly well out of this arrangement, at least if they were young and beautiful, and hence lucky enough to be recruited into the harem of a king, emperor or other powerful male.

One slave captured in Eastern Europe even went on to become effective queen of the Ottoman Empire at the height of its power. Hurrem Sultan, as she came to be known, was, of course, exceptional, but only in degree. Members of royal harems may have been secluded, but they also lived in some luxury.

Indeed, even in puritanical North America, where concubinage was very much frownded upon, van den Berghe reports that “slavery was much tougher on men than on women”, since: 

Slavery drastically reduced the fitness of male slaves; it had little or no such adverse effect on the fitness of female slaves whose masters had a double interest – financial and genetic – in having them reproduce at maximum capacity” (p133) 

Van den Berghe even tentatively ventures: 

It is perhaps not far-fetched to suggest that, even today, much of the ambivalence in relations between black men and women in America… has its roots in the highly asymmetrical mating system of the slave plantation” (p133).[32]

Miscegenation and Intermarriage in Modern America 

Yet, curiously, however, patterns of interracial dating in contemporary America are anomalous – at least if we believe the pervasive myth that America is a ‘systemically racist’ society where black people are still oppressed and discriminated against

On the one hand, genetic data confirms that, historically, matings between white men and black women were more frequent than the reverse, since African-American mitochondrial DNA, passed down the female line, is overwhelmingly African in origin, whereas their Y chromosomes, passed down the male line, are often European in origin (Lind et al 2007). 

However, recent census data suggests that this pattern is now reversed. Thus, black men are now about two and a half times as likely to marry white women as black women are to marry white men (Fryer 2007; see also Sailer 1997). 

This seemingly suggests white American males are actually losing out in reproductive competition to black males. 

This observation led controversial behavioural geneticist Glayde Whitney to claim: 

By many traditional anthropological criteria African-Americans are now one of the dominant social groups in America – at least they are dominant over whites. There is a tremendous and continuing transfer of property, land and women from the subordinate race to the dominant race” (Whitney 1999: p95). 

However, this conclusion is difficult to square with the continued disproportionate economic deprivation of much of black America. In short, African-Americans may be reproductively successful, and perhaps even, in some respects, socially privileged, but, despite benefiting from systematic discrimination in employment and admission to institutions of higher education, they are clearly also, on average, economically much worse-off as compared to whites and Asians in modern America.  

Instead, perhaps the beginnings of an explanation for this paradox can be sought in van den Berghe’s own later collaboration with anthropologist, and HBD blogger, Peter Frost

Here, in a co-authored paper, van den Berghe and Frost argue that, across cultures, there is a general sexual preference for females with somewhat lighter complexion than the group average (van den Berghe and Frost 1986). 

However, as Frost explains in a more recent work, Fair Women, Dark Men: The Forgotten Roots of Racial Prejudice, preferences with regard to male complexion are more ambivalent (see also Feinman & Gill 1977). 

Thus, whereas, according to the title of a novel, two films and a hit Broadway musical, ‘Gentlemen Prefer Blondes’ (who also reputedly, and perhaps as a consequence, have more fun), the idealized male romantic partner is instead tall, dark and handsome

In subsequent work, Frost argues that ecological conditions in sub-Saharan Africa permitted high levels of polygyny, because women were economically self-supporting, and this increased the intensity of selection for traits (e.g. increased muscularity, masculinity, athleticism and perhaps outgoing, sexually-aggressive personalities) which enhance the ability of African-descended males to compete for mates and attract females (Frost 2008). 

In contrast, Frost argues that there was greater selection for female attractiveness (and perhaps female chastity) in areas such as Northern Europe and Northeast Asia, where, to successfully reproduce, women were required to attract a male willing to provision them during cold winters throughout their gestation, lactation and beyond (Frost 2008). 

This then suggests that African males have simply evolved to be, on average, more attractive to women, whereas European and Asian females have evolved to be more attractive to men

This speculation is supported by a couple of recent studies of facial attractiveness, which found that black male faces were rated as most attractive to members of the opposite sex, but that, for female faces, the pattern was reversed (Lewis 2011; Lewis 2012). 

These findings could also go some way towards explaining patterns of interracial dating in the contemporary west (Lewis 2012). 

The Most Explosive Aspect of Interethnic Relations” 

However, such an explanation is likely to be popular neither with racialists, for whom miscegenation is anathema, nor with racial egalitarians, for whom, as a matter of sacrosanct dogma, all races must be equal in all things, even aesthetics and sex appeal.[33]

Thus, when evolutionary psychologist Satoshi Kanazawa made a similar claim in 2011 in a blog post (since deleted), outrage predictably ensued, the post was swiftly deleted, his then-blog dropped by its host, Psychology Today, and the author reprimanded by his employer, the London School of Economics, and forbidden from writing any blog or non-scholarly publications for a whole year. 

Yet all of this occurred within a year of the publication of the two papers cited above that largely corroborated Kanazawa’s finding (Lewis 2011; Lewis 2012). 

Yet such a reaction is, in fact, little surprise. As van den Berghe points out: 

It is no accident that the most explosive aspect of interethnic relations is sexual contact across ethnic (or racial) lines” (p75). 

After all, from a sociobiological perspective, competition over reproductive access to fertile females is Darwinian conflict in its most direct and primordial form

Van den Berghe’s claim that interethnic sexual contact is “the most explosive aspect” of interethnic relations also has support from the history of racial conflict in the USA and elsewhere. 

The spectre of interracial sexual contact, real or imagined, has motivated several of the most notorious racially-motivated ‘hate-crimes’ of American history, from the torture-murder of Emmett Till for allegedly propositioning a white woman, to the various atrocities of the reconstruction-era Ku Klux Klan in defence of the ostensible virtue of ‘white womanhood, to the recent Charleston church shooting, ostensibly committed in revenge for the allegedly disproportionate rate of rape of white women by black man.[34]

Meanwhile, interracial sexual relations are also implicated in some of American history’s most infamous alleged miscarriages of justice, from the Scottsboro Boys and Groveland Four cases, and the more recent Central Park jogger case, all of which involved allegations of interracial rape, to the comparatively trivial conduct alleged, but by no means trivial punishment imposed, in the so-called Monroe ‘kissing case

Allegations of interracial rape also seem to be the most common precursor of full-blown race riots

Thus, in early-twentieth century America, the race riots in Springfield, Illinois in 1908, in Omaha, Nebraska in 1919, in Tulsa, Oklahoma in 1921 and in Rosewood, Florida in 1923 were all ignited, at least in part, by allegations of interracial rape or sexual assault

Meanwhile, on the other side of the Atlantic, multi-racial Britain’s first modern post-war race riot, the 1958 Notting Hill riot in London 1958, began with a public argument between an interracial couple, when white passers-by joined in on the side of the white woman against her black Jamaican husband (and pimp) before turning on them both. 

Meanwhile, Britain’s most recent unambiguous race riot, the 2005 Birmingham riot, an entirely non-white affair, was ignited by the allegation that a black girl had been gang-raped by South Asians.

Meanwhile, at least in the west, whites no longer seem participate in race riots, save as victims. However, an exception was the 2005 Cronulla riots in Sydney, Australia, which were ignited by the allegation that Middle Eastern males were sexually harassing white Australian girls on Sydney beaches. 

Similarly, in Britain, though riots have yet to result, the spectre of so-called Muslim grooming gangs, preying on, and pimping out, underage white British girls in northern towns across the England, has arguably done more to ignite anti-Muslim sentiment among whites in the UK than a whole series of Jihadist terrorist attacks on British civilian targets

Thus, in Race: The Reality of Human Differences (which I have reviewed here, here and here) Sarich and Miele caution that miscegenation, often touted as the universal panacea to racism simply because, if practiced sufficiently widely, it would eventually eliminate all racial differences, or at least blur the lines between racial groups, may actually, at least in the short-term, actually incite racist attacks. 

This, they argue, is because: 

Viewed from the racial solidarist perspective, intermarriage is an act of race war. Every ovum that is impregnated by the sperm of a member of a different race is one less of that precious commodity to be impregnated by a member of its own race and thereby ensure its survival” (Race: The Reality of Human Differences: p256) 

This “racial solidarist perspective” is, of course, a crudely group selectionist view of Darwinian competition, and it leads Sarich and Miele to hypothesize: 

Paradoxically, intermarriage, particularly of females of the majority group with males of a minority group, is the factor most likely to cause some extremist terrorist group to feel the need to launch such an attack” (Race: The Reality of Human Differences: p255). 

In other words, in sociobiological terms, ‘Robert’, a character from one of Michel Houellebecq’s novels, has it right when he claims: 

What is really at stake in racial struggles… is neither economic nor cultural, it is brutal and biological: It is competition for the cunts of young women” (Platform: p82). 

Endnotes

[1] Admittedly, the Croatian War of Independence is indeed sometimes said to have been triggered, or at least precipitated, by a football match between Dinamo Zagreb and Red Star Belgrade, and the riot that occurred at the ground on that day. However, this war was, of course, ethnic in origin, fought between Croats and Serbians, and the football match served as a triggering event only because the two teams were overwhelmingly supported supported by Croats and Serbians respectively.
This leads to an interesting observation – namely that rivalries such as those between supporters of different football teams tend to become especially malignant and acrimonious when support for one team or the other comes to be inextricably linked to ethnic identity.
Thus it is surely no accident that, in the UK, the most intense rivalry between groups of football supporters is that between between supporters of Ragners and Celtic in Glasgow, at least in part because the rivalry has become linked to religion, which was, at least until recently, a marker for ancestry and ethnicity, while an apparently even more intense rivalry was that between Linfield and Belfast Celtic in Northern Ireland, which was also based on a parallel religious and ethnic divide, and ultimately became so acrimonious that one of the two teams had to withdraw from domestic football and ultimately ceased to exist.

[2] Actually, however, contrary to Brigandt’s critique, it is clear that van den Berghe intended his “biological golden rule” only as a catchy and memorable aphorism, crudely summarizing Hamilton’s rule, rather than a quantitative scientific law akin to, or rivalling, Hamilton’s Rule itself. Therefore, this aspect of Brigandt’s critique is, in my view, misplaced. Indeed, it is difficult to see how this supposed rule could be applied as a quantitative scientific law, since relatedness, on the one hand, and altruism, on the other, are measured in different currencies. 

[3] Thus, van den Berghe concedes that: 

In many cases, the common descent acribed to an ethny is fictive. In fact, in most cases, it is partly fictive” (p27). 

[4] The question of racial nationalism (i.e. encompassing all members of a given race, not just those of a single ethnicity or language group) is actually more complex. Certainly, members of the same race do indeed share some degree of kinship, in so far as they are indeed (almost by definition) on average more closely biologically related to one another than to members of other races – and indeed that relatedness is obviously apparent in their phenotypic resemblance to one another. This suggests that racial nationalist movements such as that of, say, UNIA or of the Japanese imperialists, might have more potential as a viable form of nationalism than do attempts to unite racially disparate ethnicities, such as civic nationalism in the contemporary USA. The same may also be true of Oswald Mosley’s Europe a Nation campaign, at least while Europe remained primarily monoracial (i.e. white). However, any such racial nationalism would incorporate a far larger and more culturally, linguistically and genetically disparate group than any form of nationalism that has previously proven capable of mobilizing support.
Thus, Marcus Garvey’s attempt to create a kind of pan-African ethnic identity enjoyed little success and was largely restricted to North America, where African-Americans, do indeed share a common language and culture in addition to their race. Similarly, the efforts of Japanese nationalists to mobilize a kind of pan-Asian nationalism in support of their imperial aspirations during the first half of the twentieth century was an unmitigated failure, though this was partly because of the brutality with which they conquered and suppressed the other Asian nationalities whose support for pan-Asianism they intermittently and half-heartedly sought to enlist.
On the other hand, it is sometimes suggested that, in the early twentieth century, a white supremacist ideology was largely taken for granted among whites. However, while to some extent true, this shared ideology of white supremacism did not prevent the untold devastation wrought by the European wars of the early twentieth century, namely World Wars I and II, which Patrick Buchanan has collectively termed The Great Civil War of the West.
Thus, European nationalisms usually defined themselves by opposition to other European peoples and powers. Thus, just as Irish nationalism is defined largely by opposition to Britain, and Scottish nationalism by opposition to England, so English (and British) nationalism has itself traditionally been directed against rival European powers such as France and Germany (and formerly Spain), while French nationalism seems to have defined itself primarily in opposition to the Germans and the British, and German nationalism in opposition to the French and Slavs, etc.
It is true that, in the USA, a kind of pan-white American nationalism did seem to prevail in the early twentieth century, albeit initially limited to white protestants, and excluding at least some recent European immigrants (e.g. Italians, Jews). This is, however, a consequence of the so-called melting pot, and really only amounts to yet another parochial nationalism, namely that of a newly-formed ethnic group – white Americans.
At any rate, today white American nationalism is, at most, decidedly muted in form – a kind of implicit white racial consciousness, or, to coin a phrase, the nationalism that dare not speak its name. Thus, Van den Berghe observes: 

In the United States, the whites are an overwhelming majority, so much so that they cannot be meaningfully conceived of as a ruling group at all. The label ‘white’ in the United States does not correspond to a well-defined ethnic or racial group with a high degree of social organization or even self-consciousness, except regionally in the south” (p183). 

Van den Berghe wrote this in 1981. Today, of course, whites are no longer such an “overwhelming majority” of the US population. On the contrary, they are already well on the way to becoming a minority in America, a milestone that is likely to be reached over the coming decades.
Yet, curiously, white ‘racially consciousness’ is seemingly even more muted and implicit today than it was back when van den Berghe authored his book – and this is seen even in the South, which van den Berghe cited as an exception and lone bastion of white identity politics.
True, White Southerners may vote as a solidly for Republican candidates as they once did for the Democrats. However, overt appeals to white racial interests are now as anathema in the South as elsewhere.
Thus, as recently as 1990, a more or less open white racialist like David Duke was able to win a majority of the white vote in Louisiana in his run for the Senate. Today, this is unimaginable.
If the reason that whites lack any ‘racial consciousness’ is indeed, as van den Berghe claims, because they represent such an “overwhelming majority” of the American population, then it is interesting to speculate if and when, during the ongoing process of white demographic displacement, this will cease to be the case.
One thing seems certain: If and when it does ever occur, it will be too late to make any difference to the ongoing process of demographic displacement that some have termed ‘The Great Replacement’ or a third demographic transition.

[5] Of course, a preference for those who look similar to oneself (or one’s other relatives) may itself function as a form of kin recognition (i.e. of recognizing who is kin and who is not). This is referred to in biology as phenotype matching. Moreover, as Richard Dawkins has speculated in The Selfish Gene (reviewed here), racial feeling could conceivably have evolved through a misfiring of such a crude heuristic (The Selfish Gene: p100).

[6] Actually, I suspect that, on average, at least historically, both mothers and fathers may indeed, on average, have provided rather less care for their mixed-race offspring than for offspring of the same race as themselves, simply because mixed-race offspring were more likely to be born out of wedlock, not least because interracial marriage was, until recently, strongly frowned upon, and, in some jurisdictions, either not legally permitted or even outright criminalized, and both mothers and fathers tended to provide less care for illegitimate offspring, fathers because they often refused to acknowledge their illegitimate offspring and had little or no contact with them and may not even have been aware of their existence, and mothers because, lacking paternal support, they usually had no means of raising their illegitimate offspring alone and hence often gave them up for adoption or fostering.

[7] On the other hand, in his paper, An integrated evolutionary perspective on ethnicity, controversial antiSemitic evolutionary psychologist Kevin Macdonald disagrees with this conclusion, citing personal communication from geneticist and anthropologist Henry Harpending for the argument that: 

Long distance migrations have easily occurred on foot and over several generations, bringing people who look different for genetic reasons into contact with each other. Examples include the Bantu in South Africa living close to the Khoisans, or the pygmies living close to non-pygmies. The various groups in Rwanda and Burundi look quite different and came into contact with each other on foot. Harpending notes that it is ‘very likely’ that such encounters between peoples who look different for genetic reasons have been common for the last 40,000 years of human history; the view that humans were mostly sessile and living at a static carrying capacity is contradicted by history and by archaeology. Harpending points instead to ‘starbursts of population expansion’. For example, the Inuits settled in the arctic and exterminated the Dorsets within a few hundred years; the Bantu expansion into central and southern Africa happened in a millennium or less, prior to which Africa was mostly the yellow (i.e., Khoisan) continent, not the black continent. Other examples include the Han expansion in China, the Numic expansion in northern America, the Zulu expansion in southern Africa during the last few centuries, and the present day expansion of the Yanomamo in South America. There has also been a long history of invasions of Europe from the east. ‘In the starburst world people would have had plenty of contact with very different looking people’” (Macdonald 2001: p70). 

[8] Others have argued that the differences between Tutsi and Hutu are indeed largely a western creation, part of the divide and rule strategy supposedly deliberately employed by European colonialists, as well as a theory of Tutsi racial superiority promulgated by European racial anthropologists known as the Hamitic theory of Tutsi origins, which suggested that the Tutsi had migrated from the Horn of Africa, and had benefited from Caucasoid ancestry, as reflected in their supposed physiological differences from the indigenous Hutu (e.g. lighter complexions, greater height, narrower noses).
On this view, the distinction between Hutu and Tutsi was originally primarily socioeconomic rather than racial, and, at least formerly, the boundaries between the two groups were quite fluid.
I suspect this view is nonsense, reflecting political correctness and the leftist tendency to excuse any evidence of dysfunction or oppression in non-Western cultures as necessarily of product of the malign influence of western colonizers. (Most preposterously, even the Indian caste system has been blamed on British colonizers, although it actually predated them, in one form or another, by several thousand years.)
With respect to the division between Tutsi and Hutu, there are not only morphological differences between the two groups in average stature, nose width and complexion, but also substantial differences in the prevalence of genes for lactose tolerance and sickle-cell. These results do indeed seem to suggest that, as predicted by the reviled ‘Hamitic theory’, the Tutsi do indeed have affinities with populations from the Horn of Africa and East Africa. Modern genome analysis tends to confirm this conclusion. 

[9] Exceptions, where immigrant groups retain their distinctive language for multiple generations, occur where immigrants speaking a particular language arrive in sufficient numbers, and are sufficiently isolated in ethnic enclaves and ghettos, that they mix primarily or exclusively with people speaking the same language as themselves. A related exception is in respect of economically, politically or socially dominant minorities, such as alien colonizers, as well as market-dominant or middleman minorities, who often resist assimilation into the mainstream culture precisely so as to maintain their cultural separateness and hence their privileged position within society, and who also, partly for this reason, take steps to socialize, and ensure their offspring socialize, primarily among their own group. 

[10] Some German-Americans were also interred during World War II. However, far fewer were interred than among Japanese-Americans, especially on a per capita basis.
Nevertheless, some German-Americans were treated very badly indeed, yet the latter, unlike the Japanese, have yet to receive a government apology or compensation. Moreover, there was perhaps justification for the differing treatment accorded Japanese- and German-Americans, since the latter were generally longer established and, being white, were also more successfully integrated into mainstream American society, and there was perceived to be a real threat of enemy sabotage.
Also, with regard to van den Berghe’s observation that nuclear atomic weapons were used only against Japan, this is rather misleading. Nuclear weapons could not have been used against Germany, since, by the time of the first test detonation of a nuclear device, Germany had already surrendered. Yet, in fact, the Manhattan Project seems to have been begun with the Germans very much in mind as a prospective target. (Many of the scientists involved were Jewish, many having fled Nazi-occupied Europe for America, and hence their hostility towards the Nazis, and perhaps Germans in general, is easy to understand.)
Whether it is true that, as van den Berghe claims, atomic bombs were never actually likely to be “dropped over, say, Stuttgart or Dortmund” is a matter of supposition. Certainly, there were great animosity towards the Germans in America, as illustrated by the Morgenthau Plan, which, although ultimately never put into practice, was initially highly influential in directing US policy in Europe and even supported by President Roosevelt.
On the other hand, Roosevelt’s references to ‘the Nazis, the Fascists, and the Japanese’ might simply reflect the fact that there was no obvious name for the faction or regime in control of Japan during the Second World War, since, unlike in Germany and Italy, no named political party had seized power. I am therefore unconvinced that a great deal can necessarily be read into this.

[11] This was especially so in historical times, before the development of improved technologies of long-distance transportation (ships, aeroplanes) enabled more distantly related populations to come into contact, and hence conflict with one another (e.g. blacks and whites in the USA and South Africa, South Asians and English in the UK or under the British Raj). Thus, the ancient Indian treatise on statecraft and strategy, Arthashastra, observed that a ruler’s natural enemies are his immediate neighbours, whereas his next-but-one neighbours, being immediate neighbours of his own immediate neighbours, are his natural allies. This is sometimes credited as the origin of the famous aphorism, The enemy of my enemy is my friend.

[12]  The idea that neighbouring groups tend to be in conflict with one another precisely because, being neighbours, they are also in close contact, and hence competition, with one another, ironically posits almost the exact opposite relationship between ‘contact’ and intergroup relations than that posited by the famous contact theory of mid-twentieth psychology, which posited that increased contact between members of different racial and ethnic groups would lead to reduced prejudice and animosity.
This, of course, depends, at least partly, on the nature of the ‘contact’ in question. Contact that involves territorial rivalry, economic competition and war, obviously exacerbates conflict and animosity. In contrast, proponents of contact theory typically had in mind personal contact, rather than, say, the sort of impersonal, but often deadly, contact that occurs between rival belligerent combatants in wartime.
In fact, however, even at the personal level, contact can take many different forms, and often functions to increase inter-ethnic animosity. Hence the famous proverb, ‘familiarity breeds contempt’.
Indeed, social psychologists now concede that only ‘positive’ interactions with members with members of other groups (e.g. friendship, cooperation, acts of altruism, mutually beneficial trade) reduces animosity and conflict.
In contrast, negative interactions (e.g. being robbed, mugged or attacked by members of another group) only serves to reinforce, exacerbate, or indeed create intergroup animosity. This, of course, reduces the contact hypothesis to little more than common sense – positive experiences with a given group lead to positive perceptions of that group; negative interactions to negative perceptions.
This in turn suggests that stereotypes are often based on real experiences and therefore tend to be true – if not of all individuals, then at least at the statistical, aggregate group level.
I would add that, anecdotally, even positive interactions with members of disdained outgroups do not always shift perceptions regarded the disdained outgroup as a whole. Instead, the individuals with whom one enjoys positive interactions, and even friendships, are often seen as exceptions to the rule (‘one of the good ones’), rather than representative of the demographic to which they belong. Hence the familiar phenomenon of even virulent racists having friendships and sometimes even heroes among members of races whom they generally otherwise disdain. 

[13] However, Van den Berghe acknowledges that racially diverse societies have lived in “relative harmony” in places such as Latin America, where government gives no formal political recognition to racial groups (e.g. racial preferences and quotas for members of certain races) and where the latter do not organize on a racial basis, such that government is, in van den Berghe’s terminology, “non-racial” rather than “multiracial” (p190). However, this is perhaps a naïvely benign view of race relations in Latin American countries such as Brazil, which is, despite the fluidity of racial identity and lack of clear dividing lines between races, nevertheless now viewed by most social scientists, not so much as a model racial democracy, so much as a racially-stratified pigmentocracy , where skin tone correlates with social status. It is also arguably an outdated view of race relations in Latin America, because, perhaps due to indirect cultural and political influence emanating from the USA, ethnic groups in much of Latin America (e.g. blacks in Brazil, indigenous populations in Bolivia) increasingly do organize and agitate on a racial basis.

[14] I am careful here not to refer to refer the dominant culture as that of either a ‘host population’ or a ‘majority population’, or the subordinate group as a ‘minority group’ or an incoming group of migrants. This is because sometimes newly-arrived settlers successfully assimilate the indigenous populations among whom they settle, and sometimes it is the majority group who ultimately assimilate to the norms and culture of the minority. Thus, for example, the Anglo-Saxons imposed their Germanic language on the indigenous inhabitants of what is today England, and indeed ultimately most of the inhabitants of Scotland, Wales and Ireland as well, even though they likely never represented a majority of the population even in England, and may have made only a comparatively modest contribution to the ancestry of the people whom we today call ‘English’.

[15] Interestingly, and no doubt controversially, Van den Berghe argues that blacks in the USA do not have any distinctive cultural traits that distinguish them from the white American mainstream, and that their successful assimilation has been prevented only by the fact that, until very recently, whites have refused to ‘assimilate’ them. He is particularly skeptical regarding the notion of any cultural inheritances from Africa, dismissing “the romantic search for survivals of African Culture” as “elusive” (p177).
Indeed, for van den Berghe, the whole notion of a distinct African-American culture is “largely ideological and romantic” (p177). “Afro-Americans are,” he argues, “culturally ‘Anglo-Saxon’” and hence paradoxically ”as Anglo as anyone… in America” (p177). He concludes:

The case for ‘black culture’ rests… largely on the northern ghetto lumpenproletariat, a class which has no direct counterpart. Even in that group, however, much of the distinctiveness is traceable to their southern, rural origins” (p177). 

This reference to “southern rural origins” anticipates Thomas Sowell’s later black redneck hypothesis. Certainly, many aspects of black culture, such as dialect (e.g. the use of terms such as y’all and ain’t and the pronunciation of ‘whores’ as ‘hoes’) and stereotypical fondness for fried chicken, are obvious inheritances from Southern culture rather than distinctively black, let alone an inheritance from Africa. Thus, van den Berghe observes:

Ghetto lumpenproletariat blacks in Chicago, Detroit and New York may seem to have a distinct subculture of their own compared collectively to their white neighbors, but the black Mississippi sharecropper is not very different, except for his skin pigment, from his white counterparts” (p177). 

Any remaining differences not attributable to their Southern origins are, van den Berghe claims, not “African survivals, but adaptation to stigma” (p177). Here, van den Berghe perhaps has in mind the inverse morality, celebration of criminality, and bad nigger’ archetype prevalent in, for example, gangsta rap music. Thus, van den Berghe concludes that: 

Afro-Americans owe their distinctiveness overwhelmingly to the fact that they have been first enslaved and then stigmatized as a pariah group. They lack a territorial base, the necessary economic, and political resources and the cultural and linguistic pluralism ever to constitute a successful nation. Their pluralism is strictly a structural pluralism inflicted on them by racism. A stigma is hardly an adequate basis for successful nationalism” (p184). 

[16] Thus, Elizabeth Warren was a law professor who became a Democratic Party Senator and Presidential candidate, and had described herself as ‘American Indian, and been cited by her University employers as an ethnic minority, in order to benefit from informal affirmative action, despite having only a very small amount of Native American ancestry. Krug and Dolezal, meanwhile, taking advantage of the one drop rule, both identified as African-American, Krug, a history professor and leftist activist, taking advantage of her Middle-Eastern appearance, itself likely a reflection of her Jewish ancestry. Dolezal, however, was formerly a white, blonde girl, but, through the simple expedient of getting a perm and tan, managed to become an adjunct professor of black studies at a local university and local chapter president of the NAACP in an overwhelmingly white town and state. Whoever said blondes have more fun? 

[17] It has even given rise to a popular new hairstyle among young white males attempting to escape the stigma of whiteness by adopting a racially ambiguous appearance – the mulatto perm

[18] Interestingly, the examples cited by Paddy Hannam in his piece on the phenomenon, The rise of the race fakers also seem to have been female (Hannam 2021). Steve Sailer wisely counsels caution with regard to the findings of this study, noting that anyone willing to lie about their ethnicity on their college application, is also likely even more willing to lie in an anonymous survey (Sailer 2021 ; see also Hood 2007). 

[19] Actually, the Northern Ireland settlement is often classed as centripetalist rather than consociationalist. However, the distinction is minimal, with the former arrangement representing a modification of the latter designed to encourage cross-community cooperation, and prevent, or at least mitigate, the institutionalization and ossification of the ethnic divide that is perceived to occur under consociationalism, where constitutional recognition is accorded to the divide between the two (or more) communities. There is, however, little evidence that centripetalism have ever actually been successful in encouraging cross-community cooperation, beyond what is necessitated by the consitutional system, let alone encouraging assimilation of the rival communities and the depoliticization of ethnic identity. 

[20] The reason for the difference in the attitudes of leftists and liberals towards majority-rule in Northern Ireland and South Africa respectively seems to reflect the fact that, whereas in Northern Ireland, the majority protestant population were perceived of as the dominant oppressor’ group, the black majority in South Africa were perceived of as oppressed.
However, it is hard to see why this would mean black majority-rule in South Africa would be any less oppressive of South Africa’s white, coloured, and Asian minorities than Protestant majority rule had been of Catholics in Ulster. On the contrary, precisely because the black majority in South Africa perceive themselves as having been ‘oppressed’ in the past, they are likely to be especially vengeful and feel justified in seeking recompense for their earlier perceived oppression. This indeed seems to be what is occurring in South Africa, and Zimbabwe, today. 
Interestingly, van den Berghe, writing in 1981 was wisely prophetic regarding the long-term prospects for both apartheid – and for white South Africans. Thus, on the one hand he predicted: 

Past experience with decolonization elsewhere in Africa, especially in Zimbabwe (which is in almost every respect a miniature version of South Africa) seems to indicate that the end of white domination is in sight. The only question is whether it will take the form of a prolonged civil war, a negotiated partition or a frantic white exodus. The odds favor, I think, a long escalating war of attrition accompanied by a gradual economic winddown and a growing white emigration” (p174). 

Thus, van den Berghe was right in so far as he predicted the looming end of the apartheid system – though hardly unique in making this prediction. However, he was wrong in his predictions as to how this end would come about. On the other hand, however, with ongoing farm murders and the overtly genocidal rhetoric of populist politicians like Julius Malema, van den Berghe was probably right regarding the long-term prognosis of the white community in South Africa when he observed: 

Five million whites perched precariously at the tip of a continent inhabited by 400 millions blacks, with no friends in sight. No matter what happens whites will lose heavily, perhaps their very lives, or at least their place in the African sun that they love so much” (p172). 

However, perhaps surprisingly, van den Berghe denies that apartheid was entirely a failure: 

Although apartheid failed in the end, it was a rational course for the Afrikaners to take, given their collective aims, and probably did postpone the day of reckoning by about 30 years” (p174).

[21] The only other polity that perhaps has a competing claim to representing the world’s model consociationalist democracy is Switzerland. However, van den Berghe emphasizes that Switzerland is very much a special case, the secret of its success being that:

Switzerland is one of those rare multiethnic states that did not originate either in conquest or in the breakdown of multinational empires” (p194).

It managed to avoid conquest by its richer and more powerful neighbours simply because:

The Swiss had the dual advantage in resisting outside conquest: favorable terrain and lack of natural resources” (p194)

Also, it provided valuable services to these neighbours, first providing mercenaries to fight in their armed forces and later specialising in the manufacture of watches and what van den Berghe terms “the management of shady foreigners’ ill-gotten capital” (p194).
In reality, however, although divided linguistically and religiously, Switzerland does not, in van den Berghe’s constitute true consociationalism, since the country, with originated as confederation of fomerly independent hill tribes, remains highly decentralized, and power is shared, not by ethnic groups, but rather between regional cantons. Therefore, van den Berghe concludes:

The ethnic diversity of Switzerland is only incidental to the federalism, it does not constitute the basis for it” (p196-7).

In addition, most cantons, where much of the real power lies, are themselves relatively monoethnic and monoliguistic, at least as compared to the country as a whole.

[22] Indeed, since the Slavs of Eastern Europe were the last group in Europe to be converted to Christianity, and it was forbidden by Papal decree to enslave fellow-Christians or sell Christian slaves to non-Christians (i.e. Muslims, among whom there was a great demand for European slaves), Slavs were preferentially targeted by Christians for enslavement, and even those non-Slavic people who were enslaved or sold into bondage were often falsely described as Slavs in order to justify their enslavement and sale to Muslim slaveholders. The Slavs, for geographic reasons, were also vulnerable to capture and enslavement directly by the Muslims themselves.

[23] Another reason that it proved difficult to enslave the indigenous inhabitants of the Americas, according to van den Berghe, is the lifestyle of the latter prior to colonization. Thus, prior to the arrival of Euopean colonists, the indigenous people in many parts of the Americas were still relatively primitive, many subsisting, in whole or in part, as nomadic or semi-nomadic hunter-gatherers. This meant, not only that they had low population densities and were hence few in number and vulnerable to infectious diseases introduced by European colonizers, but also that:

Such aborigines as existed were mobile, elusive and difficult to control. They typically had a vast hinterland into which they could escape labor exploitation” (p93).

Thus, van den Berghe reports, when, in what is today Brazil, Portuguese colonists led raiding expeditions in an attempt to capture and enslave natives, so many of the latter “escaped, committed suicide or died of disease” that the attempt was soon abandoned (p93).
Perhaps more interestingly, van den Berghe also argues that another reason that it proved difficult to enslave nomadic peoples was that:

Nomads typically are unused to being exploited since their own societies are often relatively egalitarian, ill-adapted to steady hard labor and lacking in the skills useful to colonial exploiters (as cultivators, for example). They are, in short, lovers of freedom and make very poor colonial underlings… They are regarded by their conquerors as lazy, shiftless and unreliable, as an obstacle to development and as a nuisance to be displaced” (p93).

In contrast, whereas sub-Saharan Africa is usually stereotyped, not entirely inaccurately, as technologically backward as compared to other cultures, and this very backwardness as facilitating their enslavement, in fact, van den Berghe explains, it was the relatively socially advanced nature of West African societies that permitted the transatlantic slave trade to be so successful.

Contrary to general opinion, Africans were so successfully enslaved, not because they belonged to primitive cultures, but because they had a complex enough technology and social organization to sustain heavy losses of manpower without appreciable depopulation. Even the heavy slaving of the 18th century made only a slight impact on the demography of West Africa. The most heavily raided areas are still today among the most densely populated” (p126).

[24] Although this review is based on the 1987 edition, The Ethnic Phenomenon was first published in 1981, whereas Orlando Peterson’s Slavery and Social Death came out just a year later in 1982.

[25] In the antebellum American South, much is made of the practice of slave-owners selling the spouses and offspring of their slaves to other masters, thereby breaking up families. On the basis of van den Berghe’s arguments, this might actually have represented an effective means of preventing slaves from putting down roots and developing families and slave communities, and might therefore have helped perpetuate the institution of slavery.
However, even assuming that such practices would indeed have had this effect, it is doubtful that there was any such deliberate long-term policy among slaveholders to break up families in this way. On the contrary, van den Berghe reports:

It is not true that slave owners systematically broke up slave couples… On the contrary, it was in their interest to foster stable slave families for the sake of morale, and to discourage escape” (p133). 

Thus, though it certainly occurred and may indeed have been tragic where it did occur, slaveholders generally preferred to keep slave families intact, precisely because, in forming families, slaves would indeed ‘put down roots’ and hence be less likely to try to escape, lest, in the process, they would leave other family members behind to face the vengeance of their former owners alone and without any protection and support they might otherwise have been in a position to offer. The threat of breaking up families, however, surely remained a useful tool in the arsenal of slaveholders to maintain control over slaves. 

[26] While acknowledging, and indeed emphasizing, the virulence of western racialism, van den Berghe, bemoaning the intrusion of “moralism” (and, by extension, ethnomasochism) into scholarship, has little time for the notion that western slavery was intrinsically more malign than forms of slavery practised in other parts of the world or at other times in history (p116). This, he dismisses as “the guilt ascription game: whose slavery was worse?” (p128). Male slaves in the Islamic world, for example, were routinely castrated before being sold (p117). 
Thus, while it is true that slaves in the American South had unusually low rates of manumission (i.e. the granting of freedom to slaves), they also enjoyed surprisingly high standards of living, were well-fed and enjoyed long lives. Indeed, not only did slaves in the American South enjoy standards of living superior to those of most other slave populations, they even enjoyed, by some measures, higher standards of living than many non-slave populations, including industrial workers in Europe and the Northern United States, and poor white Southerners, during the same time period (The End of Racism: p88-91; see also Time on the Cross: the Economics of American Slavery). 
Ironically, living standards were so high for the very same reason that rates of manumission were so low – namely, slaves, especially after the abolition and suppression of the transatlantic slave-trade (but also even before then due to the costs of transportation during the middle passage) were an expensive commodity. Masters therefore fully intended to get their money’s worth out of their slaves, not only by rarely granting them their freedom, but also ensuring that they lived a long and healthy life.
In this endeavour, they were surprisingly successful. Thus, van den Berghe reports, in the fifty years that followed the prohibition on the import of new slaves into the USA in 1908, the black population of the USA nevertheless more than tripled (p128). In short, slaves may have been property, but they were valuable property – and slaveholders made every effort to protect their investment.
Ironically, therefore, indentured servants (themselves, in America, often white, and later, in Africa, usually South or East Asian) were, during the period of their indenture, often worked harder, and forced to live in worse conditions, than were actual slaves. This was because, since they were indentured for only a set number of years before they would be free, there was less incentive on the part of their owners to ensure that they lived a long and healthy life.   
Van den Berghe concludes: 

“The blanket ascription of collective racial guilt for slavery to ‘whites’ that is so dear to many liberal social scientists is itself a product of the racist mentality produced by slavery. It takes a racist to ascribe causality and guilt to racial categories” (p130). 

Indeed, as Dinesh D’Souza in The End of Racism, and Thomas Sowell in his essay ‘The Real History of Slavery’ included in the collection Black Rednecks and White Liberals, both emphasize, whereas all civilizations have practised slavery, what was unique about western civilization was that it was the first civilization ever known to have abolished slavery (at, as it ultimately turned out, no little economic cost to itself).
Therefore, even if liberals and leftists do insist that we play what van den Berghe disparagingly calls “the guilt ascription game”, then white westerners actually come out rather well in the comparison. 

[27] Indeed, in most cultures and throughout most of history, the use of female slaves as concubines was, not only widespread, but also perfectly socially acceptable. For example, in the Islamic world, the use of female slaves as concubines was entirely open and accepted, not only attracting literally no censure or criticism in the wider society or culture, but also receiving explicit prophetic sanction in the Quran. For this reason, in the Islamic world, females slaves tended to be in greater demand than males, and usually commanded a higher price.
In contrast, most slaves transported to the Americas were male, since males were more useful for hard, intensive agricultural labour and, in puritanical North America, sexual contact with between slaveholder and slave was very much frowned upon, even though it certainly occurred. Thus, van den Berghe cynically observes:  

Concubinage with slaves was somewhat more clandestine and hypocritical in the English and Dutch colonies than in the Spanish, Portuguese and French colonies where it was brazen, but there is no evidence that the actual incidence of interbreeding was any higher in the Catholic countries” (p132). 

Partial corroboration for this claim is provided by historian Eugene Genovese, who, in his book Roll, Jordan, Roll: The World the Slaves Made, reports that, in New Orleans slave markets:

First-class blacksmiths were being sold for $2,500 and prime field hands for about $1,800, but a particularly beautiful girl or young woman might bring $5,000” (Roll, Jordan, Roll: p416).

[28] Actually, exploitation can still be an adaptive strategy, even in respect of close biological relatives. This depends of the precise relative gain and loss in fitness to both the exploiter (the slave owner) and his victim (the slave), and their respective coefficient of relatedness, in accordance with Hamilton’s rule. Thus, it is possible that a slaveholder’s genes may benefit more from continuing to exploit his slaves as slaves than by freeing them, even if the latter are also his kin. Possibly the best strategy will often be a compromise of, say, keeping your slave-kin in bondage, but treating them rather better than other non-related slaves, or freeing them after your death in your will. 
Of course, this is not to suggest that individual slaveholders consciously (or subconsciously) perform such a calculation, nor even that their actual behaviour is usually adaptive. Slaveholding is likely an ‘environmental novelty’ to which we are yet to have evolved adaptive responses

[29] Others suggest that Thomas Jefferson himself did not father any offspring with Sally Hemmings and that the more likely father is Jefferson’s wayward younger brother Randolph, who would, of course, share the same Y chromosome as his elder brother. For present purposes, this is not especially important, since, either way, Heming’s offspring would be blood relatives of Jefferson to some degree, hence likely influencing his decision to free them or permit them to escape.

[30] Quite how this destruction can be expected to have manifested itself is not spelt out by van den Berghe. Perhaps, with each passing generation, as slaves became more and more closely biologically related to their masters, more and more slaves would have been freed until there were simply no more left. Alternatively, perhaps, as slaves and slaveowners increasingly became biological kin to one another, the institution of slavery would gradually have become less oppressive and exploitative until ultimately it ceased to constitute true slavery at all. At any rate, in the Southern United States this (supposed) process was forestalled by the American Civil War and Emancipation Proclamation, and neither does it appear to have occurred in Latin America.  

[31] Another area of conflict between Marxism and Darwinism is the assumption of the former that somehow all conflict and exploitation will end in a future posited communist utopia. Curiously, although healthily cynical about exploitation under Soviet-style communism (p60), van den Berghe describes himself as an anarchist (van den Berghe 2005). However, anarchism seems even more hopelessly utopian than communism, given humanity’s innate sociality and desire to exploit reproductive competitors. In short, a Hobbesian state of nature is surely no one’s utopia (except perhaps Ragnar Redbeard). 

[32] The idea that there is “ambivalence in relations between black men and women in America” seems anecdotally plausible, given, for example, the delightfully misogynistic lyrics found in much African-American rap music. However, it is difficult to see how this could be a legacy of the plantation era, when everyone alive today is several generations removed from that era and living in a very different sexual and racial milieu. Today, black men do rather better in the mating market place than do black women, with black men being much more likely to marry non-black women than black women are to marry non-black men, suggesting that black men have a larger dating pool from which to choose (Sailer 1997; Fryer 2007).
Moreover, black men and women in America today are, of course, the descendants of both men and women. Therefore, even if black women did have a better time of it that black men in the plantation era, how would black male resentment be passed down the generations to black men today, especially given that most black men are today raised primarily by their mothers in single-parent homes and often have little or no contact with their fathers?

[33] Indeed, being perceived as attractive, or at least not as ugly, seems to be rather more important to most women that does being perceived as intelligent. Therefore, the question of race differences in attractiveness is seemingly almost as controversial as that of race differences in intelligence. This, then, leads to the delightfully sexist Sailer’s first law of female journalism, which posits that: 

The most heartfelt articles by female journalists tend to be demands that social values be overturned in order that, Come the Revolution, the journalist herself will be considered hotter-looking.” 

[34] A popular alt-right meme has it that there are literally no white-on-black rapes. This is, of course, untrue, and reflects the misreading of a table in a US departnment of Justice report that actually involved only a small sample. In fact, the government does not currently release data on the prevalence of interracial rape. Nevertheless, the US Department of Justice report (mis)cited by some white nationalists does indeed suggest that black-on-white rape is much more common than white-on-black rape in the contemporary USA, a conclusion corroborated by copious other data (e.g. Lebeau 1985).
Thus, in his book Paved with Good Intentions, Jared Taylor reports:

“In a 1974 study in Denver, 40 percent of all rapes were of whites by blacks, and not one case of white-on-black-rape was found. In general, through the 1970s, black-on-white rape was at last ten times more common than white-on-black rape… In 1988 there were 9,406 cases of black-on-white rape and fewer than ten cases of white-on-black rape. Another researcher concludes that in 1989, blacks were three or four times more likely to commit rape than whites and that black men raped white women thirty times as often as white men raped black women” (Paved with Good Intentions: p93). 

Indeed, the authors of one recent textbook on criminology even claim that: 

“Some researchers have suggested, because of the frequency with which African Americans select white victims (about 55 percent of the time), it [rape] could be considered an interracial crime” (Criminology: A Global Perspective: p544). 

Similarly, in the US prison system, where male-male rape is endemic, such assaults disproportionately involve non-white assaults on white inmates, as discussed by the Human Rights Watch report, No Escape: Male Rape in US Prisons

References

Brigandt (2001) The homeopathy of kin selection: an evaluation of van den Berghe’s sociobiological approach to ethnicity. Politics and the Life Sciences 20: 203-215. 
Feinman & Gill (1977) Sex differences in physical attractiveness preferences, Journal of Social Psychology 105(1): 43-52. 
Frost (2008) Sexual selection and human geographic variation. Special Issue: Proceedings of the ND Annual Meeting of the Northeastern Evolutionary Psychology Society. Journal of Social, Evolutionary, and Cultural Psychology, 2(4): 169-191 
Fryer (2007) Guess Who’s Been Coming to Dinner? Trends in Interracial Marriage over the 20th Century, Journal of Economic Perspectives 21(2), pp. 71-90 
Hannam (2021) The rise of the race fakers. Spiked-Online.com, 5 November. 
Hamilton (1964) The genetical evolution of social behaviour I and II, Journal of Theoretical Biology 7:1-16,17-52. 
Hood (2017) The privilege no one wants, American Renaissance, December 11.
Johnson (1986) Kin selection, socialization and patriotism. Politics and the Life Sciences 4(2): 127-154. 
Johnson (1987) In the Name of the Fatherland: An Analysis of Kin Term Usage in Patriotic Speech and Literature. International Political Science Review 8(2): 165-174.
Johnson, Ratwik and Sawyer (1987) The evocative significance of kin terms in patriotic speech pp157-174 in Reynolds, Falger and Vine (eds) The Sociobiology of Ethnocentrism: Evolutionary Dimensions of Xenophobia, Discrimination, Racism, and Nationalism (London: Croom Helm). 
Lebeau (1985) Rape and Racial Patterns. Journal of Offender Counseling Services Rehabilitation, 9(1- 2): 125-148 
Lewis (2011) Who is the fairest of them all? Race, attractiveness and skin color sexual dimorphism. Personality & Individual Differences 50(2): 159-162. 
Lewis (2012) A Facial Attractiveness Account of Gender Asymmetries in Interracial Marriage PLoS One. 2012; 7(2): e31703. 
Lind et al (2007) Elevated male European and female African contributions to the genomes of African American individuals. Human Genetics 120(5) 713-722 
Macdonald 2001 An integrative evolutionary perspective on ethnicity. Poiltics & the Life Sciences 20(1):67-8. 
Rushton (1998a). Genetic similarity theory, ethnocentrism, and group selection. In I. Eibl-Eibesfeldt & F. K. Salter (Eds.), Indoctrinability, Warfare, and Ideology: Evolutionary perspectives (pp. 369-388). Oxford: Berghahn Books. 
Rushton (1998b). Genetic similarity theory and the roots of ethnic conflict. Journal of Social, Political, and Economic Studies, 23, 477-486. 
Rushton, (2005) Ethnic Nationalism, Evolutionary Psychology and Genetic Similarity Theory, Nations and Nationalism 11(4): 489-507. 
Sailer (1997) Is love colorblind? National Review, July 14. 
Sailer (2021) Do 48% of White Male College Applicants Lie About Their Race? Interesting, if It Replicates. Unz Review, October 21. 
Salmon (1998) The Evocative Nature of Kin Terminology in Political Rhetoric. Politics & the Life Sciences, 17(1): 51-57.   
Salter (2000) A Defense and Extension of Pierre van den Berghe’s Theory of Ethnic Nepotism. In James, P. and Goetze, D. (Eds.)  Evolutionary Theory and Ethnic Conflict (Praeger Studies on Ethnic and National Identities in Politics) (Westport, Connecticut: Greenwood Press). 
Salter (2002) Estimating Ethnic Genetic Interests: Is It Adaptive to Resist Replacement Migration? Population & Environment 24(2): 111–140. 
Salter (2008) Misunderstandings of Kin Selection and the Delay in Quantifying Ethnic Kinship, Mankind Quarterly 48(3): 311–344. 
Tooby & Cosmides (1989) Kin selection, genic selection and information dependent strategies Behavioral and Brain Sciences 12(3): 542-544 
Van den Berghe (2005) Review of On Genetic Interests: Family, Ethny and Humanity in the Age of Mass Migration by Frank Salter Nations and Nationalism 11(1) 161-177 
Van den Berghe & Frost (1986) Skin color preference, sexual dimorphism, and sexual selection: A case of gene-culture co-evolution? Ethnic and Racial Studies, 9: 87-113.
Whitney G (1999) The Biological Reality of Race. American Renaissance, October 1999.

‘The Bell Curve’: A Book Much Read About, But Rarely Actually Read

The Bell Curve: Intelligence and Class Structure in American Life by Richard Herrnstein and Charles Murray (New York: Free Press, 1994). 

There’s no such thing as bad publicity’ – or so contends a famous adage of the marketing industry. 

The Bell Curve: Intelligence and Class Structure in America’ by Richard Herrnstein and Charles Murray is perhaps a case in point. 

This dry, technical, academic social science treatise, full of statistical analyses, graphs, tables, endnotes and appendices, and totalling almost 900 pages, became an unlikely nonfiction bestseller in the mid-1990s on a wave of almost universally bad publicity in which the work was variously denounced as racist, pseudoscientific, fascist, social Darwinist, eugenicist and sometimes even just plain wrong. 

Readers who hurried to the local bookstore eagerly anticipating an incendiary racialist polemic were, however, in for a disappointment. 

Indeed, one suspects that, along with The Bible and Stephen Hawkins’ A Brief History of Time, ‘The Bell Curve’ became one of those bestsellers that many people bought, but few managed to finish. 

The Bell Curve’ thus became, like another book that I have recently reviewed, a book much read about, but rarely actually read – at least in full. 

As a result, as with that other book, many myths have emerged regarded the content of ‘The Bell Curve’ that are quite contradicted when one actually takes the time and trouble to read it for oneself. 

Subject Matter 

The first myth of ‘The Bell Curve’ is that it was a book about race differences, or, more specifically, about race differences in intelligence. In fact, however, this is not true. 

Thus, ‘The Bell Curve’ is a book so controversial that the controversy begins with the very identification of its subject-matter. 

On the one hand, the book’s critics focused almost exclusively on subject of race. This led to the common perception that ‘The Bell Curve’ was a book about race and race differences in intelligence.[1]

Ironically, many racialists seem to have taken these leftist critics at their word, enthusiastically citing the work as support for their own views regarding race differences in intelligence.  

On the other hand, however, surviving co-author Charles Murray insisted from the outset that the issue of race, and race differences in intelligence, was always peripheral to he and co-author Richard Herrnstein’s primary interest and focus, which was, he claimed, on the supposed emergence of a ‘Cognitive Elite’ in modern America. 

Actually, however, both these views seem to be incorrect. While the first section of the book does indeed focus on the supposed emergence of a ‘Cognitive Elite’ in modern America, the overall theme of the book seems to be rather broader. 

Thus, the second section of the book focuses on the association between intelligence and various perceived social pathologies, such as unemployment, welfare dependency, illegitimacy, crime and single-parenthood. 

To the extent the book has a single overarching theme, one might say that it is a book about the social and economic correlates of intelligence, as measured by IQ tests, in modern America.  

Its overall conclusion is that intelligence is indeed a strong predictor of social and economic outcomes for modern Americans – high intelligence with socially desirable outcomes and low intelligence with socially undesirable ones. 

On the other hand, however, the topic of race is not quite as peripheral to the book’s themes as sometimes implied by Murray and others. 

Thus, it is sometimes claimed only a single chapter dealt with race. Actually, however, two chapters focus on race differences, namely chapters 13 and 14, respectively titled ‘Ethnic Differences in Cognitive Ability’ and ‘Ethnic Inequalities in Relation to IQ’. 

In addition, a further two chapters, namely chapters 19 and 20, entitled respectively ‘Affirmative Action in Higher Education’ and ‘Affirmative Action in the Workplace’, deal with the topic of affirmative action, as does the final appendix, entitled ‘The Evolution of Affirmative Action in the Workplace’ – and, although affirmative action has been employed to favour women as well as racial minorities, it is with racial preferences that Herrnstein and Murray are primarily concerned. 

However, these chapters represent only 142 of the book’s nearly 900 pages. 

Moreover, in much of the remainder of the book, the authors actually explicitly restrict their analysis to white Americans exclusively. They do so precisely because the well documented differences between the races in IQ as well as in many of the social outcomes whose correlation with IQ the book discusses would mean that race would have represented a potential confounding factor that they would otherwise have to take steps to control for. 

Herrnstein and Murray therefore took to decision to extend their analysis to race differences near the end of their book, in order to address the question of the extent to which differences in intelligence, which they have already demonstrated to be an important correlate of social and economic outcomes among whites, are also capable of explaining differences in achievement as between races. 

Without these chapters, the book would have been incomplete, and the authors would have laid themselves open to the charge of political-correctness and of ignoring the elephant in the room

Race and Intelligence 

If the first controversy of ‘The Bell Curve’ concerns whether it is a book primarily about race and race differences in intelligence, the second controversy is over what exactly the authors concluded with respect to this vexed and contentious issue. 

Thus, the same leftist critics who claimed that ‘The Bell Curve’ was primarily a book about race and race differences in intelligence, also accused the authors of concluding that black people are innately less intelligent than whites

Some racists, as I have already noted, evidently took the leftists at their word, and enthusiastically cite the book as support and authority for this view. 

However, in subsequent interviews, Murray always insisted he and Herrnstein had actually remained “resolutely agnostic” on the extent to which genetic factors underlay the IQ gap. 

In the text itself, Herrnstein and Murray do indeed declare themselves “resolutely agnostic” with regard to the extent of the genetic contribution to the test score gap (p311).

However, just couple of sentences before they use this very phrase, they also appear to conclude that genes are indeed at least part of the explanation, writing: 

It seems highly likely to us that both genes and the environment have something to do with racial differences [in IQ]” (p311). 

This paragraph, buried near the end of chapter 13, during an extended discussion of evidence relating to the causes of race differences in intelligence, is the closest the authors come to actually declaring any definitive conclusion regarding the causes of the black-white test score gap.[2]

This conclusion, though phrased in sober and restrained terms, is, of course, itself sufficient to place its authors outside the bounds of acceptable opinion in the early-twenty-first century, or indeed in the late-twentieth century when the book was first published, and is sufficient to explain, and, for some, justify, the opprobrium heaped upon the book’s surviving co-author from that day forth. 

Intelligence and Social Class

It seems likely that races which evolved on separate continents, in sufficient reproductive isolation from one another to have evolved the obvious (and not so obvious) physiological differences between races that we all observe when we look at the faces, or bodily statures, of people of different races (and that we indirectly observe when we look at the results of different athletic events at the Olympic Games), would also have evolved to differ in psychological traits, including intelligence

Indeed, it is surely unlikely, on a priori grounds alone, that all different human races have evolved, purely by chance, the exact same level of intelligence. 

However, if races differ in intelligence are therefore probable, the case for differences in intelligence as between social classes is positively compelling

Indeed, on a priori grounds alone, it is inevitable that social classes will come to differ in IQ, if one accepts two premises, namely: 

1) Increased intelligence is associated with upward social mobility; and 
2) Intelligence is passed down in families.

In other words, if more intelligent people tend, on average, to get higher-paying jobs than those of lower intelligence, and the intelligence of parents is passed on to their offspring, then it is inevitable that the offspring of people with higher-paying jobs will, on average, themselves be of higher intelligence than are the offspring of people with lower paying jobs.  

This, of course, follows naturally from the infamous syllogism formulated by ‘Bell Curve’ co-author Richard Herrnstein way back in the 1970s (p10; p105). 

Incidentally, this second premise, namely that intelligence is passed down in families, does not depend on the heritability of IQ in the strict biological sense. After all, even if heritability of intelligence were zero, intelligence could still be passed down in families by environmental factors (e.g. the ‘better’ parenting techniques of high IQ parents, or the superior material conditions in wealthy homes). 

The existence of an association between social class and IQ ought, then, to be entirely uncontroversial to anyone who takes any time whatsoever to think about the issue. 

If there remains any room for reasoned disagreement, it is only over the direction of causation – namely the question of whether:  

1) High intelligence causes upward social mobility; or 
2) A privileged upbringing causes higher intelligence.

These two processes are, of course, not mutually exclusive. Indeed, it would seem intuitively probable that both factors would be at work. 

Interestingly, however, evidence demonstrates the occurrence only of the former. 

Thus, even among siblings from the same family, the sibling with the higher childhood IQ will, on average, achieve higher socioeconomic status as an adult. Likewise, the socioeconomic status a person achieves as an adult correlates more strongly with their own IQ score than it does with the socioeconomic status of their parents or of the household they grew up in (see Straight Talk About Mental Tests: p195). 

In contrast, family, twin and adoption studies and of the sort conducted by behavioural geneticists have concurred in suggesting that the so-called shared family environment (i.e. those aspects of the family environment shared by siblings from the same household, including social class) has but little effect on adult IQ. 

In other words, children raised in the same home, whether full- or half-siblings or adoptees, are, by the time they reach adulthood, no more similar to one another in IQ than are children of the same degree of biological relatedness brought up in entirely different family homes (see The Nurture Assumption: reviewed here). 

However, while the direction of causation may still be disputed by intelligent (if uninformed) laypeople, the existence of an association between intelligence and social class ought not, one might think, be in dispute. 

However, in Britain today, in discussions of social mobility, if children from deprived backgrounds are underrepresented, say, at elite universities, then this is almost invariably taken as incontrovertible proof that the system is rigged against them. The fact that children from different socio-economic backgrounds differ in intelligence is almost invariably ignored. 

When mention is made of this incontrovertible fact, leftist hysteria typically ensues. Thus, in 2008, psychiatrist Bruce Charlton rightly observed that, in discussion of social mobility: 

A simple fact has been missed: higher social classes have a significantly higher average IQ than lower social classes (Clark 2008). 

For his trouble, Charlton found himself condemned by the National Union of Students and assorted rent-a-quote academics and professional damned fools, while even the ostensibly ‘right-wing’ Daily Mail newspaper saw fit to publish a headline Higher social classes have significantly HIGHER IQs than working class, claims academic, as if this were in some way a controversial or contentious claim (Clark 2008). 

Meanwhile, when, in the same year, a professor at University College a similar point with regard the admission of working-class students to medical schools, even the then government Health Minister, Ben Bradshaw, saw fit to offer his two cents worth (which were not worth even that), declaring: 

It is extraordinary to equate intellectual ability with social class” (Beckford 2008). 

Actually, however, what is truly extraordinary is that any intelligent person, least of all a government minister, would dispute the existence of such a link. 

Cognitive Stratification 

Herrnstein’s syllogism leads to a related paradox – namely that, as environmental conditions are equalized, heritability increases. 

Thus, as large differences in the sorts of environmental factors known to affect IQ (e.g. malnutrition) are eliminated, so differences in income have come to increasingly reflect differences in innate ability. 

Moreover, the more gifted children from deprived backgrounds who escape their humble origins, then, given the substantial heritability of IQ, the fewer such children will remain among the working-class in subsequent generations. 

The result is what Herrnstein and Murray call the ‘Cognitive Stratification’ of society and the emergence of what they call a ‘Cognitive Elite’. 

Thus, in feudal society, a man’s social status was determined largely by ‘accident of birth’ (i.e. he inherited the social station of his father). 

Women’s status, meanwhile, was determined, in addition, by what we might call ‘accident of marriage’ – and, to a large extent, it still is

However, today, a person’s social status, at least according to Herrnstein and Murray, is determined primarily, and increasingly, by their level of intelligence. 

Of course, people are not allocated to a particular social class by IQ testing itself. Indeed, the use of IQ tests by employers and educators has been largely outlawed on account of its disparate impact (or indirect discrimination’, to use the equivalent British phrase) with regard to race (see below). 

However, the skills and abilities increasingly valued at a premium in western society (and, increasingly, many non-western societies as well), mean that, through the operation of the education system and labour market, individuals are effectively sorted by IQ, even without anyone ever actually sitting an IQ test. 

In other words, society is becoming increasingly meritocratic – and the form of ostensible ‘merit’ upon which attainment is based is intelligence. 

For Herrnstein and Murray, this is a mixed blessing: 

That the brightest are identified has its benefits. That they become so isolated and inbred has its costs” (p25). 

However, the correlation between socioeconomic status and intelligence remains imperfect. 

For one thing, there are still a few highly remunerated, and very high-status, occupations that rely on skills that are not especially, if at all, related to intelligence.  I think here, in particular, of professional sports and the entertainment industry. Thus, leadings actors, pop stars and sports stars are sometimes extremely well-remunerated, and very high-status, but may not be especially intelligent.  

More importantly, while highly intelligent people might be, by very definition, the only ones capable of performing cognitively-demanding, and hence highly remunerated, occupations, this is not to say all highly intelligent people are necessarily employed in such occupations. 

Thus, whereas all people employed in cognitively-demanding occupations are, almost by definition, of high intelligence, people of all intelligence levels are capable of doing cognitively-undemanding jobs.

Thus, a few people of high intellectual ability remain in low-paid work, whether on account of personality factors (e.g. laziness), mental illness, lack of opportunity or sometimes even by choice (which choice is, of course, itself a reflection of personality factors). 

Therefore, the correlation between IQ and occupation is far from perfect. 

Job Performance

The sorting of people with respect to their intelligence begins in the education system. However, it continues in the workplace. 

Thus, general intelligence, as measured by IQ testing, is, the authors claim, the strongest predictor of occupational performance in virtually every occupation. Moreover, in general, the higher paid and higher status the occupation in question, the stronger the correlation between performance and IQ. 

However, Herrnstein and Murray are at pains to emphasize, intelligence is a strong predictor of occupational performance even in apparently cognitively undemanding occupations, and indeed almost always a better predictor of performance than tests of the specific abilities the job involves on a daily basis. 

However, in the USA, employers are barred from using testing to select among candidates for a job or for promotion unless they can show the test has ‘manifest relationship’ to the work, and the burden of proof is on the employer to show such a relationship. Otherwise, given their disparate impact’ with regard to race (i.e. the fact that some groups perform worse), the tests in question are deemed indirectly discriminatory and hence unlawful. 

Therefore, employers are compelled to test, not general ability, but rather the specific skills required in the job in question, where a ‘manifest relationship’ is easier to demonstrate in court. 

However, since even tests of specific abilities almost invariably still tap into the general factor of intelligence, races inevitably score differently even on these tests. 

Indeed, because of the ubiquity and predictive power of the g factor, it is almost impossible to design any type of standardized test, whether of specific or general ability or knowledge, in which different racial groups do not perform differently. 

However, if some groups outperform others, the American legal system presumes a priori that this reflects test bias rather than differences in ability. 

Therefore, although the words all men are created equal are not, contrary to popular opinion, part of the US constitution, the Supreme Court has effectively decided, by legal fiat, to decide cases as if they were. 

However, just as a law passed by Congress cannot repeal the law of gravity, so a legal presumption that groups are equal in ability cannot make it so. 

Thus, the bar on the use of IQ testing by employers has not prevented society in general from being increasingly stratified by intelligence, the precise thing measured by the outlawed tests. 

Nevertheless, Herrnstein and Murray estimate that the effective bar on the use of IQ testing makes this process less efficient, and cost the economy somewhere between 80 billion to 13 billion dollars in 1980 alone (p85). 

Conscientiousness and Career Success

I am skeptical of Herrnstein and Murray’s conclusion that IQ is the best predictor of academic and career success. I suspect hard work, not to mention a willingness to toady, toe the line, and obey orders, is at least as important in even the most cognitively-demanding careers, as well as in schoolwork and academic advancement. 

Perhaps the reason these factors have not (yet) been found to be as highly correlated with earnings as is IQ is that we have not yet developed a way of measuring these aspects of personality as accurately as we can measure a person’s intelligence through an IQ test. 

For example, the closest psychometricians have come to measuring capacity for hard work is the personality factor known as conscientiousness, one of the Big Five factors of personality revealed by psychometric testing. 

Conscientiousness does indeed correlate with success in education and work (e.g. Barrick & Mount 1991). However, the correlation is weaker than that between IQ and success in education and at work. 

However, this may be because personality is less easily measured by current psychometric methods than is intelligence – not least because personality tests generally rely on self-report, rather than measuring actual behaviour

Thus, to assess conscientiousness, questionnaires ask respondents whether they ‘see themselves as organized’, ‘as able to follow an objective through to completion’, ‘as a reliable worker’, etc. 

This would be the equivalent of an IQ test that, instead of directly testing a person’s ability to recognize patterns or manipulate shapes by having them do just this, simply asked respondents how good they perceived themselves as being at recognizing patterns, or manipulating shapes. 

Obviously, this would be a less accurate measure of intelligence than a normal IQ test. After all, some people lie, some are falsely modest and some are genuinely deluded. 

Indeed, according to the Dunning Kruger effect, it is those most lacking in ability who most overestimate their abilities – precisely because they lack the ability to accurately assess their ability (Kruger & Dunning 1999). 

In an IQ test, on the other hand, one can sometimes pretend to be dumber than one is, by deliberately getting questions wrong that one knows the answer to.[3]

However, it is not usually possible to pretend to be smarter than one is by getting more questions right simply because one would not know what are the right answers. 

Affirmative Action’ and Test Bias 

In chapters nineteen and twenty, respectively entitled ‘Affirmative Action in Higher Education’ and ‘Affirmative Action in the Workplace’, the authors discuss so-called affirmative action, an American euphemism for systematic and overt discrimination against white males. 

It is well-documented that, in the United States, blacks, on average, earn less than white Americans. On the other hand, it is less well-documented that whites, on average, earn less than people of IndianChinese and Jewish ancestry

With the possible exception of Indian-Americans, these differences, of course, broadly mirror those in average IQ scores. 

Indeed, according to Herrnstein and Murray, the difference in earnings between whites and blacks, not only disappears after controlling for differences in IQ, but is actually partially reversed. Thus, blacks are actually somewhat overrepresented in professional and white-collar occupations as compared to whites of equivalent IQ. 

This remarkable finding Herrnstein and Murray attribute to the effects of affirmative action programmes, as black Americans are appointed and promoted beyond what their ability merits because through discrimination. 

Interestingly, however, this contradicts what the authors wrote in an earlier chapter, where they addressed the question of test bias (pp280-286). 

There, they concluded that testing was not biased against African-Americans, because, among other reasons, IQ tests were equally predictive of real-world outcomes (e.g. in education and employment) for both blacks and whites, and blacks do not perform any better in the workplace or in education than their IQ scores predict. 

This is, one might argue, not wholly convincing evidence that IQ tests are not biased against blacks. It might simply suggest that society at large, including the education system and the workplace, is just as biased against blacks as are the hated IQ tests. This is, of course, precisely what we are often told by the television, media and political commentators who insist that America is a racist society, in which such mysterious forces as ‘systemic racism’ and ‘white privilege’ are pervasive. 

In fact, the authors acknowledge this objection, conceding:  

The tests may be biased against disadvantaged groups, but the traces of bias are invisible because the bias permeates all areas of the group’s performance. Accordingly, it would be as useless to look for evidence of test bias as it would be for Einstein’s imaginary person traveling near the speed of light to try to determine whether time has slowed. Einstein’s traveler has no clock that exists independent of his space-time context. In assessing test bias, we would have no test or criterion measure that exists independent of this culture and its history. This form of bias would pervade everything” (p285). 

Herrnstein and Murray ultimately reject this conclusion on the grounds that it is simply implausible to assume that: 

“[So] many of the performance yardsticks in the society at large are not only biased, they are all so similar in the degree to which they distort the truth-in every occupation, every type of educational institution, every achievement measure, every performance measure-that no differential distortion is picked up by the data” (p285). 

In fact, however, Nicholas Mackintosh identifies one area where IQ tests do indeed under-predict black performance, namely with regard to so-called adaptive behaviours – i.e. the ability to cope with day-to-day life (e.g. feed, dress, clean, interact with others in a ‘normal’ manner). 

Blacks with low IQs are generally much more functional in these respects than whites or Asians with equivalent low IQs (see IQ and Human Intelligence: p356-7).[4]

Yet Herrnstein and Murray seem to have inadvertently, and evidently without realizing it, identified yet another sphere where standardized testing does indeed under-predict real-world outcomes for blacks. 

Thus, if indeed, as Herrnstein and Murray claim, blacks are somewhat overrepresented among professional and white-collar occupations relative to their IQs, this suggests that blacks do indeed do better in real-world outcomes than their test results would predict and, while Herrnstein and Murray attribute this to the effect of discrimination against whites, it could instead surely be interpreted as evidence that the tests are biased against blacks. 

Policy Implications? 

What, then, are the policy implications that Herrnstein and Murray draw from the findings that they report? 

In The Blank Slate: The Modern Denial of Human Nature, cognitive science, linguist and popular science writer Steven Pinker popularizes the notion that recognizing the existence of innate differences between individuals and groups in traits such as intelligence does not necessarily lead to ‘right-wing’ political implications. 

Thus, a leftist might accept the existence of innate differences in ability, but conclude that, far from justifying inequality, this is all the more reason to compensate the, if you like, ‘cognitively disadvantaged’ for their innate deficiencies, differences which are, being innate, hardly something for which they can legitimately be blamed. 

Herrnstein and Murray reject this conclusion, but acknowledge it is compatible with their data. Thus, in an afterword to later editions, Murray writes: 

If intelligence plays an important role in determining how well one does in life, and intelligence is conferred on a person through a combination of genetic and environmental factors over which that person has no control, the most obvious political implication is that we need a Rawlsian egalitarian state, compensating the less advantaged for the unfair allocation of intellectual gifts” (p554).[5]

Interestingly, Pinker’s notion of a ‘hereditarian left’, and the related concept of Bell Curve liberals, is not entirely imaginary. On the contrary, it used to be quite mainstream. 

Thus, it was the radical leftist post-war Labour government that imposed the tripartite system on schools in the UK in 1945, which involved allocating pupils to different schools on the basis of their performance in what was then called the 11-plus exam, conducted at with children at age eleven, which tested both ability and acquired knowledge. This was thought by leftists to be a fair system that would enable bright, able youngsters from deprived and disadvantaged working-class backgrounds to achieve their full potential.[6]

Indeed, while contemporary Cultural Marxists emphatically deny the existence of innate differences in ability as between individuals and groups, Marx himself, laboured under no such delusion

On the contrary, in advocating, in his famous (plagiarized) aphorism From each according to his ability; to each according to his need, Marx implicitly recognized that individuals differ in “ability”, and, given that, in the unrealistic communist utopia he envisaged, environmental conditions were ostensibly to be equalized, these differences he presumably conceived of as innate in origin. 

However, a distinction must be made here. While it is possible to justify economic redistributive policies on Rawlsian grounds, it is not possible to justify affirmative action

Thus, one might well reasonably contend that the ‘cognitively disadvantaged’ should be compensated for their innate deficiencies through economic redistribution. Indeed, to some extent, most Western polities already do this, by providing welfare payments and state-funded, or state-subsidized, care to those whose cognitive impairment is such as to qualify as a disability and hence render them incapable of looking after or providing for themselves. 

However, we are unlikely to believe that such persons should be given entry to medical school such that they are one day liable to be responsible for performing heart surgery on us or diagnosing our medical conditions. 

In short, socialist redistribution is defensible – but affirmative action is definitely not! 

Reception and Readability 

The reception accorded ‘The Bell Curve’ in 1994 echoed that accorded another book that I have also recently reviewed, but that was published some two decades earlier, namely Edward O. Wilson’s Sociobiology: The New Synthesis

Both were greeted with similar indignant moralistic outrage by many social scientists, who even employed similar pejorative soundbites (‘genetic determinism’, reductionism, ‘biology as destiny’), in condemning the two books. Moreover, in both cases, the academic uproar even spilled over into a mainstream media moral panic, with pieces appearing the popular press attacking the two books. 

Yet, in both cases, the controversy focused almost exclusively on just a small part of each book – the single chapter in Sociobiology: The New Synthesis focusing on humans and the few chapters in ‘The Bell Curve’ discussing race. 

In truth, however, both books were massive tomes of which these sections represented only a small part. 

Indeed, due to their size, one suspects most critics never actually read the books in full for themselves, including, it seemed, most of those nevertheless taking it upon themselves to write critiques. This is what led to the massive disconnect between what most people thought the books said, and their actual content. 

However, there is a crucial difference. 

Sociobiology: The New Synthesis was a long book of necessity, given the scale of the project Wilson set himself. 

As I have written in my review of that latter work, the scale of Wilson’s ambition can hardly be exaggerated. He sought to provide a new foundation for the whole field of animal behaviour, then, almost as an afterthought, sought to extend this ‘New Synthesis’ to human behaviour as well, which meant providing a new foundation, not for a single subfield within biology, but for several whole disciplines (psychology, sociology, economics and cultural anthropology) that were formerly almost unconnected to biology. Then, in a few provocative sentences, he even sought to provide a new foundation for moral philosophy, and perhaps epistemology too. 

Sociobiology: The New Synthesis was, then, inevitably and of necessity, a long book. Indeed, given that his musings regarding the human species were largely (but not wholly) restricted to a single chapter, one could even make a case that it was too short – and it is no accident that Wilson subsequently extended his writings with regard to the human species to a book length manuscript

Yet, while Sociobiology was of necessity a long book, ‘The Bell Curve: Intelligence and Class Structure in America’ is, for me, unnecessarily overlong. 

After all, Herrnstein and Murray’s thesis was actually quite simple – namely that cognitive ability, as captured by IQ testing, is a major correlate of many important social outcomes in modern America. 

Yet they reiterate this point, for different social outcomes, again and again, chapter after chapter, repeatedly. 

In my view, Herrnstein and Murray’s conclusion would have been more effectively transmitted to the audience they presumably sought to reach had they been more succinct in their writing style and presentation of their data. 

Had that been the case then perhaps rather more of the many people who bought the book, and helped make it into an unlikely nonfiction bestseller in 1994, might actually have managed to read it – and perhaps even been persuaded by its thesis. 

For casual readers interested in this topic, I would recommend instead Intelligence, Race, And Genetics: Conversations With Arthur R. Jensen (which I have reviewed herehere and here). 

Endnotes

[1] For example, Francis Wheen, a professional damned fool and columnist for the Guardian newspaper (which two occupations seem to be largely interchangeable) claimed that: 

The Bell Curve (1994), runs to more than 800 pages but can be summarised in a few sentences. Black people are more stupid than white people: always have been, always will be. This is why they have less economic and social success. Since the fault lies in their genes, they are doomed to be at the bottom of the heap now and forever” (Wheen 2000). 

In making this claim, Wheen clearly demonstrates that he has read few if any of those 800 pages to which he refers.

[2] Although their discussion of the evidence relating to the causes, genetic or environmental, of the black-white test score gap is extensive, it is not exhaustive. For example, Phillipe Rushton, the author of Race Evolution and Behavior (reviewed here and here) argues that, despite the controversy their book provoked, Herrnstein and Murray actually didn’t go far enough on race, omitting, for example, any real discussion, save a passing mention in Appendix 5, of race differences in brain size (Rushton 1997). On the other hand, Herrnstein and Murray also did not mention studies that failed to establish any correlation between IQ and blood groups among African-Americans, studies interpreted as supporting an environmentalist interpretation of race differences in intelligence (Loehlin et al 1973Scarr et al 1977). For readers interested in a more complete discussion of the evidence regarding the relative contributions of environment and heredity to the differences in IQ scores of different races, see my review of Richard Lynn’s Race Differences in Intelligence: An Evolutionary Analysis, available here.

[3] For example, some of those accused of serious crimes have been accused of deliberately getting questions wrong on IQ tests in order to qualify as mentally subnormal when before the courts for sentencing in order to be granted mitigation of sentence on this ground, or, more specifically, in order to evade the death penalty

[4] This may be because whites or Asians with such low IQs are more likely to have such impaired cognitive abilities because of underlying conditions (e.g chromosomal abnormalitiesbrain damage) that handicap them over and above the deficit reflected in IQ score alone. On the other hand, blacks with similarly low IQs are still within the normal range for their own race. Therefore, rather than suffering from, say, a chromosomal abnormality or brain damage, they are relatively more likely to simply be at the tail-end of the normal range of IQs within their group, and hence normal in other respects.

[5] The term Rawlsian is a reference to political theorist John Rawles version of social contract theory, whereby he poses the hypothetical question as to what arrangement of political, social and economic affairs humans would favour if placed in what he called the original position, where they would be unaware of, not only their own race, sex and position in to the socio-economic hierarchy, but also, most important for our purposes, their own level of innate ability. This Rawles referred to as ‘veil of ignorance’.

[6] The tripartite system did indeed enable many working-class children to achieve a much higher economic status than their parents, although this was partly due to the expansion of the middle-class sector of the economy over the same time-period. It was also later Labour administrations who largely abolished the 11-plus system, not least because, unsurprisingly given the heritability of intelligence and personality, children from middle-class backgrounds tended to do better on it than did children from working-class backgrounds.

References 

Barrick & Mount 1991 The big five personality dimensions and job performance: a meta-analysis. Personnel Psychology 44(1):1–26 
Beckford (2008) Working classes ‘lack intelligence to be doctors’, claims academicDaily Telegraph, 04 Jun 2008. 
Clark 2008 Higher social classes have significantly HIGHER IQs than working class, claims academic Daily Mail, 22 May 2008. 
Kruger & Dunning (1999) Unskilled and Unaware of It: How Difficulties in Recognizing One’s Own Incompetence Lead to Inflated Self-AssessmentsJournal of Personality and Social Psychology 77(6):1121-34 
Loehlin et al (1973) Blood group genes and negro-white ability differencesBehavior Genetics 3(3): 263-270  
Rushton, J. P. (1997). Why The Bell Curve didn’t go far enough on race. In E. White (Ed.), Intelligence, political inequality, and public policy (pp. 119-140). Westport, CT: Praeger. 
Scarr et al (1977) Absence of a relationship between degree of white ancestry and intellectual skills within a black population. Human Genetics 39(1):69-86 . 
Wheen (2000) The ‘science’ behind racismGuardian, 10 May 2000. 

John R Baker’s ‘Race’: “A Reminder of What Was Possible Before the Curtain Came Down”

‘Race’, by John R. Baker, Oxford University Press, 1974.

John Baker’s ‘Race’ represents a triumph of scholarship across a range of fields, including biology, ancient history, archaeology, history of science, psychometrics and anthropology.

First published by Oxford University Press in 1974, it also marks a watershed in Western thought – the last time a major and prestigious publisher put its name to an overtly racialist work.

As science writer Marek Kohn writes:

Baker’s treatise, compendious and ponderous, is possible the last major statement of traditional race science written in English” (The Race Gallery: p61).

Inevitably for a scientific work first published over forty years ago, ‘Race’ is dated. In particular, the DNA revolution in population genetics has revolutionized our understanding of the genetic differences and relatedness between different human populations.

Lacking access to such data, Baker had only indirect phenotypic evidence (i.e. the morphological similarities and differences between different peoples), as well as historical and geographic evidence, with which to infer such relationships and hence construct his racial phylogeny and taxonomy.

Phenotypic similarity is obviously a less reliable method of determining the relatedness between groups than is provided by genome analysis, since there is always the problem of distinguishing homology from analogy and hence misinterpreting a trait that has independently evolved in different populations as evidence of relatedness.[1]

However, I found only one case of genetic studies decisively contradicting Baker’s conclusions. Thus, whereas Baker classes the Ainu People of Japan as Europid (p158; p173; p424; p625), recent genetic studies suggest that the Ainu have little or no genetic affinities to Caucasoid populations and are most closely related to other East Asians.[2]

On the other hand, however, Baker’s omission of genetic data means that, unusually for a scientific work, in the material he does cover, ‘Race’ scarcely seems to have dated at all. This is because the primary focus of Baker’s book – namely, morphological differences between races – is a field of study that has become politically suspect and in which new research has now all but ceased.[3]

Yet in the nineteenth- and early-twentieth century, when the discipline of anthropology first emerged as a distinct science, the study of race differences in morphology was the central focus of the entire science of anthropology.

Thus, Baker’s ‘Race’ can be viewed as the final summation of the accumulated findings of the ‘old-stylephysical anthropology of the nineteenth and early-twentieth centuries, published at the very moment this intellectual tradition was in its death throes.

Accessibility

Baker’s ‘Race’ is indeed a magnum opus. Unfortunately, however, at over 600 pages, embarking on reading ‘Race’ might seem almost like a lifetime’s work in and of itself.

Not only is it a very long book, but, in addition, much of the material, particularly on morphological race differences and their measurement, is highly technical, and will be readily intelligible only to the dwindling band of biological anthropologists who, in the genomic age, still study such things.

This inaccessibility is exacerbated by the fact that Baker does not use endnotes, except for his references, and only very occasionally uses footnotes. Instead, he includes even technical and peripheral material in the main body of his text, but indicates that material is technical or peripheral by printing it in a smaller font-size.[4]

Baker’s terminology is also confusing.[5] He prefers the ‘-id’ suffix to the more familiar ‘-oid’ and ‘-ic’ (e.g. ‘Negrid‘ and ‘Nordid‘ rather than ‘Negroid’ and ‘Nordic‘) and eschews the familiar terms Caucasian or Caucasoid, on the grounds that:

The inhabitants of the Caucasus region are very diverse and very few of them are typical of any large section of Europids” (p205).

However, his own preferred alternative term, ‘Europid’, is arguably equally misleading as it contributes to the already common conflation of Caucasian with white European, even though, as Baker is at pains to emphasize elsewhere in his treatise, populations from the Middle East, North Africa and even the Indian subcontinent are also ‘Europid’ (i.e. Caucasoid) in Baker’s judgement.

In contrast, the term Caucasoid, or even Caucasian, causes little confusion in my experience, since it is today generally understood as a racial term and not as a geographical reference to the Caucasus region.[6]

At any rate, a similar criticism could surely be levelled at the term ‘Mongoloid’ (or, as Baker prefers, ‘Mongolid’), since Mongolian people are similarly quite atypical of other East Asian populations, and, despite the brief ascendancy of the Mongol Empire, and its genetic impact (as well as that previous waves of conquest by horse peoples of the Eurasian Steppe), were formerly a rather marginal people confined to the arid fringes of the indigenous home range of the so-called Mongoloid race, which had long been centred in China, the self-styled Middle Kingdom.[7]

Certainly, the term ‘Caucasoid’ makes little etymological sense. However, this is also true of a lot of words which we nevertheless continue to make use of. Indeed, since all words change in meaning over time, the original meaning of a word is almost invariably different to its current accepted usage.[8]

Yet we continue to use these words so as to make ourselves intelligible to others, the only alternative being to invent an entirely new language all of our own which only we would be capable of understanding.

Unfortunately, however, too many racial theorists, Baker included, have insisted on creating entirely new racial terms of their own coinage, or sometimes new entire lexicons, which, not only causes confusion among readers, but also leads the casual reader to underestimate the actual degree of substantive agreement between different authors, who, though they use different terms, often agree regarding both the identity of, and relationships between, the major racial groupings.[9]

Historical Focus

Another problem is the book’s excessive historical focus.

Judging the book by its contents page, one might imagine that Baker’s discussion of the history of racial thought is confined to the first section of the book, titled “The Historical Background” and comprising four chapters that total just over fifty pages.

However, Baker acknowledges in the opening page of his preface that:

Throughout this book, what might be called the historical method has been adopted as a matter of deliberate policy” (p3).

Thus, in the remainder of the book, Baker continues to adopt an historical perspective, briefly charting the history behind the discovery of each concept, archaeological discovery, race difference or method of measuring race differences that he introduces.

In short, it seems that Baker is not content with writing about science; he wants to write history of science too.

A case in point is Chapter Eight, which, despite its title (“Some Evolutionary and Taxonomic Theories”), actually contains very little on modern taxonomic or evolutionary theory, or even what would pass for ‘modern’ when Baker wrote the book over forty years ago.

Instead, the greater part of the chapter is devoted to tracing the history of two theories that were, even at the time Baker was writing, already wholly obsolete and discredited (namely, recapitulation theory and orthogenesis).

Let me be clear, Baker himself certainly agrees that these theories are obsolete and discredited, as this is his conclusion at the end of the respective sections devoted to discussion of these theories in his chapter on “Evolutionary and Taxonomic Theories”.

However, this only begs the question as to why Baker chooses to devote so much space in this chapter to discussing these theories in the first place, given that both theories are discredited and also of only peripheral relevance to his primary subject-matter, namely the biology of race.

Anyone not interested in these topics, or in history of science more generally, is well advised to skip the majority of this chapter.

The Historical Background

Readers not interested in the history of science, and concerned only with contemporary state-of-the-art science (or at least the closest an author writing in 1974 can get to modern state-of-the-art science) may also be tempted to skip over the whole first section of the book, entitled, as I have said, “The Historical Background”, and comprised of four chapters or, in total, just over fifty pages.

These days, when authoring a book on the biology of race, it seems to have become almost de rigueur to include an opening chapter, or chapters, tracing the history of race science, and especially its political misuse during nineteenth and early twentieth-centuries (e.g. under the Nazis).[10]

The usual reason for including these chapters is for the author or authors to thereby disassociate themselves from the earlier supposed misuse of race science for nefarious political purposes, and emphasize how their own approach is, of course, infinitely more scientific and objective than that of their sometimes less than illustrious intellectual forebears.

However, Baker’s discussion ofThe Historical Background” is rather different, and refreshingly short on disclaimers, moralistic grandstanding and benefit-of-hindsight condemnations that one usually finds in such potted histories.

Instead, Baker strives to give all views, howsoever provocative, a fair hearing in as objective and sober a tone as possible.[11]

Only Lothrop Stoddard, strangely, is dismissed altogether. The latter is, for Baker, an “obviously unimportant” thinker, whose book “contains nothing profound or genuinely original” (p58-9).

Yet this is perhaps unfair. Whatever the demerits of Stoddard’s racial taxonomy (“oversimplified to the point of crudity,” according to Baker: p58), Stoddard’s geopolitical and demographic predictions have proven prescient.[12]

Overall, Baker draws two general conclusions regarding the history of racial thought in the nineteenth and early twentieth century.

First, he observes how few of the racialist authors whom he discusses were anti-Semitic. Thus, Baker reports:

Only one of the authors, Lapouge, strongly condemns the Jews. Treitschke is moderately anti-Jewish; Chamberlain, Grant and Stoddard mildly so; Gobineau is equivocal” (p59).

The rest of the authors whom he discusses evince, according to Baker, “little or no interest in the Jewish problem”, the only exception being Friedrich Nietzsche, who is “primarily an anti-egalitarian, but [who] did not proclaim the inequality of ethnic taxa”, and who, in his comments regarding the Jewish people, or at least those selectively quoted by Baker, is positively gushing in his praise.

In fact, however, Nietzsche’s views regarding the Jewish people are rather more complex than Baker allows, including as they do both critical comments and no few backhanded complements, since he primarily blames the Jews for the invention of Christianity and of the slave morality that was its legacy.

Thus, anti-Semitism often goes hand-in-hand with philosemitism. Thus, both Nietzsche and Count de Gobineau indeed wrote passages that, at least when quoted in isolation, seem highly complementary regarding the Jewish people. However, it is well to bear in mind that Hitler did as well, the latter writing in Mein Kampf:

The mightiest counterpart to the Aryan is represented by the Jew. In hardly any people in the world is the instinct of self- preservation developed more strongly than in the so-called ‘chosen’. Of this, the mere fact of the survival of this race may be considered the best proof” (Mein Kampf, Manheim translation).[13]

Thus, as a character from a Michel Houellebecq novel observes:

All anti-Semites agree that the Jews have a certain superiorityIf you read anti-Semitic literature, you’re struck by the fact that the Jew is considered to be more intelligent, more cunning, that he is credited with having singular financial talents – and, moreover, greater communal solidarity. Result: six million dead” (Platform: p113) 

Baker’s second general observation is similarly curious, namely that:

None of the authors mentioned in these chapters claims superiority for the whole of the Europid race: it is only a subrace, or else a section of the Europid race not clearly defined in terms of physical anthropology, that is favoured” (p59).

In retrospect, this seems anomalous, especially given that the so-called Nordic race, on whose behalf racial supremacy was most often claimed, actually came relatively late to civilization, which began in the Middle East, North Africa and South Asia, arriving in Europe only with the Mediterranean civilizations of Greece and Rome, and in Northern Europe later still.

However, this focus on the alleged superiority of certain European subraces rather than Caucasians as a whole likely reflects the fact that, during the time period in which these works were written, European peoples and nations were largely in competition and conflict with other European peoples and nations.

Only in European overseas colonies were Europeans in contact and conflict with non-European races, and, even here, the main obstacle to imperial expansion was, not so much the opposition of the often primitive non-European races whom the Europeans sought to colonize, but rather that of rival colonizers from other European nations.

Therefore, it was the relative superiority of different European populations which was naturally of most concern to Europeans during this time period.

In contrast, the superiority of the Caucasian race as a whole was of comparably little interest, if only because it was something that these writers already took very much for granted, and hence hardly worth wasting ink or typeface over.

The Rise of Racial Egalitarianism

There are two curious limitations that Baker imposes on his historical survey of racial thought. First, at the beginning of Chapter Three (From Gobineau to Houston Chamberlain’), he announces:

The present chapter and the next [namely, those chapters dealing with the history of racial thinking from the mid-nineteenth century up until the early-twentieth century] differ from the two preceding ones… in the more limited scope. It is are concerned only with the growth of ideas that favoured belief in the inequality of ethnic taxa or are supposedrightly or wronglyto have favoured this belief” (p33).

Given that I have already criticised ‘Race’ as overlong, and as having an excessive historical focus, I might be expected to welcome this restriction. However, Baker provides no rationale for this self-imposed restriction.

Certainly, it is rare, and enlightening, to read balanced, even sympathetic, accounts of the writings of such infamous racialist thinkers as Gobineau, Galton and Chamberlain, whose racial views are today usually dismissed as so preposterous as hardly to merit serious consideration. Moreover, in the current political climate, such material even acquires a certain allure of the forbidden’.

However, thinkers championing racial egalitarianism have surely proven more influential, at least in the medium-term. Yet such enormously influential thinkers as Franz Boas and Ashley Montagu pass entirely unmentioned in Baker’s account.[14]

Moreover, the intellectual antecedents of Nazism have already been extensively explored by historians. In contrast, however, the rise of the dogma of racial equality has passed largely unexamined, perhaps because to examine its origins is to expose the weakness of its scientific basis and its fundamentally political origins.[15]

Yet the story of how the theory of racial equality was transformed from a maverick, minority opinion among scientists and laypeople alike into a sacrosanct contemporary dogma which a person, scientist or layperson, can question only at severe cost to their career, livelihood and reputation is surely one worth telling.

The second restriction that Baker imposes upon his history is that he concludes it, prematurely, in 1928. He justifies closing his survey in this year on the grounds that this date supposedly:

Marks the close of the period in which both sides in the ethnic controversy were free to put forward their views, and authors who wished to do so could give objective accounts of the evidence pointing in each direction” (p61).

Yet this cannot be entirely true, for, if it were, then Baker’s own book could never have been published – unless, of course, Baker regards his own work as something other than an “objective account of the evidence pointing in each direction”, which seems doubtful.

Certainly, the influence of what is now called political correctness is to be deplored for impact on science, university appointments, the allocation of research funds and the publishing industry. However, there has surely been no abrupt watershed but rather a gradual closing of the western mind over time.

Thus, it is notable that other writers have cited dates a little later than that quoted by Baker, often coinciding with the defeat of Nazi Germany and exposure of the Nazi genocide, or sometimes the defeat of segregation in the American South.

Indeed, not only was this process gradual, it has also proceeded apace in the years since Baker’s ‘Race’ first came off the presses, such that today such a book would surely never would have been published in the first place, certainly not by as prestigious a publisher as Oxford University Press (who, surely not uncoincidently, soon gave up the copyright).[16]

Moreover, Baker is surely wrong to claim that it is impossible:

To follow the general course of controversy on the ethnic problem, because, for the reason just stated [i.e. the inability of authors of both sides to publicise their views], there has been no general controversy on the subject” (p61).

On the contrary, the issue remains as incendiary as ever, with the bounds of acceptable opinion seemingly ever narrowing and each year a new face falling before the witch hunters of the contemporary racial inquisition.

Biology

Having dealt in his first section with what he calls “The Historical Background”, Baker next turns to what he calls “The Biological Background”. He begins by declaring, rightly, that:

Racial problems cannot be understood by anyone whose interests and field of knowledge stop short at the limit of purely human affairs” (p3).

This is surely true, not just of race, but of all issues in human biology, psychology, sociology, anthropology and political science, as the recent rise of sociobiology and evolutionary psychology attests. Indeed, Baker even coins a memorable and quotable aphorism to this effect, when he declares:

No one knows Man who knows only Man” (p65).

However, Baker sometimes takes this thinking rather too far, even for my biologically-inclined tastes.

Certainly, he is right to emphasise that differences among human populations are analogous to those found among other species. Thus, his discussion of racial differences among our primate cousins are of interest, but also somewhat out-of-date.[17]

However, his intricate and fully illustrated nine-page description of race differences among the different subspecies of crested newt stretched the patience of this reader (p101-109).

Are Humans a Single Species?

Whereas Baker’s seventh chapter (“The Meaning of Race”) discusses the race concept, the preceding two chapters deal with the taxonomic class immediately above that of race, namely ‘species’.

For sexually-reproducing organisms, ‘species’ is usually defined as the largest group of organisms capable of breeding with one another and producing fertile offspring in the wild.

However, as Baker explains, things are not quite so simple.

For one thing, over evolutionary time, one species transforms into another gradually with no abrupt dividing line where one species suddenly becomes another (p69-72). Hence the famous paradox, Which came first: the chicken or the egg?.

Moreover, in respect of extinct species, it is often impossible to know for certain whether two ostensible ‘species’ interbred with one another (p72-3). Therefore, in practice, the fossils of extinct organisms are assigned to either the same or different species on morphological criteria alone.

This leads Baker to distinguish different species concepts. These include:

  • Species in the paleontological sense” (p72-3);
  • Species in the morphological sense” (p69-72); and
  • Species in the genetical sense”, i.e. as defined by the criterion of interfertility (p72-80).

On purely morphological criteria, Baker questions humanity’s status as a single species:

“Even typical Nordids and typical Alpinids, both regarded as subraces of a single race (subspecies), the Europid, are very much more different from one another in morphological characters—for instance in the shape of the skull—than many species of animals that never interbreed with one another in nature, though their territories overlap” (p97).

Thus, later on, Baker claims:

Even a trained anatomist would take some time to sort out correctly a mixed collection of the skulls of Asiatic jackals (Canis aureus) and European red foxes (vulpes vulpes), unless he had made a special study of the osteology of the Canidae; whereas even a little child, without any instruction whatever, could instantly separate the skulls of Eskimids from those of Lappids” (p427).

That morphological differences between human groups do indeed often exceed those between closely-related but non-interbreeding species of non-human animal has recently been quantitatively confirmed by Vincent Sarich and Frank Miele in their book, Race the Reality of Human Differences (which I have reviewed here, here and here).

However, even if one defines ‘species’ strictly by the criterion of interfertility (i.e. in Baker’s terminology, “species in the genetical sense”) matters remain less clear than one might imagine.

For one thing, there are the phenomena of ring species, such as the herring gull and lesser black-backed gull.

These two ostensible species (or subspecies), both found in the UK, do not interbreed with one another, but each does interbreed with intermediaries that, in turn, interbreed with the other, such that there is some indirect gene-flow between them. Interestingly, the species ranges of the different intermediaries form a literal ring around the Arctic, such that genes will travel around the Artic before passing from lesser black-backed gull to herring gull or vice versa (p76-79).[18]

Indeed, even the ability to produce fertile offspring is a matter of degree. Thus, some pairings produce fertile offspring only rarely.

For example, often, Baker reports, “sterility affects [only] the heterogametic sex [i.e. the sex with two different sex chromosomes]” (p95). Thus, in mammals, sterility is more likely to affect male offspring. Indeed, this pattern is so common that it even has its own name, being known as Haldane’s Rule, after the famous Marxist-biologist JBS Haldane who first noted this pattern.

Other times, Baker suggests, interfertility may depend on the sex of the respective parents. For example, Baker suggests that, whereas sheep may sometimes successfully reproduce with he-goats, rams may be unable to successfully reproduce with she-goats (p95).[19]

Moreover, the fertility of offspring is itself a matter of degree. Thus, Baker reports, some hybrid offspring are not interfertile with one another, but can reproduce with one or other of the parental stocks. Elsewhere, the first generation of hybrids are interfertile but not subsequent generations (p94).

Indeed, though it was long thought impossible, it has recently been confirmed that, albeit only very rarely, even mules and hinnies can successfully reproduce, despite donkeys and horses, the two parental stocks, having, like goats and sheep, a different number of chromosomes (Rong et al 1985; Kay 2002).

Yet, as Darwin observed as far back as 1871 when himself discussing the question as to whether human races are to be regarded as belonging to entirely separate species:

Even a slight degree of sterility between any two forms when first crossed, or in their offspring, is generally considered as a decisive test of their specific distinctness” (The Descent of Man).

Thus, Baker concludes:

There is no proof that hybridity among human beings is invariably eugenesic, for many of the possible crosses have not been made, or if they have their outcome does not appear to have been recorded. It is probable on inductive grounds that such marraiges would not be infertile, but it is questionable whether the hybridity would necessarily be eugenesic. For instance, statistical study might reveal a preponderance of female offpsring” (p97-8).

However, any degree of infertility among human interracial couples is likely to be very slight. After all, today interracial relationships are increasingly common in Britain and America, and not noticeably less fecund than other unions. On the contrary, the number of biracial people, the products of such relationships, are themselves growing precipitously in number in both countries.

In practice, a very slight degree of reduced fertility among phenotypically distinct forms, as might conceivably occur among human interracial couples, would be unlikely to cause biologists to assign the different forms to different species, not least since, in the absense of close study, the slight degree of reduced fertility would probably never be detected in the first place.

Is there then any evidence of reduced fertility among mixed-race couples? Not a great deal.

Today interracial relationships are increasingly common, and not noticeably less fecund than other unions. On the contrary, the number of biracial people, the products of such relationships, are themselves growing precipitously in number in , at least in Britain and America.

On the other hand, possibly blood type incompatibility between mother and developing foetus might be more common in interracial unions due to racial variation in the prevalence of different blood groups.

Also, one study did find a greater prevalence of birth complications, more specifically caesarean deliveries, among Asian women birthing offspring fathered by white men (Nystrom et al 2008).

However, this is a simple reflection of the differences in physical size between whites and Asians, with smaller-framed Asian women having difficulty birthing larger half-white offspring. Thus, the same study also found that white women birthing offspring fathered by Asian men actually have lower rates of caesarean delivery than did women bearing offspring fathered by men of the same race as themselves (Stanford University Medical Center 2008).[20]

Also, one study from Iceland rather surprisingly found that the highest pregnancy rates were found among couples who were actually quite closely related to one another, namely equivalent to third- or fourth-cousins, with less closely related spouses enjoying reduced pregnancy rates (Helgason et al 2008; see also Labouriau & Amorim 2008).

On the other hand, however, David Reich, in Who We Are and How We Got Here reports that, whereas there was evidence of selection against Neanderthal genes in the human genome (that had resulted from ancient hybridization between anatomically modern humans and Neanderthals) owing to the deleterious effects of these genes, there was no evidence of selection against European genes (or African genes) among African-Americans, a racially-mixed population:

In African Americans, in studies of about thirty thousand people, we have found no evidence for natural selection against African or European ancestry” (Who We Are and How We Got Here: p48; Bhatia et al 2014).

This lack of selection against either European-derived (or African-derived) genes in African-Americans suggests that discordant genes did not result in reduced fitness among African-Americans.[21] 

Humans – A Domesticated Species?

A final complication in defining species is that some species of nonhuman animal, wildly recognised as separate species because they do not interbreed in the wild, nevertheless have been known to successfully interbreed in captivity.

A famous example are lions and tigers. While they have never been known to interbreed in the wild, if only because they rarely if ever encounter one another, they have interbred in captivity, producing hybrid offspring in the form of so-called ligers and tigons.

This is, for Baker, of especial relevance to question of human races since, according to Baker, we ourselves are a domesticated species. Thus, he approvingly quotes Blumenbach’s claim that:

Man is ‘of all living beings the most domesticated” (p95).

Thus, with regard to the question of whether humans represent a single species, Baker reaches the following controversial conclusion:

The facts of human hybridity do not prove that all human races are to be regarded as belonging to a single ‘species’. The whole idea of species is vague because the word is used with such different meanings, none of which is of universal application. When it is used in the genetical sense [i.e. the criterion of interfertility] some significance can be attached to it, in so far as it applies to animals existing in natural conditions… but it does not appear to be applicable to human beings, who live under the most extreme conditions of domestication” (p98).

Thus, Baker goes so far as to question whether:

Any two kinds of animals, differing from one another so markedly in morphological characters (and in odour) as, for instance, the Europid and Sanid…, and living under natural conditions, would accept one another as sexual partners” (p97).

Certainly, in our ‘natural environment’ (what evolutionary psychologists call the environment of Evolutionary adaptedness or EEA), many human races would never have interbred, if only for the simple reason that they would never come into contact with one another.

On the contrary, they were separated from one another by the very geographic obstacles (oceans, deserts, mountain-ranges) that reproductively isolated them from one another and hence permitted their evolution into distinct races.

Thus, Northern Europeans surely never mated with sub-Saharan Africans for the simple reason that the former were confined to Northern Europe and surrounding areas while the latter were largely confined to sub-Saharan Africa, such that they are unlikely ever to have interacted.

Only with the invention of technologies facilitating long-distance travel (e.g. ocean-going ships, aeroplanes) would this change.

However, if Northern Europeans never interbred with sub-Saharan Africans, both groups surely did interbreed with their immediate neighbours, who, in turn, interbred with their intermediate neighbours who may, in turn, have interbred indirectly with the other group, since even the Sahara Desert, formerly regarded as the boundary between what were then called the Caucasiod and Negroid races, was far from a complete barrier to gene flow, even in ancient times.

Indeed, there may even have been gene flow between Eurasia and the Americas at the Bering Strait. Only perhaps Australian Aboriginals may to have been completely reproductively isolated for millennia.

There may therefore have been some indirect gene flow even between even distantly related populations as Northern Europeans and sub-Saharan Africans, even if no Nordic European ever encountered, let alone mated with, a black African. This, together with the continuous clinal nature of racial differentiation across the world that resulted from this interbreeding, was the key point emphasized by Darwin in The Descent of Man in support of his conclusion that all human races ought indeed to be considered a single species.

Moreover, Baker’s assertion that modern humans are a domesticated species, although a fashionable viewpoint today, is questionable.

Whether humans can indeed be said to be domesticated depends on how one defines domesticated. If we are domesticated, then humans are surely unique in having domesticated ourselves (or at least one another).[22]

Defining Race

Ultimately then, the question of whether the human race is a single species is a purely semantic dispute. It depends how one defines the word ‘species’.

Likewise, whether human races can be said to exist ultimately depends on one’s definition of the word ‘race’.

Using the word ‘race’ interchangeably with that of ‘subspecies’, Baker provides no succinct definition. Instead, he simply explains:

If two populations [within a species] are so distinct that one can generally tell from which region a specimen was obtained, it is usual to give separate names to the two races” (p99).

Neither does he provide a neat definition of any particular race. On the contrary, he is explicit in emphasizing:

The definition of any particular race must be inductive in the sense that it gives a general impression of the distinctive characters, without professing to be applicable in detail to every individual” (p99).

Is Race Real?

At the conclusion of his chapter on “Hybridity and The Species Question”, Baker seems to reach what was, even in 1974, an incendiary conclusion – namely that, whether using morphological criteria or the criterion of interfertility, it is not possible to conclusively prove that all extant human populations belong to a single species (see above).

Nevertheless, in the remainder of the book, Baker proceeds on the assumption that differences among human groups are indeed subspecific (i.e. racial) in nature and that we do indeed form a single species.

Indeed, Baker criticises the notion that the existence persons of mixed racial ancestry, and the existence of clinal variation between races, disproves the existence of human races by observing that, if races did not interbreed with one another, then they would not be mere different races, but rather entirely separate species, according to the usual definition of this term. Thus, Baker explains:

Subraces and even races sometimes hybridise where they meet, but this almost goes without saying: for if sexual revulsion against intersubracial or interracial marriages were complete, one set of genes would have no chance of intermingling with the other, and the ethnic taxa would be species by the commonly accepted definition. It cannot be too strongly stressed that intersubracial and interracial hybridization is so far from indicating the unreality of subraces and races, that it is actually a sine qua non of the reality of these ethnic taxa” (p12).

This, Baker argues, is because:

It is the fact that intermediaries do occur that defines the race” (p99).

Thus, in nonhuman species among whom subspecies are recognized, there usually exist similar hybrid or intermediary populations around the boundaries of each distinct subspecies. Indeed, this phenomenon is so recurrent that there is even a biological term for it namely intergradation.

Yet this does not cause biologists to conclude that the subspecies in question either do not exist or that their boundaries are somehow arbitrarily delineated and artificial, let alone that subspecies is a biologically meaningless term.

Some people seem to think that, since races tend to blend into one another and hence have blurred boundaries (i.e. what biologists refer to as clinal variation), they do not really exist. Yet Baker objects:

In other matters, no one questions the reality of categories between which intermediaries exist. There is every graduation, for instance, between green and blue, but no one denies these words should be used” (p100).

However, this is perhaps an unfortunate example, since, as psychologists and physicists agree, colours, as such, do not exist.

Instead, the spectrum of light varies continuously. Distinct colours are imposed on this continuous variation only by the human brain and visual system.[23]

Using colour as an analogy for race is also potentially confusing because colour is already often conflated with race. Thus, races are referred to by their ostensible colours (e.g. blacks, whites, browns etc.) and the very word ‘colour’ is sometimes even used as a synonym, or perhaps euphemism, for race, even though, as Baker is at pains to emphasize, races differ in far more than skin colour.

Using colour as an analogy for race differences is only likely to exacerbate this confusion.

Yet Baker’s other examples are similarly problematic. Thus, he writes:

“The existence of youths and human hermaphrodites does not cause anyone to disallow the use of the words, ‘boy’, ‘man’ and ‘woman’” (p100).

However, hermaphrodites, unlike racial intermediaries, are extremely rare. Meanwhile, words such as ‘boy’ and ‘youth’ are colloquial terms, not really scientific ones. As anthropologist John Relethford observes:

We tend to use crude labels in everyday life with the realization that they are fuzzy and subjective. I doubt anyone thinks that terms such as ‘short’, ‘medium’ and ‘tall’ refer to discrete groups, or that humanity only comes in three values of height” (Relethford 2009: p21).

In short, we often resort to vague and impressionistic language in everyday conversation. However, for scientific purposes, we must surely try, wherever possible, to be more precise.

Rather than alluding to colour terms or hermaphrodites, perhaps a better counterexample, if only because it is certain to provoke annoyance, cognitive dissonance and doublethink among leftist race-denying sociologists, is that of social class. Thus, as biosocial criminologist Anthony Walsh demands:

Is social class… a useless concept because of its cline-like tendency to merge smoothly from case to case across the distribution, or because its discrete categories are determined by researchers according to their research purposes and are definitely not ‘pure’” (Race and Crime: A Biosocial Analysis: p6).

However, the same leftist social scientists who insist the race concept is an unscientific social construction, nevertheless continue to employ the concept of social class almost as if it were entirely unproblematic.

However, the objection that races do not exist because races are not discrete categories, but rather have blurred boundaries, is not entirely fallacious.

After all, sometimes intermediaries can be so common that they can no longer be said to be intermediaries at all and all that can be said to exist is continuous clinal variation, such that wherever one chose to draw the boundary between one race and another would be entirely arbitrary.

With increased migration and intermarriage, we may fast be approaching this point.[24]

However, just because the boundaries between racial groups are blurred, this does not mean that the differences between them, whether physiological or psychological, do not exist. To assume otherwise would represent a version of the continuum fallacy or sorties paradox, also sometimes called the fallacy of the heap or fallacy of the beard.

Thus, even if races do not exist, race differences still surely do – and, just as skin colour varies on a continuous, clinal basis, so might average IQbrain-size and personality!

Anticipating Jared Diamond

Remarkably, Baker even manages to anticipate certain erroneous objections to the race concept that had not, to my knowledge, even been formulated at the time of his writing, perhaps because they are so obviously fallacious to anyone without an a priori political commitment to the denying the validity of the race concept.

In particular, Jared Diamond (1994), in an influential and much-cited paper, argues that racial categories are meaningless because, rather than being classified by skin colour, races could just as easily be grouped on the basis of traits such as the prevalence of genes for sickle-cell or lactose tolerance, which would lead us to adopting very different classifications.

Actually, Baker argues, the importance of colour for racial classification has been exaggerated.

In the classification of animals, zoologists lay little emphasis on differences of colour… They pay far more attention to differences in grosser structure” (p159).

Indeed, he quotes no lesser authority than Darwin himself as observing:

Colour is generally esteemed by the systematic naturalist as unimportant (p148).

African_albino
A Negro albino: Proof that race is more than ‘skin deep’

Certainly, he is at pains to emphasise that, among humans, differences between racial groups go far beyond skin colour. Indeed, he observes, one has only to look at an African albino to realize as much:

An albino… Negrid who is fairer than any non-albino European, [yet] appears even more unlike a European than a normal… Negrid” (p160).

Likewise, some populations from the Indian subcontinent are very dark in skin tone, yet they are, according to Baker, predominantly Caucasoid (p160), as, he claims, are the Aethiopid subrace of the Horn of Africa (p225).[25]

Thus, Baker laments how:

An Indian, who may show close resemblance to many Europeans in every structural feature of his body, and whose ancestors established a civilization long before the inhabitants of the British Isles did so, is grouped as ‘coloured’ with persons who are very different morphologically from any European or Indian, and whose ancestors never developed a civilization” (p160).

Yet, in contrast, of the San Bushmen of Southern Africa, he remarks:

The skin is only slightly darker than that of the Mediterranids of Southern Europe and paler than that of many Europids whose ancestral home is in Asia or Africa” (p307).

But no one would mistake them for Caucasoid.

What then of the traits, namely the prevalence of the sickle-cell gene or of lactose tolerance, that would, according to Diamond, produce very different taxonomies?

For Baker, these are what he calls “secondary characters” that cannot be used for the purposes of racial classification because they are not present among all members of any group, but differ only in their relative prevalence (p186).

Moreover, he observes, the sickle-cell gene is likely to have “arisen independently in more than one place” (p189). It is therefore evidence, not of common ancestry, but of convergent evolution, or what Baker refers to as “independent mutation” (p189).

It is therefore irrelevant from the perspective of cladistic taxonomy, whereby organisms are grouped, not on the basis of shared traits as such, but rather of shared ancestry. From the perspective of cladistic taxonomy, shared traits are relevant only to the extent they are (interpreted as) evidence of shared ancestry.

The same is true for lactose tolerance, which seems to have evolved independently in different populations in concert with the development of dairy farming, in a form of gene-culture co-evolution.

Indeed, lactose tolerance appears to have evolved through somewhat different genetic mechanisms (i.e. mutations in different genes) in different populations, seemingly a conclusive demonstration that it evolved independently in these different lineages (Tishkoff et al 2007).

As Baker warns:

One must always be on the lookout for the possibility of independent mutation wherever two apparently unrelated taxa resemble one another by the fact that some individuals in both groups reveal the presence of the same gene” (p189).

In evolutionary biology, this is referred to as distinguishing analogy from homology.

Thus, for example, authors Vincent Sarich and Frank Miele, in their book Race: The Reality of Human Differences (which I have reviewed here, here and here) observe:

There are two groups of people [i.e. races] with the conbination of dark skin and frizzy hair—sub-Saharan Africans and Melanesians. The latter have often been called Oceanic Negroes,’ implying a special relationship with Africans. The blood-group data, however, show that they are about as different from Africans as they could be” (Race: The Reality of Human Differences: p134).

But Diamond’s proposed classification is even more preposterous than these early pre-Darwinian non-cladistic taxonomic schemes, since he proposes to classify races on the basis of a single trait in isolation, the trait in question (either lactose tolerance or the sickle-cell gene) being chosen either arbitrarily or, more likely, to illustrate the point that Diamond is attempting to make.

Yet even pre-Darwinian taxonomies proposed to classify species, not on the basis of a single trait, but rather on the basis of a whole suit of traits that intercorrelate together.

In short, Diamond proposes to classify races on the basis of a single character that has evolved independently in distantly related populations, instead of a whole suite of inter-correlated traits indicative of common ancestry.

Interestingly, a similar error may underlie an even more frequently cited paper by Marxist-geneticist Richard Lewontin, which argued the vast majority of genetic variation was within-group rather than between-group – since Lewontin, like Diamond, also relied on ‘secondary characters’ such as blood-groups to derive his estimates (Lewontin 1972).[26]

The reason for the recurrence of this error, Baker explains, is that:

Each of the differences that enable one to distinguish all the most typical individuals of any one taxon from those of another is due, as a general rule, to the action of polygenes, that is to say, to the action of numerous genes, having small cumulative effects” (p190).

Yet, unlike traits resulting from a few alleles, polygenes are not amenable to simple Mendelian analysis.

Therefore, this leads to the “unfortunate paradox” whereby:

The better the evidence of relationship or distinction between ethnic taxa, the less susceptible are the facts to genetic analysis” (p190).

As a consequence, Baker laments:

Attention is focussed today on those ‘secondary differences’… that can be studied singly and occur in most ethnic taxa, though in different proportions in different taxa… The study of these genes… has naturally led, from its very nature, to a tendency to minimise or even disregard the extent to which the ethnic taxa of man do actually differ from one another” (p534).

Finally, Baker even provides a reductio ad absurdum of Diamond’s approach, observing:

From the perspective of taste-deficiency the Europids are much closer to the chimpanzee than to the Sinids and Paiwan people; yet no one would claim that this resemblance gives a true representation of relationship” (p188).

However, applying the logic of Diamond’s article, we would be perfectly justified and within our rights to use this similarity in taste deficiency in order to classify Caucasians as a sub-species of chimpanzee!

Subraces

The third section of Baker’s book, “Studies of Selected Human Groups”, focusses on the traditional subject-matter of physical anthropology – i.e. morphological differences between human groups.[27]

Baker describes the physiological differences between races in painstaking technical detail. These parts of the book makes for an especially difficult read, as Baker carefully elucidates both how anthropologists measure morphological differences, and the nature and extent of the various physiological differences between the races discussed revealed by these methods.

Yet, curiously, although many of his measures are quantitative in nature, Baker rarely discusses whether differences are statistically significant.[28] Yet without statistical analysis, all of Baker’s reports of quantitative measurements of differences in the shapes and sizes of the skulls and body parts of people of different races represent little more than subjective impressions.

This is especially problematic in his discussion of so-called ‘subraces’ (subdivisions within the major continental races, such as Nordics and the Meditaranean race, both supposed subdivisions within the Caucasiod race), where differences could easily be dismissed as, if not wholly illusory, then at least as clinal in nature and as not always breeding true.

Yet nowhere in his defence of the reality of subracial differences does Baker cite statistics. Instead, his argument is wholly subjective and qualitative in nature:

In many parts of the world where there have not been any large movements of population over a long period, the reality of subraces is evident enough” (p211).

One suspects that, given increased geographic mobility, those parts of the world are now reduced in number.

Thus, even if subracial differences were once real, with increased migration and intermarriage, they are fast disappearing, at least within Europe.

Is ‘White’ a Social Construct?

One other interesting observation may be made with regard to Bakers proposed racial taxonomy. Save when quoting from other earlier authors who did use these terms, Baker himself never once refers to white people or the ‘the white race’. 

Not only does he, as we have seen, reject the use of colour for the purposes of racial classification, he also does not seem to recognize white people as constituting a useful racial category in the first place. Thus, not only do the terms white people’ or the white race’ receive no mention in his racial taxonomy either as a race or a subrace, neither is any synonym covering roughly the same set of people included (p624-5).

Of course, Baker’s Europid race might appear, from its name, to cover much the same ground, since the ancestral homelands of those today classed as white are roughly coextensive with the geographical boundaries of Europe.

In fact, however, its meaning is much broader, as Baker uses the word Europid to refer to what earlier anthropologists more typically called the Caucasian race, and, as he is himself at pains to emphasize, the indigenous inhabitants of North Africa, the Middle East and, at least according to Baker, even South Asia are all classified as Caucasoid/Europid (p160), and Baker even argues that those he terms the Aethiopids of the Horn of Africa are also predominantly Caucasoid/Europid (p225).

While indigenous Europeans are grouped together with North Africans, South Asians and Arabs as Europid, they are also subdivided among themselves into such supposed subraces as Nordid, Mediterranid, Osteuropid, Dinarid and Alpinid. Yet none of these terms is equivalent to what we today habitually call white people , and the indigenous homelands of at least some of these subraces, notably the Mediterranid, extend outside of the European continent into North Africa and the Middle East, and include some peoples whom we would today hesitate to call white, who are unlikely to themselves identify as such, and who would certainly not be recognized as white by most white racists.

This conclusion seems to have been shared by most other early-twentieth century physical anthropologists. For example, Carleton Coon, in The Races of Europe, contended that:

The Mediterranean racial zone stretches unbroken from Spain across the Straits of Gibraltar to Morocco, and thence eastward to India. A branch of it extends far southward on both sides of the Red Sea into southern Arabia, the Ethiopian highlands, and the Horn of Africa (The Races of Europe: p401).

Unlike Baker, Coon does indeed use the phrase the white race’, and indeed regards his 1939 book as a study of this race. However, he clearly intends this phrase to carry a rather broader meaning than that with which it is usually invested today, since he regards, for example, even the Gallas, the Somalis, the Ethiopians, and the inhabitants of Eritrea as all being white or near white”, a view that would hardly endear him to most contemporary white racists (The Races of Europe: p445).   

Thus, while he would certainly reject the idea that race is a mere social construct as preposterous, I suspect that Baker, along with other early twentieth-century racial anthropologists, might actually agree with the race deniers that the concept of a white race, at least as it is defined and demarcated in the Anglosphere today, is indeed an artificial construct with little biological validity, which owes more to geographical and even religious factors (i.e. the traditional boundary between Chistendom and the Islamic world) than it does to measuable phenotypic, or, for that matter, genetic, differences.

In contrast, although the politcally correct orthodoxy holds that terms such as ‘Caucasian’ or ‘Caucasoid’ (or, to use Baker’s preferred term ‘Europid’) reflect a scientifically obsolete and discredited basis for racial classification, this racial category actually seems to have been broadly corroborated by modern studies in population genetics.

Thus, geneticist David Reich, in his 2018 book, Who We Are and How We Got Here, reports:

Today, the peoples of West Eurasia—the vast region spanning Europe, the Near East, and much of central Asia—are genetically highly similar. The physical similarity of West Eurasian populations was recognized in the eighteenth century by scholars who classified the people of West Eurasia as ‘Caucasoids’… The whole-genome data at first seem to validate some of the old categories… Populations within West Eurasia are typically around seven times more similar to one another than West Eurasians are to East Asians. When frequencies of mutations are plotted on a map, West Eurasia appears homogeneous, from the Atlantic façade of Europe to the steppes of central Asia. There is a sharp gradient of change in central Asia before another region of homogeneity is reached in East Asia” (Who We Are and How We Got Here: p93).

This is probably because the term ‘Caucasoid’ was hardly an arbitrary invention of eighteenth- and nineteenth-century racists, but rather reflected, not only real phenotypic resemblance among populations, but also geographic factors, the indigenous homelands of the ostensible race being circumscribed by relatively impassable geographic obstacles – such as the Sahara Desert, Himalayas, Siberia and Atlantic Ocean – which represented barriers to human movement and hence gene flow throughout much of human history and prehistory.

In contrast, the ostensible boundaries of the indigenous homelands so-called ‘white race’ are, at least today, usually equated with the boundaries of the European continent. But, whereas the Sahara, Himalayas, Siberia and Atlantic were long barriers to gene-flow, at least some of the boundaries of the European continent – namely the Mediterranean Sea, Strait of Gibraltar and Turkish Straits – were long hubs of trade, migration, population movement and conquest. It is thus unsurprising that populations on either side of these boundaries, far from being racially distinct, resemble one another both phenotypically and genetically.

Studies of Selected Human Groups

This third section of the book focuses on certain specific selected human populations. These are presumably chosen because Baker feels that they are representative of certain important elements of human evolution, racial divergence, or are otherwise of particular interest.

Unfortunately, Baker’s choice of which groups upon which to focus seems rather arbitrary and he never explains why these groups were chosen ahead of others.

In particular, it is notable that Baker focuses primarily on populations from Europe and Africa. East Asians (i.e. Mongoloids), curiously, are entirely unrepresented.

The Jews

After a couple of introductory chapters, and one chapter focussing on “Europids” (i.e. Caucasians) as a whole, Baker’s next chapter discusses Jewish people.

In the opening paragraphs, he observes that:

In any serious study of the superiority or inferiority of particular groups of people one cannot fail to take note of the altogether outstanding contributions made to intellectual and artistic life, and to the world of commerce and finance, generation after generation by persons to whom the name of Jews is attached” (p232).

However, having taken due “note” of this, and hence followed his own advice, he says almost nothing further on the matter, either in this chapter or in those later chapters that deal specifically with the question of racial superiority (see below).

Instead, Baker first focuses on justifying the inclusion of Jews in a book about race, and hence arguing against the politically-correct notion that Jews are not a race, but rather mere practitioners of a religion.[29] Baker gives short-shrift to this notion:

There is no close resemblance between Judaism in the religious sense and a proselytizing religion such as the Roman Catholic” (p326).

In other words, Baker seems to be saying, because Judaism is not a religion that actively seeks out converts (but rather one that, if anything, discourages conversion), Jews have retained an ethnic character distinct from the host populations alongside whom they reside, without having their racial traits diluted by the incorporation of large numbers of converts of non-Jewish ancestry.

Yet, actually, even proselytizing religions like Christianity, Catholicism and Islam that do actively seek to convert nonbelievers, often come to take on an ethnic character, since, despite the possibility of conversion, offspring usually inherit (i.e. are indoctrinated in) the faith of their parents, apostates are persecuted, conversion remains, in practice, rare, and people are admonished to marry within the faith.

Thus, in polities beset by ethnic conflict, like Northern Ireland, Lebanon or the former Yugoslavia, religions often comes to represent markers for ethnicity or even something akin to ethnicities in and of themselves – i.e. reproductively-isolated, endogamous breeding populations.

Having concluded, then, that there is a racial as well as a religious component to Jewish identity, Baker nevertheless stops short of declaring the Jews a race or even what he calls a subrace.

Dismissing the now discredited Khazar hypothesis in a sentence,[30] Baker instead classes them bulk of the world’s Jewish population (i.e. the Ashkenazim) as merely part of “Armenid subrace” of the Europid race” with some “Orientalid” (i.e. Arab) admixture (p242).[31]

Thus, Baker claims:

Persons of Ashkennazic stock can generally be recognised by certain physical characters that distinguish them from other Europeans” (p238).

Jewish_Nose
Baker’s delightfully offensive illustration of Jewish nose shape, taken from Jacobs (1886).

These include a short but wide skull and a nose that is “large in all dimensions” (p239), the characteristic shape of which Baker even purports to illustrate with a delightfully offensive diagram (p241).[32]

Likewise, Baker claims that Sephardic Jews, the other main subgroup of European Jews, are likewise “distinguishable from the Ashkenazim by physical characters”, being slenderer in build, with straighter hair, narrower noses, and different sized skulls, approximately more to the Mediterranean racial type (p245-6).

But, if Sephardim and Ashkenazim are indeed “distinguishable” or “recognisable” by “physical characters”, either from one another or from other European Gentiles, as Baker claims, then with what degree of accuracy is he claiming such distinctions can be made? Surely far less than 100%.[33]

Moreover, are the alleged physiological differences that Baker posits between Ashkenazi, Sephardi, and other Europeans based on recorded quantitative measurements, and, if so, are the differences in question statistically significant? On this, Baker says nothing.

The Celts

The next chapter concerns The Celts, a term surrounding which there is so much confusion and which has been used in so many different senses – racial, cultural, ethnic, territorial and linguistic (p183) – that some historians have argued that it should be abandoned altogether.

Baker, himself British, is keen to dispel the notion that the indigenous populations of the British Isles were, at the time of the Roman invasion, a primitive people, and is very much an admirer of their artwork.

Thus, Baker writes that:

Caesar… nowhere states that any of the Britons were savage (immanis), nor does he speak specifically of their ignorance (ignorantia), though he does twice mention their indiscretion (imprudentia) in parleying” (p263).

Of course, Caesar, though hardly unbiased in this respect, did regard the indigenous Britons as less civilized than the Romans themselves. However, I suppose that barbarism is, like civilization (see below), a matter of degree.

Regarding the racial characteristics of those inhabitants of pre-Roman Britain who are today called Celts, Baker classifies them as Nordic, writing:

Their skulls scarcely differ from those of the Anglo-Saxons who subsequently dominated them, except in one particular character, namely, that the skull is slightly (but significantly) lower in the Iron Age man than in the Anglo-Saxon” (p257).[34]

Thus, dismissing the politically-correct notion that the English were, in the words of another author, “true multiracial society”, Baker claims:

“[The] Angles, Saxons, Jutes, Normans, Belgics and… Celts… were not only of one race (Europid) but of one subrace (Nordid).” (p267).

Citing remains found in an ancient cemetery in Berkshire supposedly containing the skeletons of Anglo-Saxon males but indigenous British females and hybrid offspring, he concludes that, rather than extermination, a process of intermarriage and assimilation occurred (p266). This is a conclusion largely corroborated by recent population genetic studies.

However, the indigenous pre-Celtic inhabitants of the British Isles were, Baker concludes, less Nordic than Mediterranid in phenotype.[35]

Such influences remain, Baker claims, in the further reaches of Wales and Ireland, as evidenced by the distribution of blood groups and of hair colour.

Thus, whereas the Celtic fringe is usually associated with red, auburn or ginger hair, Baker instead emphasizes the greater prevalence of dark hair among the Irish and Welsh:

The tendency towards the possession of dark hair was much more marked in Wales than in England, and still more marked in the western districts of Ireland” (p265).[36]

This conclusion is based upon the observations of nineteenth century English ethnologist John Beddoe, who travelled the British Isles recording the distribution of different hair and eye colours, reporting his findings in The Races of Britain, which was first published in 1862 and remains, to my knowledge, the only large body of data on the distribution of hair and eye colour in the British Isles to this day.

On this basis, Baker therefore concludes that:

The modern population of Great Britain probably derives mainly from the [insular] ‘Celts’… and Belgae, though a more ancient [i.e. Mediterranean] stock has left its mark rather clearly in certain parts of the country, and the Anglo-Saxons and other northerners made an additional Nordid contribution later on” (p269).

Yet recent population genetic studies suggest that even the so-called Celts, like the later Anglo-Saxons, Normans and Vikings, actually had only a quite minimal impact on the ancestry of the indigenous peoples of the British Isles.[37]

This, of course, further falsifies the politically correct, but absurd notion that the British are a nation of immigrants – which phrase is, of course, itself a recent immigrant from America, in respect of whose population the claim surely has more plausibility.

The Celts, moreover, likely arrived from on the British Isles from continental Europe by the same route as the later Anglo-Saxons and Normans – i.e. across the English channel (or perhaps the south-west corner of the North Sea), by way of Southern England. This is, after all, by far the easiest, most obvious and direct route.[38]

This leads Baker to conclude that the Celts, like the Anglo-Saxons after them, imposed their language on, but had little genetic impact on, the inhabitants of those parts of the British Isles furthest from this point of initial disembarkation (i.e. Scotland, Ireland, Wales). Thus, Baker concludes:

The Iron Age invaders transmitted the dialects of their Celtic language to the more ancient Britons whom they found in possession of the land [and] pushed back these less advanced peoples towards the west and north as they spread” (p264).

But these latter peoples, though adopting the Celtic tongue, were not themselves (primarily) descendants of the Celtic invaders. This leads Baker to conclude, following what he takes to also be the conclusion of Carleton Coon in the latter’s book The Races of Europe, that:

It is these people, the least Celtic—in the ethnic sense—of all the inhabitants of Great Britain, that have clung most obstinately to the language that their conquerors first taught them two thousand years ago” (p269).

In other words, in a racial and genetic, if not a linguistic, sense, the English are actually more Celtic than are the self-styled Celtic Nations of Scotland, Ireland and Wales!

Australian Aboriginals – a “Primitive” Race?

The next chapter is concerned with Australian Aboriginals, or, as Baker classes them, “Australids”.

In this chapter Baker is primarily concerned with arguing that Aboriginals are morphologically primitive.

Of course, the indigenous inhabitants of what is now Australia were, when Europeans first made contact with them, notoriously backward in terms of their technology and material culture.

For example, Australian Aboriginals are said the only indigenous people yet to have developed the simple bow or bow and arrow; while the neighbouring, and related, indigenous people of Tasmania, isolated from the Australian mainland by rising sea levels at the end of the last ice age but usually classed as of the same race, are said to have lacked even, arguably, the ability to make fire.

However, this is not what Baker means by referring to Aboriginals as retaining many “primitive traits. Indeed, unlike his later chapters on black Africans, Baker says nothing regarding the technology or material culture of indigenous Australians.

Instead, he talks exclusively about their morphology. In referring to them as retaining “primitive” characters, Baker is therefore using the word in the specialist phylogenetic sense. Thus, he argues that Australian Aboriginals:

Retain… physical characters that were possessed by remote ancestors but have been lost in the course of evolution by most members of the taxa that are related to it” (p272-3).

In other words, they retain traits characteristic of an earlier state of human evolution which have since been lost in other extant races.

Baker purports to identify twenty-eight such “primitive” characters in Australian aboriginals. These include prognathism (p281), large teeth (p289), broad noses (p282), and large brow ridges (p280).

Baker acknowledges that all extant races retain some primitive characters that have been lost in other races (p302). For example, unlike most other races (but not Aboriginals), Caucasoids retain scalp hair characteristic of early hominids and indeed other extant primates (p297).

However, Baker concludes:

The Australids are exceptional in the number and variety of their primitive characters and in the degree to which some of them are manifested” (p302).

Relatedly, Nicholas Wade observes that, whereas there is a general trend towards lighter and less robust bones and skulls over the course of human evolution, something referred to as gracialization, two populations at “the extremities of the human diaspora” seem to have been exempt, or isolated, from this process, namely Aboriginals and the “Fuegians at the tip of the South America” (A Troublesome Inheritance: p167-8).[39]

Of course, to be morphologically ‘primitive’ in this specialist phylogenetic sense entails no necessary pejorative imputations as are often associated with the word ‘primitive’.

However, some phylogentically primitive traits may indeed be linked to the primitive’ technology of indigenous Aboriginals at the time of first contact with Europeans.

For example, tooth size decreased over the course of human evolution as human invented technologies (e.g. cooking, tools for cutting) that made large teeth unnecessary. As science writer Marek Kohn puts it:

As the brain expanded in the course of becoming human, the teeth became smaller. Hominids lost their built-in weapons, but developed the possibility of building their own, all the way to the Bomb” (The Race Gallery: p63).

Indeed, Darwin himself observed, in The Descent of Man, that:

The early male forefathers of man were, as previously stated, probably furnished with great canine teeth; but as they gradually acquired the habit of using stones, clubs, or other weapons, for fighting with their enemies or rivals, they would use their jaws and teeth less and less. In this case, the jaws, together with the teeth, would become reduced in size” (The Descent of Man).

Therefore, it is possible, Kohn provocatively contends, that:

Aborigines have a biological adaptation to compensate for the primitiveness of their material culture… Teeth get smaller, the argument runs, when technology becomes more advanced” (The Race Gallery: p72-3).

On this view, the relatively large size of Aboriginal teeth could be associated with the primitive state of their technology.

Another phylogentically primitive Aboriginal trait that also, rather more obviously, implies lesser intelligence intelligence, is their relatively smaller brain size.

Indeed, Philippe Rushton posits a direct tradeoff between brain-size and the size of the jaw and teeth, arguing in Race, Evolution and Behavior (which I have reviewed here, here and here) that: 

As brain tissue expanded it did so at the expense of the temporalis muscles, whichclose the jaw. Since smaller temporalis muscles cannot close as large a jaw, jaw size was reduced. Consequently, there is less room for teeth” (Race, Evolution and Behavior: Preface to Third Edition: p20-1).

Thus, leading mid-twentieth century American physical anthropologist and racialist Carleton Coon reports:

The critical differences between [“the ancestors of our living races”] and us lie mostly in brain size versus jaw size – the balance between thinking thoughts and eating foods of various degrees of fineness” (Racial Adaptations: p113).

Thus, Aboriginals have, on average, Baker reports, not only larger jaws and teeth, but also smaller brains than those of Caucasians, weighing only about 85% as much (p292). The smaller average brain-size of Aboriginals is confirmed by more recent data (Beals et al 1984).

Baker also reviews some suggestive evidence regarding the internal structure of Aboriginal brains, as compared to that of Europeans, notably in the relative positioning of the lunate sulcus, again suggesting similarities with the brains of non-human primates.

In this sense, then, Australian Aboriginals ‘primitivebrains may indeed be linked to the primitive state, in the more familiar sense of the word ‘primitive’, of their technology and culture.

San Bushmen and Paedomorphy

Whereas Australian Aboriginals are morphologically “primitive” (i.e. retain characters of early hominids), the San Bushmen of Southern Africa (“Sanids”), together with the related Khoi (collectively Khoisan, or, in racial terms, Capoid) are, Baker contends, paedomorphic.

Bushman_penes
Bushmen’s paedomorphic penes

By this, Baker means that the San people retain into adulthood traits that are, in other taxa, restricted to infants or juveniles, and is more often referred to as neoteny.[40]

One example of this supposed paedomorphy is provided by the genitalia of the Sanid males:

The penis, when not erect, maintains an almost horizontal position… This feature is scarcely ever omitted in the rock art of the Bushmen, in their stylized representations of their own people. The prepuce is very long; it covers the glans completely and projects forward to a point. The scrotum is drawn up close to the root of the penis, giving the appearance that only one testis has descended, and that incompletely” (p319).[41]

Humans in general are known to be neotenous in many of our distinct characters, and we are also, of course, the most intelligent known species.

Indeed, as discussed by Desmond Morris in his 1960s human ethology classic The Naked Ape (which I have reviewed here), among the traits that have been associated with neotenty in humans are our brain size, growth patterns, hairlessness, inventiveness, upright posture, spinal curvature, smaller jaws and teeth, forward facing vaginas, lack of a penis bone, the length of our limbs and the retention of the hymen into adulthood.

However, Baker argues:

Although mankind as a whole is paedomorphous, those ethnic taxa (the Sanids among them) that are markedly more paedomorphious than the rest have never achieved the status of civilization, or anything approaching it, by their own initiative. It would seem that, when carried beyond a certain point, paedomorphosis is antagonistic to purely intellectual advance” (p324).

As to why this might be the case, he speculates in a later chapter:

Certain taxa have remained primitive or become paedomorphous in their general morphological characters and none of these has succeeded in developing a civilization. It is among these taxa in particular that one finds some indication of a possible cause of mental inferiority in the small size of the brain” (p428).

Yet this is a curious suggestion since neoteny is usually associated with increased brain growth in humans.[42]

Moreover, other authorities class East Asians as a paedomorphic race, and Baker himself classes the bulk of the population” of Japan as “somewhat paedomorphious” (p538).[43]

However, the Japanese, along with other Northeast Asians, not least the Chinese, have undoubtedly founded great civilizations and have brains as large as, or, after controlling for body-size, even larger than those of Europeans, and are generally reported to have somewhat higher IQs (see Lynn’s Race Differences in Intelligence: which I have reviewed here).

The Big Butts of Bushmen – or just of Bushwomen?

Bushman_buttocks
Bushwomen’s buttocks (or ‘steatopygia’)

Having discussed male genitalia, Baker also emphasizes the primary and secondary sexual characteristics of Sanid women – in particular their protruding buttocks (“steatopygia”) and alleged elongated labia.

The protruding buttocks of Sanid women are, Baker contends, qualitatively different in both shape and indeed composition from those of other populations, including the much-celebrated ‘big butts’ of contemporary African-Americans (p318).

Thus, whereas, among other populations, the shape of the buttocks, even if very large, are “rounded” in shape:

It is particular characteristic of the Khoisanids that the shape of the projecting part is that of a right-angled triangle, the upper edge being nearly horizontal … [and] internally… consist of masses of fat incorporated between criss-crossed sheets of connective tissue said to be joined to one another in a regular manner.

Regarding the function of these enlarged buttocks, Baker rejects any analogy with the humps of the camel, which evolved as reserves of fat upon which the animal could call in the event of famine or draught.

Unlike camels, which are, of course, adapted to a desert environment, Baker concludes:

The Hottentots, Korana, and Bushmen are not to be regarded as people adapted by natural selection to desert life” (p318).

However, today, San Bushmen are indeed largely restricted to a desert environment, namely the Kalahari desert.

However, although he does not directly discuss this, Baker presumably regards this as a recent displacement, resulting from the Bantu expansion, in the course of which the less advanced San were displaced from their traditional hunting grounds in southern Africa by Bantu agriculturalists, and permitted to eke out an undisturbed existence only in an arid desert environment of no use to Bantu agriculturalists.

Instead of having evolved as fat reserves in the event of famine, drought or scarcity, Baker instead suggests that Khoisan buttocks evolved through sexual selection.

This seems plausible, given the sexual appeal of ‘big butts even among western populations. However, recent research suggest that it is actually lumbar curvature, or lordosis, an ancient mammalian mating signal, rather than fat deposits in the buttocks as such, that is primarily responsible for the perceived attractiveness of so-called ‘big butts’ (Lewis et al 2015).

This sexual selection hypothesis is, of course, also consistent with the fact that large buttocks among the San seem to be largely, if not entirely, restricted to women.

However, Carleton Coon, in Racial Adaptations: A Study of the Origins, Nature, and Significance of Racial Variations in Humans, suggests alternatively that this sexual dimorphism could instead reflect the caloric requirements of pregnancy and lactation.[44]

The caloric demands of pregnancy and lactation are indeed the probable reason women of all races have greater fat deposits than do males.

Indeed, an analogy might be provided by female breasts, since these, unlike the mammary glands of other mammalian species, are present permanently, from puberty on, and, save during pregnancy and lactation, are composed predominantly of fatty tissues, not milk.[45]

Elusive Elongated Labia?

Hottentot apron
The only photographic evidence of the ‘Hottentot apron’?

In addition to their enlarged buttocks, Baker also discusses the alleged elongated labia of Sanid women, sometimes referred to, rather inaccurately in Baker’s view, as the “the Hottentot apron”.

Some writers have discounted this notion as a sort of nineteenth-century anthropological myth. However, Baker himself insists that the elongated labia of the San are indeed real.

His evidence, however, is less than compelling, the illustrations included in the text being limited to a full-body photograph in which the characteristic is barely visible (p311) and what seems to be a surely rather fanciful sketch (p315).

Likewise, although a Google image search produces abundant photographic evidence of Khoisan buttocks, their elongated labia prove altogether more elusive.

Perhaps the modesty of Khoisan women, or the prudery and puritanism of Victorian anthropologists and explorers, prevented the latter from recording photographic evidence for this characteristic.

However, it is perhaps telling that, even in this age of Rule 34 of the Internet (If it exists, there is porn of it. No exceptions), I have been unable to find photographic evidence for this trait.

Racial Superiority

The fourth and final section of ‘Race’ turns to the most controversial topic addressed by Baker in this most controversial of books, namely whether any racial group can be said to be superior or inferior to another, a question that Baker christens “the Ethnic Question”.

He begins by critiquing the very nature of the notion of superiority and inferiority, observing in a memorable and quotable aphorism:

Anyone who accepts it as a self-evident truth, in accordance with the American Declaration of Independence, that all men are created equal may properly be asked whether the meaning of the word ‘equal’ is self-evident” (p421).

Thus, if one is “concerned simply with the question whether the taxa are similar or different”, then, Baker concludes, “there can be no doubt as to the answer” (p421).

Indeed, this much is clear, not simply from the huge amount of data assembled by Baker himself in previous chapters, but also from simple observation.[46]

However, Baker continues:

The words ‘superior’ and ‘inferior’ are not generally used unless value judgements are concerned” (p421).

Any value judgement is, of course, necessarily subjective.

On objective criteria, each race can only be said to be, on average, superior in a specific endeavour (e.g. IQ tests, basketball, mugging, pimping, drug-dealing, tanning, making music, building civilizations). The value to be ascribed to these endeavours is, however, wholly subjective.

On these grounds, contemporary self-styled race realists typically disclaim any association between their theories and any notions of racial superiority.

Yet these race realists are often the very same individuals who emphasise the predictive power of IQ tests in determining many social outcomes (income, criminality, illegitimacy, welfare dependency) which are generally viewed in anything but value-neutral terms (see The Bell Curve: which I have reviewed here).

From a biological perspective, no species (or subspecies) is superior to any other. Each is adapted to its own ecological niche and hence presumably superior at surviving and reproducing within the specific environment in which it evolved.

Thus, sociobiologist Robert Trivers quotes his mentor Bill Druryf as observing during a discussion between the two regarding a possible biological basis for race prejudice:

Bob, once you’ve learnt to think of a herring gull as equal, the rest is easy” (Natural Selection and Social Theory: p57).

However, taken to its logical conclusion, or reductio ad absurdum, this suggests a dung beetle is equal to Beethoven!

From Physiology to Psychology

Although he alludes in passing to race differences in athletic ability, Baker, in discussing superiority, is concerned primarily with intellectual and moral achievement. Therefore, in this final section of the book, he turns from physiological differences to psychological ones.

Of course, the two are not entirely unconnected. All behaviour must have an ultimate basis in the brain, which is itself a part of an organism’s physiology. Thus:

Cranial capacity is, of course, directly relevant to the ethnic problem since it sets a limit to the size of the brain in different taxa; but all morphological differences are also relevant in an indirect way, since it is scarcely possible that any taxa could be exactly the same as one another in all the genes that control the development and function of the nervous and sensory systems, yet so different from one another in structural characters in other parts of the body” (p533-4).

Indeed, Baker observes:

Identity in habits is unusual even in pairs of taxa that are morphologically much more similar to one another than [some human races]. The subspecies of gorilla, for instance, are not nearly so different from one another as Sanids are from Europids, but they differ markedly in their modes of life” (426).

In other words, since human races differ significantly in their physiology, it is probable that they will also differ, to a roughly equivalent degree, in psychological traits, such as intelligence, temperament and personality.

Measuring Superiority?

In discussing the question of the intellectual and moral superiority of different racial groups, Baker focusses on two lines of evidence in particular:

  1. Different races’ performance in ability and attainment tests;
  2. Different races’ historical track record in founding civilizations.

Baker’s discussion of the former topic is now rather dated.

Recent findings unavailable to Baker include the discovery that East Asians score somewhat higher on IQ tests than do white Europeans (see Race Differences in Intelligence: reviewed here), and also that Ashkenazi Jews score higher still (see The Chosen People: review forthcoming).[47]

Evidence has also accumulated regarding the question of the relative contributions of heredity to racial differences in IQ, including the Minnesota transracial study (Scarr & Weinberg 1976; Weinberg et al 1992) and studies of the effects of racial admixture on IQ using blood-group data (Loehlin et al 1973; Scarr et al 1977), and, most recently, genome analysis (Lasker et al 2019). See also my review of Richard Lynns Race Difference in Intelligence: An Evolutionary Perspective’, posted here.

Readers interested in more recent research on this issue should consult Jensen and Rushton (2005) and Nisbett (2005); or Nicholas Mackintosh’s summary in Chapter Thirteen of his textbook, IQ and Human Intelligence (2nd Ed) (pp324-359); or indeed my own recent review of Richard Lynns Race Difference in Intelligence: An Evolutionary Perspective’, posted here.[48]

Criteria for Civilization and Moral Relativism

While his data on race differences in IQ is therefore now dated, Baker’s discussion of the track-record of different races in founding civilizations remains of interest today, if only because this is a topic studiously avoided by most contemporary authors, historians and anthropologists on account of its politically-incorrect nature – though Jared Diamond, in Guns, Germs and Steel, represents an important recent exception to this trend.[49]

The first question, of course, is precisely how one is to define ‘civilizations’ in the first place, itself a highly contentious issue.[50]

Thus, Baker identifies twenty-one criteria for recognising civilizations (p507-8).[51]

In general, these can be divided into two types:

  1. Scientific/technological criteria;
  2. Moral criteria.[52]

However, the latter are inherently problematic. What constitutes moral superiority itself involves a moral judgement that is necessarily subjective.

In other words, whereas technological and scientific superiority can be demonstrated objectively, moral superiority is a mere matter of opinion.

Thus, the ancient Romans, transported to our times, would surely accept the superiority of our technology – and, if they did not, we would, as a consequence of the superiority of our technology, outcompete them both economically and militarily and hence prove it ourselves.

However, they would view our social, moral and political values as decadent and we would have no way of proving them wrong.

Take, for example, Baker’s first requirement for civilization, namely that:

In the ordinary circumstances of life in public places they [i.e. members of the society under consideration] cover the external genitalia and greater part of the trunk with clothes” (p507).

This criterium is not only curiously puritanical, but also blatantly biased against tropical cultures. Whereas in temperate and arctic zones clothing is essential for survival, in the tropics the decision to wear clothing represents little more than an arbitrary fashion choice.

Meanwhile, the requirement that the people in question “do not practice severe mutilation or deformation of the body”, another moral criterion, could arguably exclude contemporary westerners from the ranks of the ranks of the civilized’, given the increasing prevalence of tattooing, flesh tunnel ear plugs and other forms of extreme bodily modification (not to mention gender reassignment surgery and other non-consensual forms of  genital mutilation) – or perhaps it is merely those among us who succumb to such fads who are not truly civilized.

The requirement that a civilization’s religious beliefs not be “purely or grossly superstitious” (p507) is also problematic. As a confirmed atheist, I suspect that all religions are, by very definition, superstitious. If some forms of Buddhism and Confucianism are perhaps exceptions, then they are perhaps simply not religions at all in the western sense.

At any rate, Christian beliefs  regarding miracles, resurrection, the afterlife, the Holy Spirit and so on surely rival those of any other religion when it comes to “gross superstition”.

As for his complaint that the religion of the Mayansdid not enter into the fields of ethics” (p526), a complaint he also raises in respect of indigenous black African religions (p384), contemporary moral philosophers generally see this as a good thing, believing that religion is best kept of moral debates.[53]

In conclusion, any person seeking to rank cultures on moral criteria will, almost inevitably, rank his own society as morally superior to all others – simply because he is judging these societies by the moral standards of his own society that he has internalized and adopted as his own.

Thus, Baker himself views Western civilization as superior to such pre-Columbian mesoamerican civilizations as the Aztecs due to the latter’s practice of mass ritual human sacrifice and cannibalism (p524-5).

However, in doing so, he is judging the cultures in question by distinctly Western moral standards. The Aztecs, in contrast, may have viewed human sacrifice as a moral imperative and may therefore have viewed European cultures as morally deficient precisely because they did not butcher enough of their people in order to propitiate the gods.

Likewise, whereas Baker views cannibalism as incompatible with civilization (p507), I personally view cannibalism as, of itself, a victimless crime. A dead person, being dead, is incapable of suffering by virtue of being eaten. Indeed, in this secular age of environmental consciousness, one might even praise cannibalism as a highly ‘sustainable’ form of recycling.

For this reason, in my own discussion of the different cultures and civilizations founded by members of different races, I will confine my discussion exclusively to scientific and technological criteria for civilization.

Sub-Saharan African Cultures

Baker’s discussion of different groups’ capacity for civilization actually begins before his final section on “Criteria for Superiority and Inferiority” in his four chapters on the race whom Baker terms Negrids – namely, black Africans from south of the Sahara, excluding Khoisan and Pygmies (p325-417).

Whereas his previous chapters discussing specific selected human populations focussed primarily, or sometimes exclusively, on their morphological peculiarities, in the last four of these chapters, focussing on African blacks, his focus shifts from morphology to culture.

Thus, Baker writes:

The physical characters of the Negrids are mentioned only briefly. Members of this race are studied in Chapters 18-21 mainly from the point of view of the social anthropologist interested in their progress towards civilization at a time when they were still scarcely influenced over a large part of their territory, by direct contact with members of more advanced ethnic taxa” (p184).

Unlike some racialist authors,[54] Baker acknowledges the widespread adoption of advanced technologies throughout much of sub-Saharan Africa prior to modern times. However, he attributes the adoption of these technologies to contact with, and borrowings from, outside non-Negroid civilizations (e.g. Arabs, Egyptians, Moors, Berbers, Europeans).

Therefore, in order to distinguish the indigenous, homegrown capacity of black Africans to develop advanced civilization, Baker relies on the reports of seven nineteenth century explorers of what he terms “the secluded area” of Africa, by which term Baker seems to mean the bulk of inland Southern, Eastern and Central Africa, excluding the Horn of Africa, the coast of West Africa and the Gulf of Guinea (p334-5).[55]

In these parts of Africa, at the time these early European explorers visited the continent, the influence of outside civilizations was, Baker reports, “non-existent or very slight” (p335). The cultural practices observed by these explorers therefore, for Baker, provide a measure of black Africans indigenous capacity for social, cultural and technological advancement.

On this perhaps dubious basis, Baker thus concludes that there is no evidence black Africans in this area ever:

  • Fully domesticated any plants (354-6) or animals (p373-7); or
  • Invented, or adopted, the wheel (p373); or other ‘mechanical’ devices with interacting parts (p354).[56]

Also largely absent throughout ‘the secluded area’, according to Baker, were:

In respect of these last two indices of civilization, however, Baker admits a couple of partial, arguable exceptions, which he discusses in the next chapter (Chapter 21). These include the ruins of Great Zimbabwe (p401-9) and a script invented in the nineteenth century (p409-11).[57]

Domesticated Plants and Animals in Africa

Let’s review these claims in turn. First, it certainly seems to be true that few if any species of either animals or plants were domesticated in what Baker calls the “the secluded area” of sub-Saharan Africa.[58]

However, with respect to plants, there may be a reason for this. Many important, early domesticates were annuals. These are plants that complete their life-cycle within a single year, taking advantage of predictable seasonal variations in the weather.

As explained by Jared Diamond, annual plants are ideal for human consumption, and for domestication, because:

Within their mere one year of life, annual plants inevitably remain small herbs. Many of them instead put their energy into producing big seeds, which remain dormant during the dry season and are then ready to sprout when the rains come. Annual plants therefore waste little energy on making inedible wood or fibrous stems, like the body of trees and bushes. But many of the big seeds… are edible by humans. They constitute 6 of the modern world’s 12 major crops” (Guns, Germs and Steel: p136).

Yet sub-Saharan Africa, being located closer to the equator, experiences less seasonal variation in climate. As a result, relatively fewer plants are annuals.

However, it is far less easy to explain why sub-Saharan Africans failed to domesticate any wild species of animal, with the possible exception of guineafowl.[59]

After all, Africa is popular as a tourist destination today in part precisely because it has a relative abundance of large wild mammals of the sort seemingly well suited for domestication.[60]

Jared Diamond argues that the African zebra, a close relative of other wild equids that were domesticated, was undomesticable because of its aggression and what Diamond terms its nasty disposition” (Guns, Germs and Steel: p171-2).[61]

However, this is unconvincing when one considers that Eurasians succeeded in domesticating such formidably powerful and aggressive wild species as wolves and aurochs.[62]

Thus, even domesticated bulls remain a physically-formidable and aggressive animal. Indeed, they were favoured adversaries in blood sports such as bullfighting and bull-baiting for precisely this reason.

However, the wild auroch, from whom modern cattle derive, was undoubtedly even more formidable, being, not only larger, more muscled and with bigger horns, but also surely even more aggressive than modern bulls. After all, one of the key functions of domestication is to produce more docile animals that are more amenable to control by human agriculturalists.[63]

Compared to the domestication of aurochs, the domestication of the zebra would seem almost straight forward. Indeed, the successful domestication of aurochs in ancient times might even cause us to reserve our judgement regarding the domesticability of such formidable African mammals as hippos and African buffalo, the possibility of whose domestication Diamond dismisses a priori as preposterous.

Certainly, the domestication of the auroch surely stands as one of the great achievements of ancient Man.

Reinventing the Wheel?

Baker also seems to be correct in his claim that black Africans never invented the wheel.

However, it must be borne in mind that the same is also probably true of white Europeans, who, rather than independently inventing the wheel for themselves, had the easier option of simply copying the design of the wheel from other civilizations and peoples, namely those from the Middle East, probably Mesopotamia, where the wheel seems to be have first been developed

Indeed, most cultures with access to the wheel never actually invented it themselves, for the simple reason that it is far easier to copy the invention of a third-party through simple reverse engineering than to independently invent afresh an already existing technology all by oneself.

This then explains why the wheel has actually been independently invented, at most, only a few times in history.

The real question, then, is not why the wheel was never invented in sub-Saharan Africa, but rather why it failed to spread throughout that continent in the same way it did throughout Eurasia.

Thus, if the wheel was known, as Baker readily acknowledges it was, in those parts of sub-Saharan Africa that were in contact with outside civilizations (notably in the Horn of Africa), then this raises the question as to why it failed to spread elsewhere in Africa prior to the arrival of Europeans. This indeed is acknowledged to remain a major enigma within the field of African history and archaeology (Law 2011; Chavez et al 2012).

After all, there are no obvious insurmountable geographical barriers preventing the spread of technologies across Africa other than the Sahara itself, and, as Baker himself acknowledges, black Africans in the ‘penetrated’ area had proven amply capable of imitating technological advances introduced from outside.

Why then did the wheel not spread across Africa in the same way it did across Eurasia? Is it possible that African people’s alleged cognitive deficiencies were responsible for the failure of this technology to spread and be copied, since the ability to copy technologies through reverse engineering itself requires some degree of intellectual ability, albeit surely less than that required for original innovation?

One might argue instead that the African terrain was unsuitable for wheeled transport. However, one of the markers of a civilization is surely its very ability to alter the terrain by large, cooperative public works engineering projects, such as the building of roads.

Thus, most of Eurasia is now suitable for wheeled transport in large part only because we, or more specifically our ancestors, have made it so.

Another explanation sometimes offered for the failure of African to develop wheeled transportation is that they lacked a suitable draft animal, horses in sub-Saharan Africa nbeing afflicted with sleeping sickness spread by the tsetse fly.

However, as we have seen above, Baker argues a race’s track record in successfully domesticating wild animals is itself indicative of the intellectual ability and character of that race. For Baker, then, the failure of sub-Saharan African to successfully domesticate any suitable species of potential draft animal (e.g. the zebra: see above) is itself indicative of, and a factor in, their inability to successfully develop advanced civilization.

At any rate, even in the absence of a suitable draft animal, wheels are still useful.

On the one hand, they can be used for non-transport-related purposes (e.g. the spinning wheel, the potter’s wheel, even water wheels). Indeed, in Eurasia the invention of the potter’s wheel is actually thought to have preceded the use of wheels for the purposes of transportation.

Moreover, even in the absence of a suitable draft animal, wheels remain very useful for transportation purposes e.g. wheelbarrows, pulled rickshaws

In other words, humans can themselves be employed as a draft animal, whether by choice or by force, and, if there is one arguable marker for civilization for which Africa did not lack, and which did not await introduction by Europeans, Moors and Arabs, it was, of course, the institution of slavery.

African Writing Systems?

What then of the alleged failure of sub-Saharan Africans to develop a system of writing? Baker refers to only a single writing system indigenous to sub-Saharan Africa, namely the Vai syllabary, invented in what is today Liberia in the nineteenth century in imitation of foreign scripts. Was this indeed the only writing system indigenous to sub-Saharan Africa?

Of course, writing has long been known in North Africa, and ancient Egypt even lays claim to have invented the first written script, namely hieroglyphs, although most archaeologists believe that they were beaten to the gun, once again, by Mesopotamia, with its cuneiform script.

However, this is obviously irrelevant to the question of black African civilization, since the populations of North Africa, including the ancient Egyptians, were largely Caucasoid.[64]

Thus, the Sahara Desert, as a relatively impassable obstacle to human movement throughout most of human history and prehistory (a geographic filter”, according to Sarich and Miele) that hence impeded gene flow, has long represented, and to some extent still represents, the boundary between the Caucasoid and Negroid races (Race: The Reality of Human Differences: p210).

What then of writing systems indigenous to sub-Saharan Africa? The wikipedia entry on writing systems of Africa lists several indigenous African writing systems of sub-Saharan Africa.

However, save for those of recent origin, almost all of these writing systems seem, from the descriptions on their respective wikipedia pages, to have been restricted to areas outside of ‘the secluded area’ of Africa as defined by Baker (p334-5).

Thus, excluding the writing systems of North Africa (i.e. Meroitic, Tifinagh and  ancient Egyptian hieroglyphs), Geze seems to have been restricted to the area around the Horn of Africa; Nsibidi to the area around the Gulf of Guinea in modern Nigeria; Adrinka to the coast of West Africa, while the other scripts mentioned in the entry are, like the Vai syllabary, of recent origin.

The only ancient writing system mentioned on this wikipedia page that was found in what Baker calls ‘the secluded area’ of Africa is Lusona. This seems to have been developed deep in the interior of sub-Saharan Africa, in parts of what is today eastern Angola, north-western Zambia and adjacent areas of the Democratic Republic of the Congo. Thus, it is almost certainly of entirely indigenous origin.

However, Lusona is described by its wikipedia article as only an ideographic tradition, that function[s] as mnemonic devices to help remember proverbs, fables, games, riddles and animals, and to transmit knowledge”.

It therefore appears to fall far short of a fully developed script in the modern sense.

Indeed, the same seems to be true, albeit to a lesser extent, of most of the indigenous writing systems of sub-Saharan Africa listed on the wikipedia page, namely Nsibidi and Adrinka, which each seem to represent only a form of proto-writing.

Only Geze seems to have been a fully-developed script, and this was used only in the Horn of Africa, which not only lies outside ‘the secluded area’ as defined by Baker, but whose population is, again according to Baker, predominantly Caucasoid (p225).

Also, Geze seems to have developed from an earlier Middle Eastern script. It is therefore not of entirely indigenous African origin.

It therefore seems to indeed be true that sub-Saharan Africans never produced a fully-developed script in those parts of Africa where they developed beyond the influence of foreign empires.

However, it must here be emphasized that the same is again probably also true of indigenous Europeans.

Thus, as with the wheel, Europeans themselves probably never independently invented a writing system, the Latin alphabet being derived from Greek script, which was itself developed from the Phoenician alphabet, which, like the wheel, first originated in the Middle East.[65]

Indeed, most writing systems were developed, if not directly from, then at least in imitation of, pre-existing scripts. Like the wheel, writing has only been independently reinvented afresh a few times in history.[66]

The question, then, as with the wheel, is, not so much why much of sub-Saharan Africa failed to invent a written script, but rather why those written scripts that were in use in certain parts of the continent south of the Sahara,  nevertheless failed to spread or be imitated over the remainder of that continent.

African Culture: Concluding Thoughts

In conclusion, it certainly seems clear that much of sub-Saharan Africa was indeed backward in those aspects of technology, social structure and culture which Baker identifies as the key components of civilization. This much is true and demands an explanation.

However, blanket statements regarding the failure of sub-Saharan Africans to develop a writing system or two-storey buildings seem, at best, a misleading simplification.

Indeed, Baker’s very notion of what he calls ‘the secluded area’ of Africa is vague and ill-defined, and he never provides a clear definition, or, better still, a map precisely delineating what he means by the term (p334-5).

Indeed, the very notion of a ‘secluded area’ is arguably misconceived, since even relatively remote and isolated areas of the continent that did not have any direct contact with non-Negroid peoples, will presumably have had some indirect influence from outside of sub-Saharan Africa, if only by contact with peoples from those regions of the continent south of the Sahara which had been influenced by foreign peoples and civilizations.

After all, as we have seen, Europeans also failed to independently develop either the wheel and writing system for themselves, instead simply copying these innovations from the neighbouring civilizations of the Middle East.

Why then were black Africans south of the Sahara, who were indeed exposed to these technologies in certain parts of their territory, nevertheless unable to do the same?

Perhaps one factor impeding the movement of technologies such as the wheel and writing across sub-Saharan Africa in pre-modern times is the relative lack of navigable waterways (e.g. rivers) in the region.

As emphasized by Tim Marshall in his book Prisoners of Geography, rivers in sub-Saharan African tended to be non-navigable, mainly because of the prevalence of large waterfalls that made transport by river a dangerous venture.

Since, in ancient and premodern times, transport by river was, at least in Eurasia, generally easier, safer and quicker than by land, Africas generally non-navigable river system may have ironically impeded the spread throughout Africa even of technologies that were themselves of use primarily for transportation, such as the wheel.

Pre-Columbian Native American Cultures

Baker’s discussion of status of the pre-Columbian civilizations, or putative civilizations, of America is especially interesting. Of these, the Mayans definitely stand out, in Baker’s telling, as the most impressive in terms of their scientific and technological achievements.

Baker ultimately concludes, however, that even the Maya do not qualify as a true civilization, largely on moral grounds – namely, their practice of mass sacrifices and cannibalism.

Yet, as we have seen, this is to judge the Mayans by distinctly western moral standards

No doubt if western cultures were to be judged by the moral values of the Mayans, we too would be judged just as harshly. Perhaps they would condemn us precisely for not massacring enough of our citizens in order to propitiate the gods.

However, even seeking to rank the Mayans based solely on their technological and scientific achievements, they still represent something of a paradox.

On the one hand, their achievements in mathematics and astronomy seem impressive.

Indeed, Baker educates us that it is was Mayans, not the Hindus or Arabs more often credited with the innovation, who first invented the concept of zero – or rather, to put the matter more precisely, “invent[ed] a ‘local value’ (or ‘place notational’) system of numeration that involved zero: that is to say, a system in which the value of each numberical symbol depended on its position in a series of such symbols, and the zero, if required, took its place in this series ” (p552).

Thus, Baker writes:

The Maya had invented the idea [of zero] and applied it to their vegisimal system [i.e. using a base of twenty] before the Indian mathematicians had thought of it and used it in denary [i.e. decimal] notation” (p522).[67]

Thus, Baker concludes:

The mathematics, astronomy, and calendar of the Middle Americans suggest unqualified acceptance into the ranks of the civilized” (p525).

However, on the other hand, according to Baker’s account:

They had no weights… no metal-bladed hoes or spades and no wheels (unless a few toys were actually provided with wheels and really formed part of the Mayan culture)” (p524).

Yet, as Baker alludes to in his rather disparaging reference to “a few toys”, it now appears the these toys were indeed part of the Maya culture.

Thus, far from failing to invent the wheel, Native Americans are one of the few peoples in the world with an unambiguous claim to have indeed invented the wheel entirely independently, since the possibility of wheels being introduced through contact with Eurasian civilizations is exceedingly remote.

Thus, the key question is, not why Native American civilizations failed to invent the wheel, for they did indeed invent the wheel, but rather why they failed to make full use of this remarkably useful invention, seemingly only employing it for seemingly frivolous items resembling toys (but whose real purpose is unknown) rather than for transport, or indeed the production of ceramics, textiles or energy.

Terrain may have been a factor. As mentioned above, one of the markers of a true civilization is arguably its very ability to alter its terrain by large-scale engineering projects such as the building of roads. However, obviously some terrains pose greater difficulties in this respect, and the geography of much of Mesoamerica is particularly uninviting.

As in respect of sub-Saharan Africa, another factor sometimes cited is the absence of a suitable draft animal.

The Inca, but not the Aztecs and Maya, did have the llama. However, llama are not strong enough to carry humans, or to pull large carts.

Of course, for Baker, as we have seen above, a races track record in domesticating non-human animals, including for use as draft animals, is itself indicative of that races ability and capacity to build and maintain advanced civilization.

However, in the Americas, most large wild mammals of the sort possibly suited for domestication as a draft animal were wiped out by the first humans to arrive on the continent, the former having evolved in complete isolation from humans, and hence being completely evolutionarily unprepared for the sudden influx of humans with their formidable hunting skills.[68]

Thus, Jared Diamond in Guns Germs and Steel observes:

Ancient Native Mexicans invented wheeled vehicles with axles for use as toys, but not for transport. This seems incredible to us until we reflect that ancient Mexicans lacked domestic animals to hitch to their wheeled vehicles, which therefore offered no advantage over human porters” (Guns Germs and Steel: p248).

However, it is simply not true that, in the absence of a draft animal, wheels vehicles offered no advantage over human porters”, as claimed by Diamond. On the contrary, as dicussed above, humans themselves can be employed as draft animals (e.g. the wheelbarrow and pulled rickshaw), and, as Diamond himself observes in a later chapter:

Human-powered wheelbarrows… enabled one or more people, still using just human muscle power, to transport much greater weights than they could have otherwise” (Guns Germs and Steel: p359).

Moreover, as again discussed above, the wheel also has other uses besides transport, one of which, the potter’s wheel, actually seems to have been adopted before the use of wheels for transportation purposes in Europe. Yet there is no evidence for the use of the potter’s wheel in the Americas prior to the arrival of Europeans. 

As for the Mayan script, this was also, according to Baker, quite limited. Thus, Baker reports:

There was no way of writing verbs, and abstract ideas (apart from number) could not be inscribed. It would not appear that the technique even of the Maya lent itself to a narrative form, except in a very limited sense. Most of the Middle Americans conveyed non-calendrical information only by speech or by the display of a series of paintings” (p524).

Indeed, he reports that “nearly all their inscriptions were concerned with numbers and the calendar” (p524).

The Middle Americans had nothing that could properly be called a narrative script” (p523-4).

Baker vs Diamond: The Rematch

However, departing from Baker’s conclusions, I regard the achievements of the Mesoamerican civilizations as, overall, quite impressive.

This is especially so, not only when one takes into account, not only their complete isolation from the Old World civilizations of Eurasia, but also of other factors identified by Jared Diamond in his rightly-acclaimed Guns, Germs and Steel.

Thus, whereas the Eurasian cultural zone is oriented largely on an east-to-west axis, spreading from China and Japan in the East, to western Europe and North Africa in the West, America is a tall, narrow continent that spreads instead from north-to-south, quite narrow in places, especially at the Isthmus of Panama, where the North American continent meets South America, which, at the narrowest point, is less than fifty miles across. 

As Diamond emphasizes, because climate varies with latitude (i.e. distance from the equator), this means that different parts of the Americas have very different climates, making the movement and transfer of crops, domesticated animals and people much more difficult.

This, together with the difficulty of the terrain, might explain why even the Incas and Aztecs, though contemporaneous, seem to have been largely if not wholly unaware of one another’s existence, and certainly had no direct contact.

As a result, Native American cultures developed, not only in complete isolation from Old World civilizations, but even largely in isolation even from one another.

Moreover, the Americas had few large domesticable mammals, almost certainly because the first settlers of the continent, on arriving, hunted them to extinction on first arrival, and the mammals, having evolved in complete isolation from humans, were entirely unprepared for the arrival of humans, with their formidable hunting skills, to whom they were wholly unadapted.

In these conditions, the achievements of the Mesoamerican civilizations, especially the Mayans, seem to me quite impressive, all things considered – certainly far more impressive than the achievements of, say, sub-Saharan Africans or Australian Aboriginals.

This is especially so in comparison to sub-Saharan Africa when one takes into consideration the fact that the latter region was neither completely isolated from Eurasian civilizations nor as narrowly oriented on a north-west as are the Americas.

Thus, as has been emphasized by astrophysicist Michael Hart in his book, Understanding Human History, Diamond’s theory is a rather more successful explanation for the technological backwardness and underdevelopment of the pre-Columbian Americas than it is for the even greater technological backwardness and underdevelopment of sub-Saharan Africa.

Thus, if these black Africans and Australian Aboriginals can then indeed be determined to possess lesser innate intellectual capacity as compared to, say, Europeans or East Asians, then I feel it is nevertheless premature to say the same of the indigenous peoples of the Americas.

Artistic Achievement

In addition to ranking cultures on scientific, technological and moral criteria, Baker also assesses the quality of their artwork (p378-81; p411-17; p545-549). However, judgements of artistic quality, like moral judgements, are necessarily subjective.

Thus, Baker disparages black African art as non-naturalistic (p381) yet also extols the decorative art of the Celtics, which is mostly non-figurative and abstract (p261-2).

However, interestingly, with regard to styles of music, Baker recognises the possibility of cultural bias, suggesting that European explorers, looking for European-style melody and harmony, failed to recognise the rhythmical qualities of African music which are, Baker remarks, perhaps unequalled in the music of any other race of mankind (p379).[69]

A Reminder of What Was Possible”?

The fact that Race’ remains a rewarding some read forty years after first publication, is an indictment of the hold of politically-correctness over both science and the publishing industry.

In the intervening years, despite all the advances of molecular genetics, the scientific understanding of race seems to have progressed but little, impeded by political considerations.

Meanwhile, the study of morphological differences between races seems to have almost entirely ceased, and a worthy successor to Baker’s ‘Race’, incorporating the latest genetic data, has, to my knowledge, yet to be published.

At the conclusion of the first section of his book, dealing with what Baker calls “The Historical Background”, Baker, bemoaning the impact of censorship and what would today be called political correctness and cancel culture on both science and the publishing industry, recommends the chapter on race from a textbook published in 1928 (namely, Contemporary Sociological Theories by Pitirim Sorokin) as “well worth reading”, even then, over forty years later, if only “as a reminder of what was still possible before the curtain went down” (p61).

Today, some forty years after Baker penned these very words and as the boundaries of acceptable opinion have narrowed yet further, I recommend Baker’s ‘Race’ in much the same spirit – as both an historical document and “a reminder of what was possible”.

__________________________

Endnotes

[1] For example, anthopologist-geneticist Vincent Sarich and science writer Frank Miele, in their book Race: The Reality of Human Differences (which I have reviewed here and here), provide a good example from the history of race science of where the convergent evolution of similar traits among different human lineages was once mistaken for evidence of homology and hence of shared ancestry, when they write:

There are two groups of people with the combination of dark skin and frizzy hairsub-Saharan Africans and Melanesians. The latter have often been called ‘Oceanic Negroes,’ implying a special relationship with Africans. The blood group data, however, showed that they are about as different from Africans as they could be” (Race: The Reality of Human Differences: p134)

Genetic studies often allow us distinguish homology from analogy, because the same or similar traits in different populations often evolve through different genetic mutations. For example, Europeans and East Asians evolved lighter complexions after leaving Africa, in part, by mutations in different genes (Norton et al 2007). Similarly, lactase persistence has evolved through mutations in different genes in Europeans than among some sub-Saharan Africans (Tishkoff et al 2009). Of course, at least in theory, the same mutation in the same gene could occur in different populations, thus providing an example of convergent evolution and homoplasy even at the genetic level. However, this is unlikely, and, with the analysis of a large number of genetic loci, especially in non-coding DNA, where mutations are unlikely to be selected for or against and hence are lost or retained at random in different populations, is unlikely to lead to errors in determining the relatedness of populations. 

[2] In his defence, the Ainu are not one of the groups upon whom Baker focuses in his discussion, and are only mentioned briefly in passing (p158; p173; p424) and at the very end of the book, in his “Table of Races and Subraces”, where he attempts to list, and classify by race, all the groups mentioned in the book, howsoever briefly (p624-5).

[3] Although we no longer need to rely on morphological criteria in order to determine the relatedness between populations, differences between racial groups in morphology and bodily structure remain an interesting, and certainly a legitimate, subject for scientific study in their own right. Unfortunately, however, the study and measurement of such differences seems to have all but ceased among anthropologists. One result is that much of the data on these topics is quite old. Thus, HBDers, Baker included, are sometimes criticized for citing studies published in the nineteenth and early-twentieth century. In principle, there is, however, nothing wrong with citing data from the nineteenth or early-twentieth century, unless critics can show that the methodology adopted have subsequently been shown to be flawed. However, it must be acknowledged that the findings of such studies with respect to morphology may no longer apply to modern populations, as a result of recent population movements and improvements in health and nutrition, among other factors. At any rate, the reason for the paucity of recent data is the taboo associated with such research.

[4] This is a style of formatting I have not encountered elsewhere. It makes it difficult to bring oneself to skip over the material rendered in smaller typeface since it is right there in the main body of the text, and indeed Baker himself claims that this material is “more technical and more detailed than the rest (but not necessarily less interesting)” (pix).

[5] Yet another source of potential terminological confusion results from the fact that, as will be apparent from many passages from the book quoted in this review, Baker uses the word “ethnic” to refer to differences that would better to termed “racial” – i.e. when referring to biologically-inherited physical and morphological differences between populations. Thus, for example, he uses the term “ethnic taxon” as “a comprehensive term that can be used without distinction for any of the taxa that are minor to species: that is to say, races, subraces and local forms” (p4). Similarly, he uses the phrase “the ethnic problem” to refer to the “whole subject of equality and inequality among the ethnic taxa of man” (p6). However, as Baker acknowledges, “English words derived from the Greek ἔθνος (ethnic, ethnology, ethnography, and others) are used by some authors in reference to groups of mankind distinguished by cultural or national features, rather than descent from common ancestors” (p4). However, in defending his adoption of this term, he notes “this usage is not universal” (p4). This usage has, I suspect, become even more prevalent in the years since the publication of Bakers book. However, in my experience, the term ethnic’ is sometimes also used as politically correct euphemism for the word ‘race’, both colloquially and in academia.

[6] In both cases, the source of potential confusion is the same, since both terms, though referring to a race, are derived from geographic terms (Europe and the Caucasus region, respectively), yet the indigenous homelands of the races in question are far from identical to the geographic region referred to by the term. The term Asian, when used as an ethnic or racial descriptor, is similarly misleading. For example, in British-English, Asian, as an ethnic term, usually refers to South Asians, since South Asians form a larger and more visible minority ethnic group in the UK than do East Asians. However, in the USA, the term Asian is usually restricted to East Asians and Southeast Asians – i.e. those formerly termed Mongoloid. The British-English usage is more geographically correct, but racially misleading, since populations of the Indian subcontinent, like those from the Middle East (also part of the Asian continent) are actually genetically closer to southern Europeans than to East Asians and were generally classed as Caucasian by nineteenth and early-twentieth century anthropologists, and are similarly classed by Baker himself. This is one reason that the term Mongoloid, despite pejorative connotations, remains useful.

[7] Moreover, the term Mongoloid is especially confusing given that it has also been employed to refer to people suffering from a developmental disability and chromosomal abnormality (Down Syndrome), and, while both usages are dated, and the racial meaning is actually the earlier one from which the later medical usage is derived, it is the latter usage which seems, in my experience, to retain greater currency, the word ‘Mongoloid’ being sometimes employed as a rather politically-incorrect insult, implying a mental handicap. Therefore, while I find annoying the euphemism treadmill whereby terms once quite acceptable terms (e.g. ‘negro’, ‘coloured people’) are suddenly and quite arbitrarily deemed offensive, the term ‘Mongoloid’ is, unlike these other etymologically-speaking, quite innocent terms, understandably offensive to people of East Asian descent given this dual meaning.

[8] For example, another ethnonym, Asian, is also etymologically problematic. Thus, the word Asia, the source of the ethnonym, Asian, derives from the Greek Ἀσία, which originally referred only to Anatolia, at the far western edge of what would now be called Asia, the inhabitants of which region are not now, nor have ever likely been, Asian in the current American sense. Indeed, the very term Asia is a Eurocentric concept, grouping together many diverse peoples, fauna, flora and geographic zones, and whose border with Europe is quite arbitrary. Another even more etymologically suspect ethonym is, of course, the term Indian (and its derivatives ‘Amerindian’, ‘Red Indian’ and ‘American Indian’) when applied to the indigenous peoples of the Americas.

[9] The main substantive differences between the rival taxonomies of different racial theorists reflect the perennial divide between lumpers and splitters. There is also the question of precisely where the line is to be drawn between one race and another in clinal variation between groups, and whether a hybrid or clinal population sometimes constitutes a separate race in and of itself.

[10] For example, in Nicholas Wade’s A Troublesome Inheritance, this history of the misuse of the race concept comes in Chapter Two, titled ‘Perversions of Science’; in Philippe Rushton’s Race, Evolution and Behavior: A Life History Perspective (which I have reviewed here, here and here), this historical account is postponed until Chapter Five, titled ‘Race and Racism in History’; in Jon Entine’s Taboo: Why Black Athletes Dominate Sports and Why We’re Afraid to Talk About it, it is delayed until Chapter Nine, titled ‘The Origins of Race Science’; whereas, in Sarich and Miele’s Race: The Reality of Human Differences (which I have reviewed here, here and here), these opening chapters discussing the history of racial science expand to fill almost half the entire book.

[11] Indeed, somewhat disconcertingly, even Hitler’s Mein Kampf is taken seriously by Baker, the latter acknowledging that “the early part of [Hitler’s] chapter dealing with the ethnic problem is quite well-written and not uninteresting” (p59) – or perhaps this is only to damn with faint praise.

[12] Thus, at the time Stoddard authored The Rising Tide of Color Against White World-Supremacy in 1920, with a large proportion of the world under the control of European colonial empires, a contemporary observer might be forgiven for assuming that what Stoddard called White World-Supremacy, was a stable, long-term, if not permanent arrangement. However, Stoddard accurately predicted the demographic transformation of the West, what some have termed The Great Replacement or A Third Demographic Transition, almost a century before this began to become a reality.

[13] The exact connotations of this passage may depend on the translation. Thus, other translators translate the passage that Manheim translates as The mightiest counterpart to the Aryan is represented by the Jew instead as The Jew offers the most striking contrast to the Aryan”, which alternative translation has rather different, and less flattering, connotations, given that Hitler famously extols the Aryan as the master race. The rest of the passage quoted remains, when taken in isolation, broadly flattering, however.

[14] To clarify, both Boas and Montagu are briefly mentioned in later chapters. For example, Boass now largely discredited work on cranial plasticity is discussed by Baker at the end of his chapter on ‘Physical Differences Between the Ethnic Taxa of Man: Introductory Remarks’ (p201-2). However, this is outside of Baker’s chapters on “The Historical Background”, and therefore Boas’s role in (allegedly) shaping the contemporary consensus of race denial is entirely unexplored by Baker. For discussion on this topic, see Carl Degler’s In Search of Human Nature; see also Chapter Two of Kevin Macdonald’s The Culture of Critique (which I have reviewed here) and Chapter Three of Sarich and Miele’s Race: The Reality of Human Differences (which I have reviewed here, here and here).

[15] Thus, there was no new scientific discovery that presaged or justified the abandonment of biological race as an important causal factor in the social and behavioural sciences. Later scientific developments, notably in genetics, were certainly later co-opted in support of this view. However, there is no coincidence in time between these two developments. Therefore, whatever the true origins of the theory of racial egalitarianism, whether one attributes it to horror at the misuse of race science by the Nazi regime, or the activism of certain influential social scientists such as Boas and Montagu, one thing is certain – namely, the abandonment, or at least increasing deemphasis, of the race category in the social and behavioural sciences was originally motivated by political rather than scientific considerations. See Carl Degler’s In Search of Human Nature; see also Chapter 2 of Kevin Macdonald’s Culture of Critique (which I have reviewed here) and Chapter Three of Sarich and Miele’s Race: The Reality of Human Differences (which I have reviewed here, here and here).

[16] That OUP gave up the copyright is, of course, to be welcomed, since it means, rather than gathering dust on the shelves of university libraries, while the few remaining copies still in circulation from the first printing rise in value, it has enabled certain dissident publishing houses to release new editions of this now classic work.

[17] Baker suggests that, at the time he wrote, behavioural differences between pygmy chimpanzees and other chimpanzees had yet to be demonstrated (p113-4). Today, however, pygmy chimpanzees are known to differ behaviourally from other chimps, being, among other differences, less prone to intra-specific aggression and more highly sexed. However, they are now usually referred to as bonobos rather than pygmy chimpanzees, and are recognized as a separate species from other chimpanzees, rather than a mere subspecies.

[18] This is, at least, how Baker describes this species complex and how it was traditionally understood. Researching the matter on the internet, however, suggests whether this species complex represents a true ring species is a matter of some dispute (e.g. Liebers et al 2006).

[19] In cases of matings between sheep and goats that result in offspring, the resulting offspring themselves are usually, if not always, infertile. Moreover, actually, according to the wikipedia page on the topic, the question of when sheep and goats can ever successfully interbreed is more complex than suggested by Baker.

[20] I have found no evidence to support the assertion in some of the older nineteenth-century literature that women of lower races have difficulty birthing offspring fathered by European men, owing to the greater brain- and head-size of European infants. Summarizing this view, contemporary Russian racialist Vladimir Avdeyev, in his impressively encyclopaedic, if extremely racist and occassionally slightly bonkers book, Raciology: The Science of the Hereditary Traits of Peoples, claims:

The form of the skull of a child is directly connected with the characteristics of the structure of the mother’s pelvis—they should correspond to each other in the goal of eliminating death in childbirth. The mixing of the races unavoidably leads to this, because the structure of the pelvis of a mother of a different race does not correspond to the shape of the head of [the] mixed infant; that leads to complications during childbirth” (Raciology: p157).

Thus, Avdeyev claims, owing to race differences in brain size:

Women on lower races [sic] endure births very easily, sometimes even without any pain, and only in highly rare cases do they die from childbirth. But this can never be said of women of lower races [sic] who birth children of white fathers” (Raciology: p157).

Thus, he quotes an early-twentieth century Russian race theorist as claiming:

American Indian women… often die in childbirth from pregnancies with a child of mixed blood from a white father, whereas pure-blooded children within them are easily born. Many Indian women know well the dangers [associated with] a pregnancy from a white man, and therefore, they prefer a timely elimination of the consequence of cross-breeding by means of fetal expulsion, in avoidance of it” (Raciology: p157-8).

This, interestingly, accords with the claim of infamous late-twentieth century race theorist J Philippe Rushton, in the ‘Preface to the Third Edition’ of his book Race, Evolution and Behavior (which I have reviewed here, here and here), that, as compared to whites and Asians, blacks have narrower hips, giving them a more efficient stride”, which provides an advantage in many athletic events, and that:

The reason Whites and East Asians have wider hips than Blacks, and so make poorer runners, is because they give birth to larger brained babies” (Race, Evolution and Behavior: p11-12).

Thus, Rushton explains elsewhere:

Increasing brain size [over the course of hominid evolution] was associated with a broadening of the pelvis. The broader pelvis provides a wider birth canal, which in turn allows for delivery of larger-brained offspring” (Odyssey: My Life as a Controversial Evolutionary Psychologist: p284-5).

However, contrary to the claim of Avdeyev, I find no support from contemporary delivery room data for the claim that women from so-called lower-races’ experience greater birth complications, and mortality rates, when birthing offspring fathered by European males.
On the contrary, it is only differences in overall body-size, not brain-size, that seem to be the key factor, with East Asian women having greater difficulties birthing offspring fathered by European males because of the smaller frames of East Asian women, even though East Asians have brains as large as or larger than those of Europeans
 (Nystrom et al 2008).
Neither is it true that, where inter-racial mating has not occurred, then, on account of the small brain-size of their babies:

Women on lower races [sic] endure births very easily, sometimes even without any pain, and only in highly rare cases do they die from childbirth(Raciology: p157).

On the contrary, data from the USA actually seems to indicate a somewhat higher rate of caesarean delivery among African-American women as compared to white American women (Braveman et al 1995; Edmonds et al 2013; Getahun et al 2009; Valdes 2020).

[21] Any selection would presumably be against the European-derived component of the African-American genome, since African-Americans are of predominantly black African ancestry. It is therefore possible that selection against the (possibly) deleterous European component of their genome was offset by other advantages possibly accruing to African-Americans with increased European ancestry (e.g. the increased intelligence supposedly associated with increased levels of European ancestry, or the social benefits formerly associated with lighter skin tone or a more Caucasoid phsyiognomy).
Examining the effects of interracial hybridization on other traits besides fertility, there are mixed results. Thus, one study reported what the authors interpreted as a hybrid vigour effect on g-factor of general intelligence among the offspring of white-Asian unions in Hawaii, as compared to the offspring of same-race couples matched for educational and occupational levels (Nagoshi & Johnson 1986). Similarly, Lewis (2010) attributed the higher attractiveness ratings accorded to the faces of mixed-race people to heterosis. Meanwhile, another study found that height was positively correlated with the distance between the birthplaces of one’s parents, itself presumably a correlate of their relatedness (Koziel et al 2011).
On the other hand, however, behavioural geneticist Glayde Whitney suggests that hybrid incompatibility may explain the worse health outcomes, and shorter average life-spans, of African Americans as compared to whites in the contemporary USA, owing to the formers mixed African and European ancestry (Whitney 1999). One specific negative health outcome for some African-Americans resulting from a history racial admixture is also suggested by Helgadottir et al (2006). On the other hand, the disproportionate success of African-Americans in professional athletics hardly seems indicative of impaired health.
It is notable that, whereas recent studies tend to emphasize the (supposed) positive genetic effects resulting from interracial unions, the older literature tends to focus on (supposed) negative effects of interracial hybridization (see Frost 2020). No doubt this reflects the differing zeitgeister of the two ages (Provine 1976; Khan 2011c).
At any rate, even assuming that it can be shown that mixed-race people either enjoy improved health outcomes as compared to monoracial people as a consequence of hybrid vigour, or impaired health outcomes due to outbreeding depression, this is not generally regarded as directly relevant to the question of whether the different human races are to be regarded as separate species. As Darwin wrote:

The inferior vitality of mulattoes is spoken of in a trustworthy work as a well-known phenomenon; but this is a different consideration from their lessened fertility; and can hardly be advanced as a proof of the specific distinctness of the parent races… The common mule, so notorious for long life and vigour, and yet so sterile, shews how little necessary connection there is in hybrids between lessened fertility and vitality” (The Descent of Man).

[22] To clarify, some other domestic species have also been described as having self-domesticated. In particular, a currently popular theory of dog domestication holds that, rather than humans adopting and domesticating wolves, wolves effectively domesticated themselves, by scavenging around human campfires to feed themselves, the tamer, less aggressive and less fearful wolves enjoying greater success in this endeavour, and hence coming to predominate.
However, although, in a sense, a form of self-domestication, this process would still have involved wolves habituating themselves to, and becoming tolerated by, and tolerant to, a different species to themselves, namely humans. In contrast, theories of human self-domestication involve humans interacting only with members of the same species, namely other humans. 

[23] Interestingly, while languages and cultures vary in the number of colours that they recognise and have words for, both the ordering of the colours recognised, and the approximate boundaries between different colours, seems to be cross-culturally universal. Thus, some languages have only two colour terms, which are always equivalent to ‘light’ and ‘dark’. Then, if a third colour terms is used, it is always equivalent to ‘red’. Next come either ‘green’ or ‘yellow’. Experimental attempts to teach colour terms not matching the familiar colours show that individuals learn these terms much less quickly than they do the colour familiar terms recognised in other languages. This, of course, suggests that our colour perception is both innately programmed into the mind and cross-culturally universal (see Berlin & Kay, Basic Color Terms: Their Universality and Evolution). 

[24] Indeed, as I discuss later, with respect to what Baker calls subraces, we may already have long previously passed this point, at least in Europe and North America. While morphological differences certainly continue to exist, at the aggregate, statistical level, between populations from different regions of Europe, there is such overlap, such a great degree of variation even within families, and the differences are so fluid, gradual and continuous, that I suspect such terms as the Nordic race, Alpine race, Mediterranid race and Dinaric race have likely outlived whatever usefulness they may once have had and are best retired. The differences are now best viewed as continuous and clinal.

[25] While Ethiopians and other populations from the Horn of Africa are indeed a hybrid or clinal population, representing an intermediate position between Caucasians and other black Africans, Baker perhaps goes too far in claiming:

Aethiopids (‘Eastern Hamites’ or ‘Erythriotes’) of Ethiopia and Somaliland are an essentially Europid subrace with some Negrid admixture (p225).

Thus, summarizing the findings of one study from the late-1990s, Jon Entine reports:

Ethiopians [represent] a genetic mixture of about 60 percent African and 40 percent Caucasian” (Taboo: Why Black Athletes Dominate Sports And Why We’re Afraid To Talk About It: p115)

The study upon which Entine based this conclusion looked only at mitochondrial DNA and Y chromosome data. More recent studies have incorporated autosomal DNA as well. However, while eschewing terms such as Caucasian’, such studies broadly confirm that there exist substantial genetic affinities between populations from the Horn of Africa and the Middle East (e.g. Ali et al 2020Khan 2011aKhan 2011bHodgson 2014).

[26] Thus, Lewontin famously showed that, when looking at individual genetic loci, most variation is within a single population, rather than between populations, or between races (Lewontin 1972). However, when looking at phenotypic traits that are caused by polygenes, it is easy to see that there are many such traits in which the variation within the group does not dwarf that between groups – for exampe, differences in skin colour as between Negroes and Nordics, or differences in stature between as Pygmies and even neighbouring tribes of Bantu. This is a point emphasized by Sarich and Miele in Race: The Reality of Human Differences (which I have reviewed here).

[27] In addition to discussing morphological differences between races, Baker also discusses differences in scent (170-7). This is a particularly emotive issue, given the negative connotations associated with smelling bad. However, given the biochemical differences between races, and the fact that even individuals of the same race, even the same family, are distinguishable by scent, it is inevitable that persons of different races will indeed differ in scent, and unsurprising that people would generally prefer the scent of their own group. There is substantial anecdotal evidence that this is indeed the case. In general, Baker reports that East Asians have less body odour, whereas both Caucasoids and blacks have greater body odour. Partly this is explained by the relative prevalence of dry and wet ear wax, which is associated with body odour, varies by population and is one of the few easily detectable phenotypic traits in humans that is determined by simply Mendelian inheritance (see McDonald, Myths of Human Genetics). Intriguingly, Nicholas Wade speculates that dry earwax, which is associated with less strong body-odour, may have evolved through sexual selection in colder climates where, due to the cold, more time is spent indoors, in enclosed spaces, where body odour is hence more readily detectable, and producing less scent may have conferred a reproductive advantage (A Troublesome Inheritance: p91). This may explain some of the variation in the prevalence of dry and wet ear wax respectively, with dry earwax predominating only in East Asia, but also being found, albeit to a lesser degree, among Northern Europeans. On the other hand, however, although populations inhabiting colder climates may spend more time indoors, populations inhabiting tropical climates might be expect to sweat more due to the greater heat and hence build up greater bodily odour.

[28] A few exceptions include where Baker discusses the small but apparently statistically significant differences between the skulls of ‘Celts’ and Anglo-Saxons (p257), and where he mentions statistically significant differences between ancient Egypian skulls and those of Negroes (p518).

[29] Baker does, however, acknowledge that:

Some Jewish communities scattered over the world are Jews simply in the sense that they adhere to a particular religion (in various forms); they are not definable on an ethnic basis” (p246).

Here, Baker has in mind various communities that are not either Ashkenazi or Sephardic (or Mizrahi), such as the Beta Israel of Ethiopia, the Lemba of Southern Africa and the Kaifeng Jews resident in China. Although Baker speaks of communities”, the same is obviously true of recent converts to Judaism

[30] Thus, of the infamous Khazar hypothesis, now almost wholly discredited by genetic data, but still popular among some anti-Zionists, because it denies the historical connection between (most) contemporary Jews and the land of Israel, and among Christian anti-Semites, because it denies that the Ashkenazim are indeed chosen people’ of the Old Testament, Baker writes:

It is clear they [the Khazars] were not related, except by religion, to any modern group of Jews” (p34).

[31] Baker thus puts the intellectual achievements of the Ashkenazim in the broader context of other groups within this same subrace, including the Assyrians, Hittites and indeed Armenians themselves. Thus, he concludes:

The contribution of the Armenid subrace to civilization will bear comparison with that of any other” (p246-7).

Some recent genetic studies have indeed suggested affinities between Ashkenazim and Armenian populations (Nebel et al 2001; Elhaik 2013).

[32] In Baker’s defence, the illustration in question is actually taken from the work of a Jewish anthropologist, Joseph Jacobs (Jacobs 1886). Jacobs findings this topic are summarized in this entry in the 1906 Jewish Encyclopedia, entitled Nose, authored by Jacobs and Maurice Fishberg, another Jewish anthropologist, which reports that the ‘hook nose’ stereotypically associated with Jewish people is actually found in only a minority of European Jews (Jacobs & Fishberg 1906).
However, such noses do seem to be more common among Jews than among at least some of the host populations among whom they reside. The
wikipedia article on Jewish noses cites this same entry from the Jewish Encyclopaedia as suggesting that the prevalence of this shape of nose is actually no greater among Jews than among populations from the Mediterranean region (hence the supposed similar shape of so-called Roman noses). However, the Jewish Encyclopaedia entry itself does not actually seem to say any such thing. Instead, it reports only that:

“[As compared with] non-Jews in Russia and Galiciaaquiline and hook-noses are somewhat more frequently met with among the Jews” (Jacobs & Fishberg 1906). 

The entry also reports that, measured in terms of their nasal index, “Jewish noses… are mostly leptorhine, or narrow-nosed” (Jacobs & Fishberg 1906). Similarly, Joseph Jacobs reports in On the Racial Characteristics of Modern Jews’:

Weisbach‘s nineteen Jews vied with the Patagonians in possessing the longest nose (71 mm.) of all the nineteen races examined by him … while they had at the same time the narrowest noses (34 mim)” (Jacobs 1886).

This data, suggesting that Jewish noses are indeed long but are also very narrow, contradicts Baker’s claim that the characteristic Ashkenazi nose is “large in all dimensions [emphasis added]” (p239). However, such a nose shape is consistent Jews having evolved in an arid desert environment, such as the Nagev or other nearby deserts, or in the Judean mountains, where the earliest distinctively Jewish settlements are thought to have developed. Thus, anthropologist Stephen Molnar writes:

Among desert and mountain peoples the narrow nose is the predominant form” (Human Variation: Races, Types and Ethnic Groups: p196).

As Baker himself observes, the nose width characteristic of a population correlates with both the temperature and humidity of the environment in which they evolved (p310-311). This is known as Thomson’s nose rule and is thought to reflect the need to warm and moisturize air before it enters the lungs in cold and dry conditions respectively.
However, interestingly, Baker reports that the correlations are much weaker among the indigenous populations of the American continent (p311). Presumably this is because humans only relatively recently populated that continent
, and therefore have yet had sufficient time to become wholly adapted to the different environments in which they find themselves.
A further factor affecting nose width is jaw size. This might explain why Australian Aboriginals have extremely wide noses despite much of the Australian landmass being dry and arid, since Aboriginals also have very large jaws (Human Variation: Races, Types and Ethnic Groups: p196).
However, it is fallacious to believe that most Australian Aborigines lived in the arid outback prior to the arrival of Europeans and their resulting displacement. In fact, prior to the arrival of Europeans, Aboriginals were probably concentrated in the same more fertile (and much less arid) areas where most white European settlers are today today themselves concentrated, since the same areas which are conducive for agriculture and settlement today also tended to provide more game and vegetation for foragering groups.

[33] Hans Eysenck refers in his autobiography to a study supposedly conducted by one of his PhD students that he claims demonstrated statistically that people, both Jewish and Gentile, actually perform at no better than chance when attempting to distinguish Jews from non-Jews, even after extended interaction with one another (Rebel with a Cause: p35). However, since he does not cite a source or reference for this study, it was presumably unpublished, and must be interpreted with caution.
Eysenck himself, incidentally, was of closeted 
half-Jewish ancestry, practising what antiSemite Kevin Macdonald calls crypsis, which may be taken to suggest he was not entirely disinterested with regard to to question of the extent to which Jews can be recognized by sight. 
The only other study I have found addressing the quite easily researchable, if politically incorrect, question of whether some people can or cannot identify Jews from non-Jews on the basis of phenotypic differences is Andrzejewski et al (2009).

[34] This is one of the few occasions in the book where I recall Baker actually mentioning whether the morphological differences between racial groupings that he describes are statistically significant.

[35] Interestingly, Stephen Oppenheimer, in his book Origins of the British, posits a link between the so-called Celtic regions of the British Isles and populations from one particular area of the Mediterranean, namely the Iberian peninsula, especially the Basques, themselves probably the descendants of the original pre-Indo-European inhabitants of the Iberian peninsula (see Oppenheimer 2006; see also Blood of the Isles).
This seemingly corroborates the otherwise implausible mythological account of the peopling of Ireland provided in Lebor Gabála Érenn, which claims that the last major migration to, and invasion of, Ireland, from which movement of people the modern Irish primarily descend, arrived from Spain in the form of the Milesians. This mythological account may derive from the similarity between the Greek and Latin words for the two regions, namely Iberia and Hibernia respectively, and between the words Gael and Galicia, and the belief of some ancient Roman writers, notably Orosius and Tacitus, that Ireland lay midway between Britain and Spain (Carey 2001).
However, while some early population genetic studies were indeed interpreted to suggest a connection between populations from Iberia and the British Isles, this interpretation seems to have been largely been discredited by more recent research.

[36] Actually, the position with regard to hair and eye colour is rather more complicated. On the one hand, hair colour does appear to be darkest in the ostensibly Celtic’ regions of the British Isles. Thus, Carleton Coon in his 1939 book, The Races of Europe, reports that, with regard to hair colour:

England emerges as the lightest haired of the four major divisions of the British Isles, and Wales as the darkest” (The Races of Europe: p385).

Likewise, Coon reports, that in Scotland:

“Jet black hair is commoner in the western highlands than elsewhere, and is statistically correlated with the greatest survival of Gaelic speech” (The Races of Europe: p387).

However, patterns of eye colour diverge from and complicate this picture. Thus, Coon reports:

“Whereas the British are on the whole lighter-haired than the Irish, they are at the same time darker-eyed” (The Races of Europe: p388).

Indeed, contrary to the notion of the Irish as a people with substantial Mediterranean racial affinities, Coon claims:

There is probably no population of equal size in the world which is lighter eyed, and bluer eyed, than the Irish” (The Races of Europe: p381).

On the other hand, the Welsh, in addition to being darker-haired than the English, are also darker-eyed, with a particularly high prevalence of dark eyes being found in certain more isolated regions of Wales (The Races of Europe: p389).
Interestingly, as far back as the time of the Roman Empire, the Silures, a Brittonic tribe occupying most of South-East Wales and known for their fierce resistance to the Roman conquest, were described by Roman writers Tacitus and Jordanes (the Romans themselves being, of course, a Mediterranean people) as “swarthy” in appearance and as possessing black curly hair.
The same is true of the, also until recently Celtic-speaking, Cornish people, who are, Coon reports, the darkest eyed of the English” (The Races of Europe: p389). Dark hair is also more common in Cornwall (The Races of Europe: p386). Cornwall is, Coon therefore reports, the darkest county in England(The Races of Europe: p396). (However, with the historically unprecedented mass migration of non-whites into the UK in the latter half of the twentieth century and beyond, this is, of course, no doubt no longer true.)
Yet another complicating factor is the prevalence of red hair, which is also associated with the Celtic’ regions of the British Isles, but is hardly a Mediterranean character, and which, like dark hair, reaches its highest prevalence in Wales (The Races of Europe: p385). Baker, for his part, does not dwell on this point, but does acknowledge
, “there is rather a high proportion of people with red hair in Wales”, something for which, he claims “no satisfactory explanation… has been provided” (p265).
However, Baker is skeptical regarding the supposed association of the ancient Celts with ginger or auburn hair. He traces this belief to a single casual remark of Tacitus. However, he suggests that the Latin word used rutilai is actually better translated as red (inclining to golden yellow), and was, he observes, also used to refer to the Golden Fleece and to gold coinage (p257). 

[37] The genetic continuity of the British people is, for example, a major theme of Stephen Oppenheimer’s The Origins of the British (see also Oppenheimer 2006). It is also a major conclusion of Bryan Sykes’s Blood of the Isles, which concludes:

We are an ancient people, and though the [British] Isles has been the target of invasion and opposed settlement from abroad ever since Julius Caesar first stepped onto the shingle shores of Kent, these have barely scratched the topsoil of our deep rooted ancestry” (Blood of the Isles: p338).

However, population genetics is an extremely fast moving science, and recent research has revised this conclusion, suggesting a replacement of around 90% of the population of the British Isles, albeit in very ancient times (around 2000BCE) associated with the spread of the Bell Beaker culture and Steppe-related ancestry, presumably deriving from the Indo-European expansion (Olalde et al 2018). Also, recent population genetic studies suggest that the Anglo-Saxons actually made a greater genetic contribution to the ancestry of the English, especially those from Eastern England, than formerly thought (e.g. Martiniano et al 2016; Schiffels et al 2016).

[38] However, in The Origins of the British, Stephen Oppenheimer proposes an alternative route of entry and point of initial disembarkation, suggesting that the people whom we today habitually refer to as ‘Celts’ arrived, not from Central Europe as traditionally thought, but rather up the Atlantic seaboard from the west coasts of France and Iberia. This is consistent with some archaeological evidence (e.g. the distribution of passage graves) suggesting longstanding trade and cultural links up the Atlantic seaboard from the Mediterranean region, through the Basque country, into Brittany, Cornwall, Wales and Ireland. This would also provide an explanation for what Baker claims is a Mediterranid component in the ancestry of the Welsh and Irish, as supposedly evidenced in distribution of blood groups and the prevalence dark hair and eye colours as recorded by Beddoe.

[39] Interestingly, in addition to gracialization having occurred least, if at all, in Fuegians and Aboriginals, Wade also reports that:

Gracialization of the skull is most pronounced in sub-Saharan Africans and East Asians, with Europeans retaining considerable robustness (A Troublesome Inheritance: p167).

This is an exception to what Steve Sailer calls ‘Rushton’s Rule of Three (see here) and, given that Wade associates gracialization with domestication and pacification (as well as with neoteny), suggests that, at least by this criteria, Europeans evince less evidence of pacification and domestication than do black Africans. This is perhaps a surprising finding given that domestication and pacification among humans are usually associated with the rise of civilization, yet, according to Baker himself, civilization was largely absent from sub-Saharan Africa prior to the arrival of Europeans (see discussion above).

[40] Actually, the meaning of the two terms is subtly different. ‘Paedomorphy’ refers to the retention of juvenile or infantile traits into adulthood. ‘Neoteny refers to one particular process whereby this end-result is achieved, namely slowing some aspects of physiological development. However, ‘paedomorphy’ can also result from another process, namely progenesis’, where, instead, some aspects of development are actually sped up, such that the developing organism reaches sexual maturity earlier, before reaching full maturity in other respects. In humans, most examples of paedomorphy result from the former process, namely ‘neoteny.

[41] These genitalia, of course, contrast with those of neighbouring Negroids, at least according to popular stereotype. For his part, Baker accepts the stereotype that black males have large penes. However, he cites no quantitative data, remarking only:

That Negrids have large penes is somtimes questioned, but those who doubt it are likely to change their minds if they will look at photographs 8, 9, 20, 23, 29, and 37 in Bernatzig’s excellently illustrated book Zwischen Weissem Nil und Belgisch-Kongo’. They represent naked male Nilotids and appear convincing” (p331).

But five photos, presumably representing just five males, hardly represents a convincing sample size. (I found several of the numbered pictures online by searching for the book’s title, and each showed only a single male.) Interestingly, Baker is rightly skeptical regarding claims of differences in the genitalia between European subraces, given the intimate nature of the measurements required, writing:

It is difficult to obtain measurements of theses parts of the body and statements about subracial differences in them must not be accepted without confirmation” (p219).

[42] Interestingly, in their book Big Brain: The Origins and Future of Human Intelligence, neuroscientists Gary Lynch and Richard Granger devote considerable discussion to a supposedly extinct species of hominid, termed Boskop Man or alternatively the Boskopoid race, who, they claim, possessed, as compared to other hominid species (ourselves included), extremely large brains, paedomorphic traits and some physical resemblence to living San Bushmen. However, anthropologist-blogger John Hawks has critiqued this claim in a blog post where he argues that the Boskops are no longer recognized as a distinct species (or subspecies) of hominid and also that the cranial capacity of those remains formerly identified as Boskop, though certainly large, has nevertheless been exaggerated. In this, Hawks cites Singer (1958), who argues that those skulls identified as Boskops’ should instead be classified as Khoisan, from whom they were formerly distinguished solely on the basis of their brain size. However, as Baker suggests, living San Bushmen have very small brains as compared to other extant human races, at least according to data cited by Richard Lynn in his book, Race Differences in Intelligence (reviewed here).

[43] Indeed, the claim that East Asians are especially paedomorphic or neotenized as compared to other races is not restricted to researchers in the racialist or hereditarian tradition. On the contrary, anthropologist  Ashley Montagu, though an early pioneer in race denial, nevertheless conceded one racial difference, namely:

The Mongoloid skull generally, whether Chinese or Japanese, has been rather more neotenized than the Caucasoid or European” (Growing Young: p18).

Similarly, no lesser leftist champion of racial egalitarianism than infamous scientific fraud and charlatan Stephen Jay Gould conceded:

It is hard to deny that Mongoloids… are the most paedomorphic of human groups (Ontogeny and Phylogeny: p134).

Interestingly, Gould made this concession in the context of arguing against the notion that the greater paedomorphosis of Caucasoids as compared to Negroids was indicative of the intellectual superiority of the former. Yet, since there is now widespread agreement among hereditarians that East Asians (but curiously not South-East Asians) score rather higher in IQ tests than do Caucasoids, his observations are actually supportive of both the link between paedomorphosis and encephalization and the hereditarian hypothesis with respect to to race differences in intelligence.
Perhaps recognizing this, in a later book Gould, while still acknowledging that Orientals, not whites, are clearly the most neotenous of human races”, rather contradicted himself just a couple of sentences later by also asserting:

The whole enterprise of ranking groups by degree of neoteny is fundamentally unjustified” (Mismeasure of Man: p121).

[44] Thus, anthropologist Carleton Coon, in Racial Adaptations: A Study of the Origins, Nature, and Significance of Racial Variations in Humans, does not even consider sexual selection as an explanation for the evolution of Khoisan steatopygia, despite their obviously dimorphic presentation. Instead, he proposes:

“[Bushman’s] famous steatopygia (fat deposits that contain mostly fibrous tissue) may be a hedge against scarce nutrients and draught during pregnancy and lactation” (Racial Adaptations: p105). 

[45] Others, however, notably Desmond Morris in The Naked Ape (which I have reviewed here and here), have implicated sexual selection in the evolution of the human female’s permanent breasts. The two hypotheses are not, however, mutually exclusive. Indeed, they may be complementary. Thus, Nancy Etcoff in Survival of the Prettiest (which I have reviewed here and here) proposes that breasts may be perceived as attractive by men precisely because they honestly advertise the presence of the fat reserves needed to sustain a pregnancy” (Survival of the Prettiest: p187). By analogy, the same could, of course, also be true of fatty buttocks.

[46] Thus, Baker demands rhetorically:

Who could conceivably fail to distinguish between a Sanid and a Europid, or between an Eskimid [Eskimo] and a Negritid [Negrito], or between a Bambutid (African Pygmy) or an Australid [Australian Aboriginal]?

[47] Baker does discuss the performance of East Asians on IQ tests, but his conclusions are ambivalent (p490-492). He concludes, for example, “the IQs of Mongolid [i.e. East Asian] children in North America are generally found to be about the same as those of Europids” (p490). Yet recent studies have revealed a slight advantage for East Asians in general intelligence. Baker also mentions the relatively higher scores of East Asians on tests of spatio-visual ability, as compared to verbal ability. However, he attributes this to their lack of proficiency in the language of their host culture, as he relied mostly on American studies of first and second-generation immigrants, or the descendants of immigrants, who were often raised in non-English-speaking homes, and hence only learnt English as a second-language (p490). However, recent studies suggest that East Asians score relatively lower on verbal ability, as compared to their scores on spatio-visual ability, even when tested in a language in which they are wholly proficient (see Race Differences in Intelligence: reviewed here).

[48] Rushton and Jensen (2005) favour the hereditarian hypothesis vis a vis race differences in intelligence, and their presentation of the evidence is biased somewhat in this direction. Nisbett’s rejoinder therefore provides a good balance, being very much biased in the opposite direction. Macintosh’s chapter is perhaps more balanced, but he still clearly favours an environmental explanation with regard to population differences in intelligence, if not with regard to individual differences. My own post on the topic is, of course, naturally enough, the most thorough and balanced treatment of this topic.

[49] Indeed, in proposing tenable environmental-geographical explanations for the rise and fall of civilizations in different parts of the world, Jared Diamond’s Guns, Germs and Steel represents a substantial challenge to Baker’s conclusions in this chapter and the two books are well worth reading together. Another recent work addressing the question of why civilizations rise and fall among different races and peoples, but reaching less politically-correct conclusions, is Michael Hart’s Understanding Human History, which seems to have been conceived of as a rejoinder to Diamond, drawing heavily upon, but also criticizing the former work.

[50] Interestingly, Baker quotes Toynbee as suggesting that:

An ‘identifying mark’ (but not a definition) [of] civilization might be equated with ‘a state of society in which there is a minority of the population, however small, that is free from the task, nor merely of producing food, but of engaging in any other form of economic activities-e.g. industry or trade” (p508).

Yet a Marxist would view this, not as a marker of civilization, but rather of exploitation. Those free from engaging in economic activity are, from a Marxist perspective, clearly extracting surplus value, and hence exploiting the labour of others. Toynbee presumably had in mind the idle rich or leisure class, as well perhaps as those whom the latter patronize, e.g. artists, though the latter, if paid for their work, are surely engaging in a form of economic activity, as indeed are the patrons who subsidize them. (Indeed, even the idle rich or leisure class engage in economic activity, if only as consumers.) However, this criterion, at least as described by Baker, is at least as capable of applying to the opposite end of the social spectrum – i.e. the welfare-dependent underclass. Did Toynbee really intend to suggest that the existence of the long-term unemployed is a distinctive marker of civilization? If so, is Baker really agreeing with him?

[51] The full list of criteria for civilization provided by Baker is as follows:

  1. In the ordinary circumstances of life in public places they cover the external genitalia and greater part of the trunk with clothes” (p507);
  2. They keep the body clean and take care to dispose of its waste elements” (p507);
  3. They do not practice severe mutilation or deformation of the body” (p507);
  4. They have knowledge of building in brick or stone, if the necessary materials are available in their territory” (p507);
  5. Many of them live in towns or cities, which are linked by roads” (p507);
  6. “They cultivate food plants” (p507);
  7. They domesticate animals and use some of the larger ones for transportif suitable species are available (p507);
  8. They have knowledge of the use of metals, if these are available” (p507);
  9. They use wheels” (p507);
  10. They exchange property by the use of money” (p507);
  11. They order their society by a system of laws, which are enforced in such a way that they ordinarily go about their various concerns in times of peace without danger of attack or arbitrary arrest” (p507);
  12. They permit accused people to defend themselves and call witnesses” (p507);
  13. They do not use torture to extract information or punishment” (p507);
  14. They do practice cannibalism” (p507);
  15. The religious systems include ethical elements and are not purely or grossly superstitious” (p507);
  16. They use a script… to communicate ideas” (p507);
  17. There is some facility in the abstract use of numbers, without consideration of actual objects” (p507);
  18. A calendar is in use” (p508);
  19. “[There are] arrangements for the instruction of the young in intellectual matters” (p508);
  20. There is some appreciation of the fine arts” (p508);
  21. Knowledge and understanding are valued as ends in themselves” (p508).

[52] Actually, some of the criteria include both technological and moral elements. For example, the second requirement, namely that the culture in question keep the body clean and take care to dispose of its waste elements”, at first seems a purely moral requirement. However, the disposal of sewage is, not only essential for the maintenance of healthy populations living at high levels of population density, but also often involves impressive feats of engineering (p507).
Similarly, the requirement that some people live in towns or cities” seems rather arbitrary. However, to sustain populations at the high population density required in towns and cities usually requires substantial technological, not to mention social and economic, development. Likewise, the building and maintenance of roads linking these settlements, also mentioned by Baker as part of the same criterion, is a technological achievement, often requiring, like the building of facilities for sewage disposal, substantial coordination of labour.

[53] Indeed, even the former Bishop of Edinburgh apparently agrees (see his book, Godless Morality: Keeping Religion out of Ethics). The classic thought-experiment used by moral philosophers to demonstrate that morality does not derive from God’s commandments is to ask devout believers whether, if, instead of commanding Thou shalt not kill, God had instead commanded Thou shalt kill, would they then consider killing a moral obligation? Most people, including devout believers, apparently concede otherwise. In fact, however, the hypothetical thought-experiment is not as hypothetical as many moral philosophers, and many Christians, seem to believe, as various passages in the Bible do indeed command mass killing and genocide (e.g. Deuteronomy 20: 16-17; Samuel 15:3; Deuteronomy 20: 13-14), and indeed rape too (Numbers 31:18).

[54] For example, in IQ and Racial Differences (1973), former president of the American Psychological Association and confirmed racialist Henry E Garrett claims:

Until the arrival of Europeans there was no literate civilization in the continent’s black belt. The Negro had no written language, no numerals, no calendar, no system of measurement. He never developed a plow or wheel. He never domesticated any animal. With the rarest exceptions, he built nothing more elaborate than mud huts and thatched stockades” (IQ and Racial Differences: p2).

[55] These explorers included David Livingston, the famous missionary, and Francis Galton, the infamous eugenicist, celebrated statistician and all-round Victorian polymath, in addition to Henry Francis FlynnPaul Du ChailluJohn Hanning Speke, Samuel Baker (the author John R Baker’s own grand-uncle) and George August Schweinfurth (p343).

[56] This, of course, depends on precisely how we define the words machine and ‘mechanical’. Thus, many authorities, especially military historians, class the simple bow as the first true ‘machine’. However, the only indigenous people known to lack even the bow and arrow at the time of their first contact with Europeans were the Australian Aboriginals of Australia and Tasmania.

[57] With regard to the ruins of Great Zimbabwe, Baker emphasizes that “the buildings in question are in no sense houses; the great majority of them are simply walls” (p402). Nor do they appear to have been part of a two-storey building (p402).
Unlike some other racialist authors who have attributed their construction to the possibly part-Jewish Lemba people, Baker attributes their construction and design to indigenous Africans (p405). However, he suggests their anomalous nature reflected that they had been constructed in (crude) imitation of buildings constructed outside of the “secluded area” of Africa by non-Negro peoples with whom the former were in a trading relationship (p407-8).
This would explain why the structures, though impressive by the standards of other constructions within the “secluded zone” of Africa from the same time-period, where buildings of brick or stone were rare and tended to be on a much smaller scale (so impressive, indeed, that, in the years since Baker’s book was published, they have even had an entire surrounding country named after them), are, by European or Middle Eastern standards of the same time period, quite shoddy. Baker also emphasizes:

The splendour and ostentation were made possible by what was poured into the country from foreign lands. One must acknowledge the administrative capacity of the rulers, but may question the utility of the ends to which much of it was put” (p409).

With regard to the technological achievements of black Africans mroe generally, Baker also acknowledges the adoption of iron smelting throughout most parts of Africa where the ore was available by the tenth century (p352; see also p373). However, while he attributes its origin to outside influence, recent research apparently suggests a much earlier, and indigenous, origin in some parts of sub-Saharan Africa. He also credits black Africans with great skill in forging iron into weapons and other tools (p353).

[58] Several plants seem to have been first domesticated in the Sahel region, and the Horn of Africa, both of which are part of sub-Saharan Africa. However, these areas lie outside of what Baker calls the “secluded area”, as I understand it. Also, populations from the Horn of Africa are, according to Baker predominantly Caucasoid (p225).

[59] The sole domestic animal that was perhaps first domesticated by black Africans is the guineafowl. Guineafowl are found wild throughout sub-Saharan Africa, but not elsewhere. It has therefore been argued, plausibly enough, that it was first domesticated in sub-Saharan Africa. However, Baker reports that the nineteenth-century explorers whose work he relies on “nowhere mention its being kept as a domestic animal by Negrids” (p375). Instead, he proposes it was probably first domesticated in Ethiopia, outside the “secluded area” as defined by Baker, and whose population are, according to Baker, predominantly Caucasoid (p225). However, he admits that there are no “early record of tame guinea-fowl in Ethiopia” (p375). 

[60] The relative absense of large wild mammals outside of sub-Saharan Afirca may partly be because such mammals have been driven to extinction or had their numbers depleted in recent times (e.g. wolves have been driven to extinction in Britain and Ireland, bison to the verge of extinction in North America). However, it is likely that Africa had a comparatively large number of large wild mammalian species even in ancient times.
This is because outside of Africa (notably in the Americas), many wild mammals were wiped out by the sudden arrival of humans with their formidable hunting skills to whom indigenous fauna were wholly unadapted. However, Africa is where humans first evolved. Therefore, prey species will have gradually evolved fear and avoidance of humans at the same time as humans themselves first evolved to become formidable hunters.
Thus, Africa, unlike other continents, never experienced a sudden influx of human hunters to whom its prey species were wholly unadapted. It therefore retains many of large wild game animals into modern times.

[61] Of course, rather conveniently for Diamonds theory, the wild ancestors of many modern domesticated animals, including horses and aurochs, are now extinct, so we have no way of directly assessing their temperament. However, we have every reason to believe that aurochs, at least, posed a far more formidable obstacle to domestication than does the zebra.

[62] Actually, a currently popular theory of the domestication of wolves/dogs holds that humans did not so much domesticate wolves/dogs as wolves/dogs domesticated themselves.

[63] Aurochs, and contemporary domestic cattle, also evince another trait that, according to Diamond, precludes their domestication – namely, it is not usually possible to keep two adult males of this species in the same field enclosure. Yet, according to Diamond, the social antelope species for which Africa is famous” could not be domesticated because:

The males of [African antelope] herds space themselves into territories and fight fiercely with one another when breeding. Hence, those antelope cannot be maintained in crowded enclosures in captivity” (Guns, Germs and Steel: p174).

Evidently, the ancient Eurasians who successfully domesticated the auroch never got around to reading Diamonds critially acclaimed bestseller. If they had, they could have learnt in advance to abandon the project as hopeless and hence save themselves the time and effort.

[64] With regard to the racial affinities of the ancient Egyptians, a source of some controversy in recent years, Baker concludes that, contrary to the since-popularized Afrocentrist Black Athena hypothesis, the ancient Egyptians were predominantly, but not wholly, Caucasoid, and that “the Negrid contribution to Egyptian stock was a small one” (p518). Indeed, there is presumably little doubt on this question, since, according to Baker, there is an abundance of well-preserved skulls from Egypt, not least due to the practice of mummifying corpses and thus:

More study has been devoted to the craniology of ancient Egypt than to that of any other country in the world” (p517).

From such data, Baker reports:

Morant showed that all the sets of ancient Egyptian skills that he analysed statistically were distinguishable by each of six criteria from Negrid skulls” (p518).

For what it’s worth, this conclusion is also corroborated by their self-depiction in artwork:

In their monuments the dynastic Egyptians represented themselves as having a long face, pointed chin with scanty beard, a straight or somewhat aquiline nose, black irises, and a reddish-brown complexion” (p518).

Similarly, in Race: the Reality of Human Differences (reviewed here, here and here), Sarich and Miele, claiming that Egyptian monuments are not mere ‘portraits but an attempt at classification’”, report that the Egyptians painted themselves as red, Asiatics or Semites as yellow, Southerns or Negroes” as black, and “Libyans, Westerners or Northerners” as “white, with blue eyes and fair beards” (Race: the Reality of Human Differences: p33).
Thus, if not actually black, neither were the ancient Egyptians exactly white either, as implausibly claimed by contemporary Nordicist Arthur Kemp, in his books, Children of Ra: Artistic, Historical, and Genetic Evidence for Ancient White Egypt and March of the Titans: The Complete History of the White Race.
In the latter work, Kemp contends that the ancient Egyptians were originally white, being part-Mediterranean (the Mediterranean race itself being now largely extinct, according to Kemp), but governed by a Nordic elite. Over time, however, he contends that they interbred with imported black African slaves and Semitic populations from the Middle East and hence the population was gradually transformed and hence Egyptian civilization degenerated.
This is, of course, a version of de Gobineau’s infamous theory that great empires inevitably decline because, through their imperial conquests, they subjugate, and hence ultimately interbreed with, the inferior peoples whom they have conquered (as well as with inward migrants attracted by higher living standards), which interbreeding supposedly dilutes the very racial qualities that permitted their original imperial glories.
Interestingly, consistent with Kemp’s theory, there is indeed some evidence that of an increase in the proportion of sub-Saharan African ancestry in Egypt since ancient times (Schuenemann et al 2017).
However, this same study demonstrating an increase in the proportion of sub-Saharan African ancestry in Egypt also showed that, contrary to Kemp’s theory, Egyptian populations always had close affinities to Middle Eastern populations (including Semites), and, in fact, owing to the increase in sub-Saharan African ancestry, and despite the Muslim conquest, actually had closer affinities to Near Eastern populations in ancient times than they do today (Schuenemann et al 2017).
Importantly, this study was based on DNA extracted from mummies, and, since mummification was a costly procedure that was almost always restricted to the wealthy, it therefore indicates that even the Egyptian elite were far from Nordic even in ancient times, as implausibly claimed by Kemp.
To his credit, Kemp does indeed amass some remarkable photographic evidence of Egyptian tomb paintings and monuments depicting figures, according to Kemp intended to represent Egyptians themselves, with blue eyes and light hair and complexions.
Admitting that Egyptian men were often depicted with reddish skin, he dismisses this as an artistic convention:

It was a common artistic style in many ancient Mediterranean cultures to portray men with red skins and women with white skins. This was done, presumably to reflect the fact that the men would have been outside working in the fields” (Children of Ra: p33). 

Actually, according to anthropologist Peter Frost, this artistic convention reflects real and innate differences, as well as differing sexually selected ideals of male and female beauty (see Dark Men, Fair Women).
Most interestingly, Kemp also includes photographs of some Egyptian mummies, including Ramses II, apparently with light-coloured hair. 
At first, I suspected this might reflect loss of pigmentation owing to the process of decay occurring after death, or perhaps to some chemical process involved in mummification.
Robert Brier, an expert on mummification, confirms that Ramses’s “strikingly blond” hair was indeed a consequence of its having been “dyed as a final step in the mummification process so that he would be young forever” (The Encyclopedia of Mummies: p153). However, he also reports in the next sentence that:

Microscopic inspection of the roots of the hair revealed that Ramses was originally a redhead” (The Encyclopedia of Mummies: p153).

Brier also confirms, again as claimed by Kemp, that one especially ancient predynastic mummy, displayed in the British Museum, was indeed nicknamed Ginger on account of its hair colour (The Encyclopedia of Mummies: p64). However, whether this was the natural hair colour of the person when alive is not clear.
At any rate, even if both Ginger and Ramses the Great were indeed natural redheads, in this respect they appear to have been very much the exception rather than the rule. Thus, Baker himself reports that
:

It would appear that their head-hair was curly, wavy, or almost straight, and very dark brown or black” (p518).

This conclusion is again based on the evidence of their mummies, and, since mummification was a costly procedure largely restricted to the wealthy, it again contradicts Kemp’s notion of a ‘Nordic elite’ ruling ancient Egypt. On this and other evidence, Baker therefore concludes:

There is general agreement… that the Europid element in the Egyptians from predynastic times onwards has been primarily Mediterranid, though it is allowed that Orientalid immigrants from Arabia made a contribution to the stock” (p518).

In short, ancient Egyptians, including Pharaohs and other elites, though certainly not black, were not really exactly white either, and certainly far from Nordic. Despite the increase in sub-Saharan African ancestry and the probable further influx of Middle Eastern DNA owing the Muslim conquest, they probably resembled modern Egyptians, especially the indigenous, Christian Copts.

[65] The same is true of the earlier runic alphabets of the Germanic peoples, the Paleohispanic scripts of the Iberian peninsula, and presumably also of the undeciphered Linear A alphabet that was in use at the outer edge of the European continent during the Bronze Age.

[66] Writing appears to have been developed first in Mesopotamia, then shortly afterwards in Egypt (though some Egyptologists claim priority on behalf of Egypt). However the relative geographic proximity of these two civilizations, their degree of contact with one anther and the coincidence in time, make it likely that one writing system was copied from the other. It then seems to have been independently developed in China. Writing was also developed, almost certainly entirely independently, in Mesoamerica. Other possible candidates for the independent development of writing include the Indus Valley civilization, and Easter Island, though, since neither script has been deciphered, it is not clear that they represent true writing systems, and the Easter Island script has also yet to be reliably dated.

[67] Actually, it is now suggested that both the Mayans and Indians may have been beaten to this innovation by the Babylonians, although, unlike the later Indians and Muslims, neither the Mayans nor the Babylonians went on to take full advantage of this innovation, by developing mathematics in a way made possible by their innovation. For this, it is Indian civilization that deserves credit. The invention of the concept by both the Maya and the Babylonians was, of course, entirely independent of one another, but the Indians, the Islamic civilization and other Eurasian civilizations probably inherited the concept ultimately from Babylonia.

[68] Interestingly, this excuse is not available in Africa. There, large mammals survived, probably because, since Africa was where anatomically modern humans first evolved, prey species evolved in concert with humans, and hence gradually evolved to fear and avoid humans, at the same time as humans themselves gradually evolved to be formidable predators. In contrast, the native species of the Americas would have been totally unprepared to protect themselves from human hunters, to whom they were completely ill-adapted, owing to the late, and, in evolutionary terms, sudden, peopling of the continent. This may be why, to this day, Africa has more large animals than any other continent.

[69] Baker also uses the complexity of a people’s language in order to assess their intelligence. Today, there seems to be an implicit assumption among many linguists that all languages are equal in their complexity. Thus, American linguists rightly emphasize the subtlety and complexity of, for example, African-American vernacular, which is certainly, by no means, merely a impoverished or corrupted version of standard English, but rather has grammatical rules all of its own, which often convey information that is lost on white Americans not conversant in this dialect.
However, there is no a priori reason to assume that all languages are equal in their capacity to express complex and abstract ideas. The size of vocabularies, for example, differs in different languages, as does the number of different tenses that are recognised. For example, the Walpiri language of some Australian Aboriginals is said to have only a few number terms, namely words for just onetwo’ and ‘many, while the Pirahã language of indigenous South Americans is said to get by with no number terms at all. Thus, when Baker contends that certain languages, notably the Arunta language of indigenous Australians, as studied by Alf Sommerfelt, and also the Akan language of Africa, are inherently impoverished in their capacity to express abstract thought, he may well be right.

________________________

References

Ali et al (2020) Genome-wide analyses disclose the distinctive HLA architecture and the pharmacogenetic landscape of the Somali population. Science Reports 10:5652.
Andrzejewski, Hall & Salib (2009) Anti-Semitism and Identification of Jewish Group Membership from Photographs Journal of Nonverbal Behavior 33(1):47-58.
Beals et al (1984) Brain size, cranial morphology, climate and time machines. Current Anthropology 25(3): 301–330
Bhatia et al (2014) Genome-wide Scan of 29,141 African Americans Finds No Evidence of Directional Selection since Admixture. American Journal of Human Genetics 95(4): 437–444.
Braveman et al (1995) Racial/ethnic differences in the likelihood of cesarean delivery, California. American Journal of Public Health 85(5): 625–630.
Carey (2001) Did the Irish come from Spain? History Ireland 9(3).
Chavez (2002) Reinventing the Wheel: The Economic Benefits of Wheeled Transportation in Early Colonial British West Africa. Africa’s Development in Historical Perspective. New York: Cambridge University Press.
Diamond (1994) Race without Color Discover Magazine, November 1st.
Edmonds et al (2013) Racial and ethnic differences in primary, unscheduled cesarean deliveries among low-risk primiparous women at an academic medical center: a retrospective cohort study. BMC Pregnancy Childbirth 13, 168.
Elhaik (2013). The missing link of Jewish European ancestry: contrasting the Rhineland and the Khazarian hypotheses. Genome Biology and Evolution 5 (1): 61–74.
Frost (2020) The costs of outbreeding what do we know? Evoandproud.blogspot.com, January 14th.
Getahun et al (2009) Racial and ethnic disparities in the trends in primary cesarean delivery based on indications. American Journal of Obstetrics and Gynecology 201(4):422.e1-7.
Helgason et al (2008) An association between the kinship and fertility of human couples. Science 319(5864):813-6.
Helgadottir et al (2006) A variant of the gene encoding leukotriene A4 hydrolase confers ethnicity-specific risk of myocardial infarction. Nature Genetics 38(1):68-74.
Hodgeson et al (2014) Early Back-to-Africa Migration into the Horn of Africa. PLoS Genetics 10(6): e1004393.
Jacobs & Fishberg (1906) ‘Nose, entry in The Jewish Encyclopedia.
Jacobs (1886) On the Racial Characteristics of Modern Jews, Journal of the Anthropological Institute, 1886, xv. 23-62.
Kay, K (2002). Morocco’s miracle muleBBC News 2 October.
Khan (2011a) The genetic affinities of Ethiopians. Discover Magazine, January 10.
Khan (2011b) A genomic sketch of the Horn of AfricaDiscover Magazine, June 10
Khan (2011c) Marry Far and Breed Tall Sons. Discover Magazine, July 7th.
Koziel et al (2011) Isolation by distance between spouses and its effect on children’s growth in height 146(1):14-9.
Labouriau & Amorim (2008) Comment on ‘An association between the kinship and fertility of human couples’ Science 12;322(5908):1634.
Lasker et al (2019) Global ancestry and cognitive abilityPsych 1(1), 431-459.
Law (2011) Wheeled transport in pre-colonial West Africa. Africa 50(3):249 – 262.
Lewis (2010) Why are mixed-race people perceived as more attractive? Perception 39(1):136-8.
Lewis et al (2015) Lumbar curvature: a previously undiscovered standard of attractiveness. Evolution and Human Behavior 36(5): 345-350.
Lewontin, (1972) The Apportionment of Human Diversity. In: Dobzhansky et al (eds) Evolutionary Biology. Springer, New York, NY.
Liebers et al (2004). “The herring gull complex is not a ring species”Proceedings of the Royal Society B: Biological Sciences 271 (1542): 893–901.
Loehlin et al (1973) Blood group genes and negro-white ability differences Behavior Genetics 3(3):263-270.
Martiniano et al (2016) Genomic signals of migration and continuity in Britain before the Anglo-Saxons. Nature Communications 7: 10326.
Nagoshi & Johnson (1986) The ubiquity of g. Personality and Individual Differences 7(2): 201-207.
Nebel et al (2001). The Y chromosome pool of Jews as part of the genetic landscape of the Middle East. American Journal of Human Genetics. 69 (5): 1095–112.
Nisbett (2005). Heredity, environment, and race differences in IQ: A commentary on Rushton and Jensen (2005). Psychology, Public Policy, and Law 11:302-310.
Norton et al (2009) Genetic evidence for the convergent evolution of light skin in Europeans and East Asians. Molecular Biology & Evolution 24(3): 710-722.
Nystrom et al (2008) Perinatal outcomes among Asian–white interracial couples. American Journal of Obstetrics and Gynecology 199(4), p382.e1-382.e6.
Olalde et al (2018) The Beaker phenomenon and the genomic transformation of northwest Europe. Nature 555: 190–196
Oppenheimer (2006) Myths of British Ancestry. Prospect Magazine, October 21.
Provine (1976) Geneticists and the biology of race crossing. Science 182(4114):790-796.
Relethford (2009) Race and global patterns of phenotypic variation. American Journal of Physical Anthropology 139(1):16-22.
Rong et al (1985) Fertile mule in China and her unusual foal. Journal of the Royal Society of Medicine. 78 (10): 821–25.
Rushton & Jensen (2005). Thirty years of research on race differences in cognitive ability. Psychology, Public Policy, and Law, 11:235-294.
Scarr et al (1977) Absence of a relationship between degree of white ancestry and intellectual skills within a black population Human Genetics 39(1):69-86.
Scarr & Weinberg (1976).IQ test performance of black children adopted by White families. American Psychologist 31:726-739.
Schiffels et al (2016) Iron Age and Anglo-Saxon genomes from East England reveal British migration history. Nature Communications 7: 10408.
Schuenemann et al (2017) Ancient Egyptian mummy genomes suggest an increase of Sub-Saharan African ancestry in post-Roman periods. Nature Communications 8:15694.
Singer (1958)  The Boskop ‘race’ problem. Man 58: 173-178.
Stanford University Medical Center (2008) Asian-white couples face distinct pregnancy risks, Stanford/Packard Eurekaltert.org, 1 October.
Tishkoff et al (2007) Convergent adaptation of human lactase persistence in Africa and Europe. Nature Genetics (1): 31-40.
Valdes (2020) Examining Cesarean Delivery Rates by Race: a Population-Based Analysis Using the Robson Ten-Group Classification System Journal of Racial and Ethnic Health Disparities.
Weinberg et al (1992). The Minnesota Transracial Adoption Study: A follow-up of IQ test performance at adolescence. Intelligence, 16, 117-135.
Whitney (1999) The Biological Reality of Race American Renaissance, October 1999.

Peter Singer’s ‘A Darwinian Left’

Peter Singer, A Darwinian Left: Politics, Evolution and Cooperation, London: Weidenfeld & Nicolson 1999.

Social Darwinism is dead. 

The idea that charity, welfare and medical treatment ought to be withheld from the poor, the destitute and the seriously ill so that they perish in accordance with the process of natural selection and hence facilitate further evolutionary progress survives only as a straw man sometimes attributed to conservatives by leftists in order to discredit them, and a form of guilt by association sometimes invoked by creationists in order to discredit the theory of evolution.[1]

However, despite the attachment of many American conservatives to creationism, there remains a perception that evolutionary psychology is somehow right-wing

Thus, if humans are fundamentally selfish, as Richard Dawkins is taken, not entirely accurately, to have argued, then this surely confirms the underlying assumptions of classical economics. 

Of course, as Dawkins also emphasizes, we have evolved through kin selection to be altruistic towards our close biological relatives. However, this arguably only reinforces conservatives’ faith in the family, and their concerns regarding the effects of family breakdown and substitute parents

Finally, research on sex differences surely suggests that at least some traditional gender roles – e.g. women’s role in caring for young children, and men’s role in fighting wars – do indeed have a biological basis, and also that patriarchy and the gender pay gap may be an inevitable result of innate psychological differences between the sexes

Political scientist Larry Arnhart thus champions what he calls a new ‘Darwinian Conservatism’, which harnesses the findings of evolutionary psychology in support of family values and the free market. 

Against this, however, moral philosopher and famed animal liberation activist Peter Singer, in ‘A Darwinian Left’, seeks to reclaim Darwin, and evolutionary psychology, for the Left. His attempt is not entirely successful. 

The Naturalistic Fallacy 

At least since David Hume, it has an article of faith among most philosophers that one cannot derive values from facts. To do otherwise is to commit what some philosophers refer to as the naturalistic fallacy

Edward O Wilson, in Sociobiology: The New Synthesis was widely accused of committing the naturalistic fallacy, by attempting to derive moral values form facts. However, those evolutionary psychologists who followed in his stead have generally taken a very different line. 

Indeed, recognition that the naturalistic fallacy is indeed a fallacy has proven very useful to evolutionary psychologists, since it has enabled them investigate the possible evolutionary functions of such morally questionable (or indeed downright morally reprehensible) behaviours as infidelityrape, warfare and child abuse while at the same time denying that they are somehow thereby providing a justification for the behaviours in question.[2] 

Singer, like most evolutionary psychologists, also reiterates the sacrosanct inviolability of the fact-value dichotomy

Thus, in attempting to construct his ‘Darwinian Left’, Singer does not attempt to use Darwinism in order to provide a justification or ultimate rationale for leftist egalitarianism. Rather, he simply takes it for granted that equality is a good thing and worth striving for, and indeed implicitly assumes that his readers will agree. 

His aim, then, is not to argue that socialism is demanded by a Darwinian worldview, but rather simply that it is compatible with such a worldview and not contradicted by it. 

Thus, he takes leftist ideals as his starting-point, and attempts to argue only that accepting the Darwinian worldview should not cause one to abandon these ideals as either undesirable or unachievable. 

But if we accept that the naturalistic fallacy is indeed a fallacy then this only raises the question: If it is indeed true that moral values cannot be derived from scientific facts, whence can moral values be derived?  

Can they only be derived from other moral values? If so, how are our ultimate moral values, from which all other moral values are derived, themselves derived? 

Singer does not address this. However, precisely by failing to address it, he seems to implicitly assume that our ultimate moral values must simply be taken on faith. 

However, Singer also emphasizes that rejecting the naturalistic fallacy does not mean that the facts of human nature are irrelevant to politics. 

On the contrary, while Darwinism may not prescribe any particular political goals as desirable, it may nevertheless help us determine how to achieve those political goals that we have already decided upon. Thus, Singer writes: 

An understanding of human nature in the light of evolutionary theory can help us to identify the means by which we may achieve some of our social and political goals… as well as assessing the possible costs and benefits of doing so” (p15). 

Thus, in a memorable metaphor, Singer observes: 

Wood carvers presented with a piece of timber and a request to make wooden bowls from it do not simply begin carving according to a design drawn up before they have seen the wood. Instead they will examine the material with which they are to work and modify their design in order to suit its grain…Those seeking to reshape human society must understand the tendencies inherent within human beings, and modify their abstract ideals in order to suit them” (p40). 

Abandoning Utopia? 

In addition to suggesting how our ultimate political objectives might best be achieved, an evolutionary perspective also suggests that some political goals might simply be unattainable, at least in the absence of a wholesale eugenic reengineering of human nature itself. 

In watering down the utopian aspirations of previous generations of leftists, Singer seems to implicitly concede as much. 

Contrary to the crudest misunderstanding of selfish gene theory, humans are not entirely selfish. However, we have evolved to put our own interests, and those of our kin, above those of other humans. 

For this reason, communism is unobtainable because: 

  1. People strive to promote themselves and their kin above others; 
  2. Only coercive state apparatus can prevent them so doing; 
  3. The individuals in control of this coercive apparatus themselves seek to promote the interests of themselves and their kin and corruptly use this coercive apparatus to do so. 

Thus, Singer laments: 

What egalitarian revolution has not been betrayed by its leaders?” (p39). 

Or, alternatively, as HL Mencken put it:

“[The] one undoubted effect [of political revolutions] is simply to throw out one gang of thieves and put in another.” 

In addition, human selfishness suggests, if complete egalitarianism were ever successfully achieved and enforced, it would likely be economically inefficient – because it would remove the incentive of self-advancement that lies behind the production of goods and services, not to mention of works of art and scientific advances. 

Thus, as Adam Smith famously observed: 

It is not from the benevolence of the butcher, the brewer, or the baker that we expect our dinner, but from their regard to their own interest.” 

And, again, the only other means of ensuring goods and services are produced besides economic self-interest is state coercion, which, given human nature, will always be exercised both corruptly and inefficiently. 

What’s Left? 

Singer’s pamphlet has been the subject of much controversy, with most of the criticism coming, not from conservatives, whom one might imagine to be Singer’s natural adversaries, but rather from other self-described leftists. 

These leftist critics have included both writers opposed to evolutionary psychology (e.g. David Stack in The First Darwinian Left), but also some other writers claiming to be broadly receptive to the new paradigm but who are clearly uncomfortable with some of its implications (e.g.  Marek Kohn in As We Know It: Coming to Terms with an Evolved Mind). 

In apparently rejecting the utopian transformation of society envisioned by Marx and other radical socialists, Singer has been accused by other leftists for conceding rather too much to the critics of leftism. In so doing, Singer has, they claim, in effect abandoned leftism in all but name and become, in their view, an apologist for and sell-out to capitalism. 

Whether Singer can indeed be said to have abandoned the Left depends, of course, on precisely how we define ‘the Left’, a rather more problematic matter than it is usually regarded as being.[3]

For his part, Singer certainly defines the Left in unusually broad terms.

For Singer, leftism need not necessarily entail taking the means of production into common ownership, nor even the redistribution of wealth. Rather, at its core, being a leftist is simply about being: 

On the side of the weak, not the powerful; of the oppressed, not the oppressor; of the ridden, not the rider” (p8). 

However, this definition is obviously problematic. After all, few conservatives would admit to being on the side of the oppressor. 

On the contrary, conservatives and libertarians usually reject the dichotomous subdivision of society into oppressed’ and ‘oppressor groups. They argue that the real world is more complex than this simplistic division of the world into black and white, good and evil, suggests. 

Moreover, they argue that mutually beneficial exchange and cooperation, rather than exploitation, is the essence of capitalism. 

They also usually claim that their policies benefit society as a whole, including both the poor and rich, rather than favouring one class over another.[4]

Indeed, conservatives claim that socialist reforms often actually inadvertently hurt precisely those whom they attempt to help. Thus, for example, welfare benefits are said to encourage welfare dependency, while introducing, or raising the level of, a minimum wage is said to lead to increases in unemployment. 

Singer declares that a Darwinian left would “promote structures that foster cooperation rather than competition” (p61).

Yet many conservatives would share Singer’s aspiration to create a more altruistic culture. 

Indeed, this aspiration seems more compatible with the libertarian notion of voluntary charitable donations replacing taxation than with the coercively-extracted taxes invariably favoured by the Left. After all, being forced to pay taxes is an example of coercion rather than true altruism. 

Nepotism and Equality of Opportunity 

Yet selfish gene theory suggests humans are not entirely self-interested. Rather, kin selection makes us care also about our biological relatives.

But this is no boon for egalitarians. 

Rather, the fact that our selfishness is tempered by a healthy dose of nepotism likely makes equality of opportunity as unattainable as equality of outcome – because individuals will inevitably seek to aid the social, educational and economic advancement of their kin, and those individuals better placed to do so will enjoy greater success in so doing. 

For example, parents with greater resources will be able to send their offspring to exclusive fee-paying schools or obtain private tuition for them; parents with better connections may be able to help their offspring obtain better jobs; while parents with greater intellectual ability may be better able to help their offspring with their homework. 

However, since many conservatives and libertarians are as committed to equality of opportunity as socialists are to equality of outcome, this conclusion may be as unwelcome on the right as on the left. 

Indeed, the theory of kin selection has even been invoked to suggest that ethnocentrism is innate and ethnic conflict is inevitable in multi-ethnic societies, a conclusion unwelcome across the mainstream political spectrum in the West today, where political parties of all persuasions are seemingly equally committed to building multi-ethnic societies. 

Unfortunately, Singer does not address any of these issues. 

Animal Liberation After Darwin 

Singer is most famous for his advocacy on behalf of what he calls animal liberation

In ‘A Darwinian Left’, he argues that the Darwinian worldview reinforces the case for animal liberation by confirming the evolutionary continuity between humans other animals. 

This suggests that there are unlikely to be fundamental differences in kind as between humans and other animals (e.g. in the capacity to feel pain) sufficient to justify the differences in treatment currently accorded humans and animals. 

It sharply contrasts account of creation in the Bible and the traditional Christian notion of humans as superior to other animals and as occupying an intermediate position between beasts and angels. 

Thus, Singer concludes: 

By knocking out the idea that we are a separate creation from the animals, Darwinian thinking provided the basis for a revolution in our attitudes to non-human animals” (p17). 

This makes our consumption of animals as food, our killing of them for sport, our enslavement of them as draft animals, or even pets, and our imprisonment of them in zoos and laboratories all ethically suspect, since these are not things that are generally permitted in respect of humans. 

Yet Singer fails to recognise that human-animal continuity cuts two ways. 

Thus, anti-vivisectionists argue that animal testing is not only immoral, but also ineffective, because drugs and other treatments often have very different effects on humans than they do on the animals used in drug testing. 

Our evolutionary continuity with non-human species makes this argument less plausible. 

Moreover, if humans are subject to the same principles of natural selection as other species, this suggests, not the elevation of animals to the status of humans, but rather the relegation of humans to just another species of animal

In short, we do not occupy a position midway between beasts and angels; we are beasts through and through, and any attempt to believe otherwise is mere delusion

This is, of course, the theme of John Gray’s powerful polemic Straw Dogs: Thoughts on Humans and Other Animals (which I have reviewed here). 

Finally, acceptance of the existence of human nature surely entails recognition of carnivory as a part of that nature. 

Of course, we must remember not to commit the naturalistic or appeal to nature fallacy.  

Thus, just because meat-eating may be natural for humans, in the sense that meat was a part of our ancestors diet in the EEA, this does not necessarily mean that it is morally right or even morally justifiable to eat meat. 

However, the fact that meat is indeed a natural part of the human diet does suggest that, in health terms, vegetarianism is likely to be nutritionally sub-optimal. 

Thus, the naturalistic fallacy or appeal to nature fallacy is not always entirely fallacious, at least when it comes to human health. What is natural for humans is indeed what we are biologically adapted to and what our body is therefore best designed to deal with.[5]

Therefore, vegetarianism is almost certainly to some degree sub-optimal in nutritional terms. 

Moreover, given that Singer is an opponent of the view that there is a valid moral distinction between acts and omissions, describing one of his core tenets in the Introduction to his book Writings on an Ethical Life as the belief that “we are responsible not only for what we do but also for what we could have prevented” (Writings on an Ethical Life: pxv), then we must ask ourselves: If he believes it is wrong for us to eat animals, does he also believe we should take positive measures to prevent lions from eating gazelles? 

Economics 

Bemoaning the emphasis of neoliberals on purely economic outcomes, Singer protests:

From an evolutionary perspective, we cannot identify wealth with self-interest… Properly understood self-interest is broader than economic self-interest” (p42). 

Singer is right. The ultimate currency of natural selection is not wealth, but rather reproductive success – and, in evolutionarily novel environments, wealth may not even correlate with reproductive success (Vining 1986). 

Thus, as discussed by Laura Betzig in Despotism and Differential Reproduction, a key difference between Marxism and sociobiology is the relative emphasis on production versus reproduction

Whereas Marxists see societal conflict and exploitation as reflecting competition over control of the means of production, for Darwinians, all societal conflict ultimately concerns control over, not the means of production, but rather what we might term the ‘means of reproduction’ – in other words, women, their wombs and vaginas

Thus, sociologist-turned-sociobiologist Pierre van den Berghe observed: 

“The ultimate measure of human success is not production but reproduction. Economic productivity and profit are means to reproductive ends, not ends in themselves” (The Ethnic Phenomenon: p165). 

Production is ultimately, in Darwinian terms, merely by which to gain the necessary resources to permit successful reproduction. The latter is the ultimate purpose of life

Thus, for all his ostensible radicalism, Karl Marx, in his emphasis on economics (‘production’) at the expense of sex (‘reproduction’), was just another Victorian sexual prude

Competition or Cooperation: A False Dichotomy? 

In Chapter  Four, entitled “Competition or Cooperation?”, Singer argues that modern western societies, and many modern economists and evolutionary theorists, put too great an emphasis on competition at the expense of cooperation

Singer accepts that both competition and cooperation are natural and innate facets of human nature, and that all societies involve a balance of both. However, he argues that different societies differ in their relative emphasis on competition or cooperation, and that it is therefore possible to create a society that places a greater emphasis on the latter at the expense of the former. 

Thus, Singer declares that a Darwinian left would: 

Promote structures that foster cooperation rather than competition” (p61) 

However, Singer is short on practical suggestions as to how a culture of altruism is to be fostered.[6]

Changing the values of a culture is not easy. This is especially so for a liberal democratic (as opposed to a despotic, totalitarian) government, let alone for a solitary Australian moral philosopher – and Singer’s condemnation of “the nightmares of Stalinist Russia” suggests that he would not countenance the sort of totalitarian interference with human freedom to which the Left has so often resorted in the past, and continues to resort to in the present (even in the West), with little ultimate success, in the past. 

But, more fundamentally, Singer is wrong to see competition and conflict as necessarily in conflict with altruism and cooperation

On the contrary, perhaps the most remarkable acts of cooperation, altruism and self-sacrifice are those often witnessed in wartime (e.g. kamikaze pilotssuicide bombers and soldiers who throw themselves on grenades). Yet war represents perhaps the most extreme form of competition and conflict known to man. 

In short, soldiers risk and sacrifice their lives, not only to save the lives of others, but also to take the lives of other others. 

Likewise, trade is a form of cooperation, but is as fundamental to capitalism as is competition. Indeed, I suspect most economists would argue that exchange is even more fundamental to capitalism than is competition

Thus, far from disparaging cooperation, neoliberal economists see voluntary exchange as central to prosperity. 

Ironically, then, popular science writer Matt Ridley also, like Singer, focuses on humans’ innate capacity for cooperation to justify political conclusions in his book, The Origins of Virtue

But, for Ridley, our capacity for cooperation provides a rationale, not for socialism, but rather for free markets – because humans, as natural traders, produce efficient systems of exchange which government intervention almost always only distorts. 

However, whereas economic trade is motivated by self-interested calculation, Singer seems to envisage a form of reciprocity mediated by emotions such as compassiongratitude and guilt
 
However, sociobiologist Robert Trivers argues in his paper that introduced the concept of reciprocal altruism to evolutionary biology that these emotions themselves evolved through the rational calculation of natural selection (Trivers 1971). 

Therefore, while open to manipulation, especially in evolutionarily novel environments, they are necessarily limited in scope. 

Group Differences 

Singer’s envisaged ‘Darwinian Left’ would, he declares, unlike the contemporary left, abandon: 

“[The assumption] that all inequalities are due to discrimination, prejudice, oppression or social conditioning. Some will be, but this cannot be assumed in every case” (p61). 

Instead, Singer admits that at least some disparities in achievement may reflect innate differences between individuals and groups in abilities, temperament and preferences. 

This is probably Singer’s most controversial suggestion, at least for modern leftists, since it contravenes the contemporary dogma of political correctness

Singer is, however, undoubtedly right.  

Moreover, his recognition that some differences in achievement as between groups reflect, not discrimination, oppression or even the lingering effect of past discrimination or oppression, but rather innate differences between groups in psychological traits, including intelligence, is by no means incompatible with socialism, or leftism, as socialism and leftism were originally conceived. 

Thus, it is worth pointing out that, while contemporary so-called cultural Marxists may decry the notion of innate differences in ability and temperament as between different racessexesindividuals and social classes as anathema, the same was not true of Marx himself

On the contrary, in famously advocating from each according to his ability, to each according to his need, Marx implicitly recognized that people differed in ability – differences which, given the equalization of social conditions envisaged under communism, he presumably conceived of as innate in origin.[7]

As Hans Eysenck observes:

“Stalin banned mental testing in 1935 on the grounds that it was ‘bourgeois’—at the same time as Hitler banned it as ‘Jewish’. But Stalin’s anti-genetic stance, and his support for the environmentalist charlatan Lysenko, did not derive from any Marxist or Leninist doctrine… One need only recall The Communist Manifesto: ‘From each according to his ability, to each according to his need’. This clearly expresses the belief that different people will have different abilities, even in the communist heaven where all cultural, educational and other inequalities have been eradicated” (Intelligence: The Battle for the Mind: p85).

Here Eysenck echoes the earlier observations of the brilliant, pioneering early twentieth century biologist, and unrepentant Marxist, JBS Haldane, who reputedly wrote in the pages of The Daily Worker in the 1940s, that:

The dogma of human equality is no part of Communism… The formula of Communism ‘from each according to his ability, to each according to his needs’ would be nonsense if abilities are equal.”

Thus, Steven Pinker, in The Blank Slate, points to the theoretical possibility of what he calls a “Hereditarian Left”, arguing for a Rawlsian redistribution of resources to the, if you like, innately ‘cognitively disadvantaged’.[8] 

With regard to group differences, Singer avoids discussing the incendiary topic of race differences in intelligence, a question too contentious for Singer to touch. 

Instead, he illustrates the possibility that not “all inequalities are due to discrimination, prejudice, oppression or social conditioning” with the marginally less incendiary case of sex differences.  

Here, it is sex differences, not in intelligence, but rather in temperament, preferences and personality that are probably more important, and likely explain occupational segregation and the so-called gender pay gap

Thus, Singer writes: 

If achieving high status increases access to women, then we can expect men to have a stronger drive for status than women” (p18). 

This alone, he implies, may explain both the universalilty of male rule and the so-called gender pay gap

However, Singer neglects to mention another biological factor that is also probably important in explaining the gender pay gap – namely, women’s attachment to infant offspring. This factor, also innate and biological in origin, also likely impedes career advancement among women. 

Thus, it bears emphasizing that never-married women with no children actually earn more, on average, than do unmarried men without children of the same age in both Britain and America.[9]

For a more detailed treatment of the biological factors underlying the gender pay gap, see Biology at Work: Rethinking Sexual Equality by professor of law, Kingsley Browne, which I have reviewed here.[10] See also my review of Warren Farrell’s Why Men Earn More, which can be found here, here and here.

Dysgenic Fertility Patterns? 

It is sometimes claimed by opponents of welfare benefts that the welfare system only encourages the unemployed to have more children so as to receive more benefits and thereby promotes dysgenic fertility patterns. In response, Singer retorts:

Even if there were a genetic component to something as nebulous as unemployment, to say that these genes are ‘deleterious’ would involve value judgements that go way beyond what the science alone can tell us” (p15).

Singer is, of course, right that an extra-scientific value judgement is required in order to label certain character traits, and the genes that contribute to them, as deleterious or undesirable. 

Indeed, if single mothers on welfare do indeed raise more surviving children than do those who are not reliant on state benefits, then this indicates that they have higher reproductive success, and hence, in the strict biological sense, greater fitness than their more financially independent, but less fecund, reproductive competitors. 

Therefore, far from being deleterious’ in the biological sense, genes contributing to such behaviour are actually under positive selection, at least under current environmental conditions.  

However, even if such genes are not ‘deleterious’ in the strict biological sense, this does not necessarily mean that they are desirable in the moral sense, or in the sense of contributing to successful civilizations and societal advancement. To suggest otherwise would, of course, involve a version of the very appeal to nature fallacy or naturalistic fallacy that Singer is elsewhere emphatic in rejecting. 

Thus, although regarding certain character traits, and the genes that contribute to them, as undesirable does indeed involve an extra-scientific “value judgement”, this is not to say that the “value judgement” in question is necessarily mistaken or unwarranted. On the contrary, it means only that such a value judgement is, by its nature, a matter of morality, not of science. 

Thus, although science may be silent on the issue, virtually everyone would agree that some traits (e.g. generosity, health, happiness, conscientiousness) are more desirable than others (e.g. selfishness, laziness, depression, illness). Likewise, it is self-evident that the long-term unemployed are a net burden on society, and that a successful society cannot be formed of people unable or unwilling to work. 

As we have seen, Singer also questions whether there can be “a genetic component to something as nebulous as unemployment”. 

However, in the strict biological sense, unemployment probably is indeed partly heritable. So, incidentally, are road traffic accidents and our political opinions – because each reflect personality traits that are themselves heritable (e.g. risk-takers and people with poor physical coordination and slow reactions probably have more traffic accidents; and perhaps more compassionate people are more likely to favour leftist politics). 

Thus, while it may be unhelpful and misleading to talk of unemployment as itself heritable, nevertheless traits of the sort that likely contribute to unemployment (e.g. intelligenceconscientiousnessmental and physical illness) are indeed heritable

Actually, however, the question of heritability, in the strict biological sense, is irrelevant. 

Thus, even if the reason that children from deprived backgrounds have worse life outcomes is entirely mediated by environmental factors (e.g. economic or cultural deprivation, or the bad parenting practices of low-SES parents), the case for restricting the reproductive rights of those people who are statistically prone to raise dysfunctional offspring remains intact. 

After all, children usually get both their genes and their parenting from the same set of parents – and this could be changed only by a massive, costly, and decidedly illiberal, policy of forcibly removing offspring from their parents.[11]

Therefore, so long as an association between parentage and social outcomes is established, the question of whether this association is biologically or environmentally mediated is simply beside the point, and the case for restricting the reproductive rights of certain groups remains intact.  

Of course, it is doubtful that welfare-dependent women do indeed financially benefit from giving birth to additional offspring. 

It is true that they may receive more money in state benefits if they have more dependent offspring to support and provide for. However, this may well be more than offset by the additional cost of supporting and providing for the dependent offspring in question, leaving the mother with less to spend on herself. 

However, even if the additional monies paid to mothers with dependent children are not sufficient as to provide a positive financial incentive to bearing additional children, they at least reduce the financial disincentives otherwise associated with rearing additional offspring.  

Therefore, given that, from an evolutionary perspective, women probably have an innate desire to bear additional offspring, it follows that a rational fitness-maximizer would respond to the changed incentives represented by the welfare system by increasing their reproductive rate.[12]

Towards A New Socialist Eugenics?

If we accept Singer’s contention that an understanding of human nature can help show us how achieve, but not choose, our ultimate political objectives, then eugenics could be used to help us achieve the goal of producing the better people and hence, ultimately, better societies. 

Indeed, given that Singer seemingly concedes that human nature is presently incompatible with communist utopia, perhaps then the only way to revive the socialist dream of communism is to eugenically re-engineer human nature itself. 

Thus, it is perhaps no accident that, before World War Two, eugenics was a cause typically associated, not with conservatives, nor even, as today, with fascism and German National Socialism, but rather with the political left, the main opponents of eugenics, on the other hand, being Christian conservatives.

Thus, early twentieth century socialist-eugenicists like H.G. Wells, Sidney Webb, Margaret Sanger and George Bernard Shaw may then have tentatively grasped what eludes contemporary leftists, Singer very much included – namely that re-engineering society necessarily requires as a prerequisite re-engineering Man himself.[13]

_________________________

Endnotes

[1] Indeed, the view that the poor and ill ought to be left to perish so as to further the evolutionary process seems to have been a marginal one even in its ostensible late nineteenth century heyday (see Bannister, Social Darwinism Science and Myth in Anglo-American Social Thought). The idea always seems, therefore, to have been largely, if not wholly, a straw man.

[2] In this, the evolutionary psychologists are surely right. Thus, no one accuses biomedical researchers of somehow ‘justifying disease’ when they investigate how infectious diseases, in an effort maximize their own reproductive success, spread form host to host. Likewise, nobody suggests that dying of a treatable illness is desirable, even though this may have been the ‘natural’ outcome before such ‘unnatural’ interventions as vaccination and antibiotics were introduced.

[3] The convenional notion that we can usefully conceptualize the political spectrum on a single dimensional left-right axis is obviously preposterous. For one thing, there is, at the very least, a quite separate liberal-authoritarian dimension. However, even restricting our definition of the left-right axis to purely economic matters, it remains multi-factorial. For example, Hayek, in The Road to Serfdom classifies fascism as a left-wing ideology, because it involved big government and a planned economy. However, most leftists would reject this definition, since the planned economy in question was designed, not to reduce economic inequalities, but rather, in the case of Nazi Germany at least, to fund and sustain an expanded military force, a war economy, external military conquest and grandiose vanity public works architectural projects. The term right-wing’ is even more problematic, including everyone from fascists, to libertarians to religious fundamentalists. Yet a Christian fundamentalist who wants to outlaw pornography and abortion has little in common with either a libertarian who wants to decriminalize prostitution and child pornography, nor with a eugenicist who wants to make abortions, for certain classes of person, compulsory. Yet all three are classed together as ’right-wing’ even though they share no more in common with one another than any does with a raving unreconstructed Marxist.

[4] Thus, the British Conservatives Party traditionally styled themselves one-nation conservatives, who looked to the interests of the nation as a whole, rather than what they criticized as the divisive ‘sectionalism’ of the trade union and labour movements, which favoured certain economic classes, and workers in certain industries, over others, just as contemporary leftists privilege the interests of certain ethnic, religious and culturally-defined groups (e.g. blacks, Muslims, feminists) over others (i.e. white males).

[5] Of course, some ‘unnatural’ interventions have positive health benefits. Obvious examples are modern medical treatments such as penicillin, chemotherapy and vaccination. However, these are the exceptions. They have been carefully selected and developed by scientists to have this positive effect, have gone through rigorous testing to ensure that their effects are indeed beneficial, and are generally beneficial only to people with certain diagnosed conditions. In contrast, recreational drug use almost invariably has a negative effect on health.
It might also be noted that, although their use by humans may be ‘unnatural’, the role of antibiotics in fighting bacterial infection is not itself ‘unnatural’, since antibiotics such as penicillin themselves evolved as a natural means by which one microorganism, namely mould, a form of fungi, fights another form of microorganism, namely bacteria.

[6] It is certainly possible for more altruistic cultures to exist. For example, the famous (and hugely wasteful) potlatch feasts of some Native American cultures, which involved great acts of both altruism and wanton waste, exemplify an extreme form of competitive altruism, analogous to conspicuous consumption, and may be explicable as a form of status display in accordance with Zahavi’s handicap principle. However, recognizing that such cultures exist does not easily translate into working out how to create or foster such cultures, let alone transform existing cultures in this direction.

[7]  Indeed, by modern politically-correct standards, Marx was a rampant racist, not to mention an anti-Semite

[8] The term Rawlsian is a reference to political theorist John Rawles version of social contract theory, whereby he poses the hypothetical question as to what arrangement of political, social and economic affairs humans would favour if placed in what he called the original position, where they would be unaware of, not only their own race, sex and position in to the socio-economic hierarchy, but also, most important for our purposes, their own level of innate ability. This Rawles referred to as ’veil of ignorance’. 

[9] As Warren Farrell documents in his excellent Why Men Earn More (which I have reviewed here, here and here), in the USA, women who have never married and have no children actually earn more than men who have never married and have no children and have done since at least the 1950s (Why Men Earn More: pxxi). More precisely, according to Farrell, never-married men without children on average earn only about 85% of their childless never-married female counterparts (Ibid: pxxiii).
The situation is similar in the UK. Thus, economist JR Shackleton reports:

Women in the middle age groups who remain single earn more than middle-aged single males” (Should We Mind the Gap? p30).

The reasons unmarried, childless women earn more than unmarried childless men are multifarious and include:

  1. Married women can afford to work less because they appropriate a portion of their husband’s income in addition to their own
  2. Married men and men with children are thus obliged to earn even more so as to financially support, not only themselves, but also their wife, plus any offspring;
  3. Women prefer to marry richer men and hence poorer men are more likely to remain single;
  4. Childcare duties undertaken by women interfere with their earning capacity.

[10]  Incidentally, Browne has also published a more succinct summary of the biological factors underlying the pay-gap that was first published in the same ‘Darwinism Today’ series as Singer’s ‘A Darwinian Left’, namely Divided Labors: An Evolutionary View of Women at Work. However, much though I admire Browne’s work, this represents a rather superficial popularization of his research on the topic, and I would recommend instead Browne’s longer Biology at Work: Rethinking Sexual Equality (which I have reviewed here) for a more comprehenseive treatment of the same, and related, topics. 

[11] A precedent for just such a programme, enacted in the name of socialism, albeit imposed consensually, was the communal rearing practices in Israeli Kibbutzim, since largely abandoned. Another suggestion along rather different lines comes from a rather different source, namely Adolf Hitler, who, believing that nature trumped nurture, is quoted in Mein Kampf as proposing: 

The State must also teach that it is the manifestation of a really noble nature and that it is a humanitarian act worthy of all admiration if an innocent sufferer from hereditary disease refrains from having a child of his own but bestows his love and affection on some unknown child whose state of health is a guarantee that it will become a robust member of a powerful community” (quoted in: Parfrey 1987: p162). 

[12] Actually, it is not entirely clear that women do have a natural desire to bear offspring. Other species probably do not have any such natural desire. After all, since they are almost certainly are not aware of the connection between sex and child birth, such a desire would serve no adaptive purpose and hence would never evolve. All an organism requires is a desire for sex, combined perhaps with a tendency to care for offspring after they are born. (Indeed, in principle, a female does not even require a desire for sex, only a willingness to submit to the desire of a male for sex.) As Tooby and Cosmides emphasize: 

Individual organisms are best thought of as adaptation-executers rather than as fitness-maximizers.” 

There is no requirement for a desire for offspring as such. Nevertheless, anecdotal evidence of so-called broodiness, and the fact that most women do indeed desire children, despite the costs associated with raising children, suggests that, in human females, there is indeed some innate desire for offspring. Curiously, however, the topic of broodiness is not one that has attracted much attention among evolutionists.

[13] However, there is a problem with any such case for a ‘Brave New Socialist Eugenics’. Before the eugenic programme is complete, the individuals controlling eugenic programmes (be they governments or corporations) would still possess a more traditional human nature, and may therefore have less than altruistic motivations themselves. This seems to suggest then that, as philosopher John Gray concludes in Straw Dogs: Thoughts on Humans and Other Animals (which I have reviewed here):  

“[If] human nature [is] scientifically remodelled… it will be done haphazardly, as an upshot of the struggles in the murky world where big business, organized crime and the hidden parts of government vie for control” (Straw Dogs: p6).

References  

Parfrey (1987) Eugenics: The Orphaned Science. In Parfrey (Ed.) Apocalypse Culture (New York: Amoc Press). 

Trivers 1971 The evolution of reciprocal altruism Quarterly Review of Biology 46(1):35-57 

Vining 1986 Social versus reproductive success: The central theoretical problem of human sociobiologyBehavioral and Brain Sciences 9(1), 167-187.

Donald Symons’ ‘The Evolution of Human Sexuality’: A Founding Work of Modern Evolutionary Psychology

The Evolution of Human Sexuality by Donald Symons (Oxford University Press 1980). 

Research over the last four decades in the field that has come to be known as evolutionary psychology has focused disproportionately on mating behaviour. Geoffrey Miller (1998) has even argued that it is the theory of sexual selection rather than that of natural selection which, in practice, guides most research in this field. 

This does not reflect merely the prurience of researchers. Rather, given that reproductive success is the ultimate currency of natural selection, mating behaviour is, perhaps along with parental investment, the form of behaviour most directly subject to selective pressures.

Almost all of this research traces its ancestry ultimately to Donald Symons’ ‘The Evolution of Human Sexuality’ by Donald Symons. Indeed, much of it was explicitly designed to test claims and predictions formulated by Symons himself in this very book.

Age Preferences

For example, in his discussion of the age at which women are perceived as most attractive by males, Symons formulated two alternative hypotheses. 

First, if human evolutionary history were characterized by fleeting one-off sexual encounters (i.e. one-night standscasual sex and hook-ups), then, he reasoned, men would have evolved to find women most attractive when the latter are at the age of their maximum fertility

For women, fertility is said to peak around when a woman reaches her mid-twenties since, although women still in their teens have high pregnancy rates, they also experience greater risk of birth complications

However, if human evolutionary history were characterized instead by long-term pair bonds, then men would have evolved to be maximally attracted to somewhat younger women (i.e. those at the beginning of their reproductive careers), so that, by entering a long-term relationship with the woman at this time, a male is potentially able to monopolize her entire lifetime reproductive output (p189). 

More specifically, males would have evolved to prefer females, not of maximal fertility, but rather of maximal reproductive value, a term borrowed from demography and population genetics which refers to a person’s expected future reproductive output given their current age. Unlike fertility, a woman’s reproductive value peaks around her mid- to late-teens.  

On the basis of largely anecdotal evidence, Symons concludes that human males have evolved to be most attracted to females of maximal reproductive value rather than maximal fertility.  

Subsequent research designed to test between Symons’s rival hypotheses has largely confirmed his speculative hunch that it is younger females in their mid- to late-teens who are perceived by males as most attractive (e.g. Kenrick and Keefe 1992). 

Why Average is Attractive

Symons is also credited as the first person to recognize that a major criterion of attractiveness is, paradoxically, averageness, or at least the first to recognize the significance of, and possible evolutionary explanation for, this discovery.[1] Thus, Symons argues that: 

“[Although] health and status are unusual in that there is no such thing as being too healthy or too high ranking… with respect to most anatomical traits, natural selection produces the population mean” (p194). 

On this view, deviations from the population mean are interpreted as the result of deleterious mutations or developmental instability, and hence bad genes.[2]

Concealed Ovulation

Support has even emerged for some of Symons’ more speculative hunches.

For example, one of Symons’ two proposed scenarios for the evolution of concealed ovulation, in which he professed “little confidence” (p141), was that this had evolved so as to impede male mate-guarding and enable females select a biological father for their offspring different from their husbands (p139-141).

Consistent with this theory, studies have found that women’s mate preferences vary throughout their menstrual cycle in a manner compatible with a so-called ‘dual mating strategy’, preferring males evidencing a willingness to invest in offspring at most times, but, when at their most fertile, preferring characteristics indicative of genetic quality (e.g. Penton-Voak et al 1999). 

Meanwhile, a questionnaire distributed via a women’s magazine found that women engaged in extra-marital affairs do indeed report engaging in ‘extra-pair copulations’ (EPCs) at times likely to coincide with ovulation (Bellis and Baker 1990).[3]

The Myth of Female Choice

Interestingly, Symons even anticipated some of the mistakes evolutionary psychologists would be led into.

Thus, he warns that researchers in modern western societies may be prone to overestimate the importance of female choice as a factor in human evolution, because, in their own societies, this is a major factor, if not the major factor, in determining marriage and sexual and romantic relationships (p203).[4]

However, in ancestral environments (i.e. what evolutionary psychologists now call the Environment of Evolutionary Adaptedness or EEA) arranged marriages were likely the norm, as they are in most premodern cultures around the world today (p168).[5]

Thus, Symons concludes: 

There is no evidence that any features of human anatomy were produced by intersexual selection [i.e. female choice]. Human physical sex differences are explained most parsimoniously as the outcome of intrasexual selection (the result of male-male competition)” (p203). 

Thus, human males have no obvious analogue of the peacock’s tail, but they do have substantially greater levels of upper-body strength and violent aggression as compared to females.[6]

This was a warning almost entirely ignored by subsequent generations of researchers before being forcefully reiterated by Puts (2010)

Homosexuality as a ‘Test-Case

An idea of the importance of Symons’s work can be ascertained by comparing it with contemporaneous works addressing the same subject-matter.

Edward O Wilson’s On Human Nature was first published in 1978, only a year before Symons’s ‘The Evolution of Human Sexuality’. 

However, whereas Symons’s book set out much of the theoretical basis for what would become the modern science of evolutionary psychology, Wilson’s chapter on “Sex” has dated rather less well, and a large portion of chapter is devoted to introducing a now faintly embarrassing theory of the evolution of homosexuality which has subsequently received no empirical support (see Bobrow & Bailey 2001).[7]

In contrast, Symons’s own treatment of homosexuality is innovative. It is also characteristic of his whole approach and illustrates why ‘The Evolution of Human Sexuality‘ has been described by David Buss as “the first major treatise on evolutionary psychology proper” (Handbook of Evolutionary Psychology: p251).

Rather than viewing all behaviours as necessarily adaptive (as critics of evolutionary psychology, such as Stephen Jay Gould, have often accused sociobiologists of doing),[8] Symons instead focuses on admittedly non-adaptive (or, indeed, even maladaptive) behaviours, not because he believes them to be adaptive, but rather because they provide a unique window on the nature of human sexuality.

Accordingly, Symons does not concern himself with how homosexuality evolved, implicitly viewing it as a rare and maladaptive malfunctioning of normal sexuality. Yet the behaviour of homosexuals is of interest to Symons because it provides a window on the nature of male and female sexuality as it manifests itself when freed from the constraints imposed by the conflicting desires of the opposite sex.

On this view, the rampant promiscuity manifested by many homosexual men (e.g. cruising and cottaging in bathhouses and public lavatories, or Grindr hookups) reflects the universal male desire for sexual variety when freed from the constraints imposed by the conflicting desires of women. 

This desire for sexual variety is, of course, obviously reproductively unproductive among homosexual men themselves. However, it evolved because it enhanced the reproductive success of heterosexual men by motivating them to attempt to mate with multiple females and thereby father multiple offspring.

Thus, a powerful ruler like with a large harem like Ismail the Bloodthirsty’ of Morocco could reputedly father as many as 888 offspring.

In contrast, burdened with pregnancy and lactation, women’s potential reproductive rate is more tightly constrained than that of men. They therefore have little to gain reproductively by mating with multiple males, since they can usually gestate, and nurse, only one offspring at a time.

It is therefore notable that, among lesbians, there is little evidence of the sort of rampant promiscuity common among gay men. Instead, lesbian relationships seem to be characterized by much the same features as heterosexual coupling (i.e. long-term pair-bonds).

The similarity of heterosexual coupling to that of lesbians, and the striking contrast with that of male homosexuals, suggests that it is women, not men, who exert decisive influence in dictating the terms of heterosexual coupling.[9]

Thus, Symons reports:

There is enormous cross-cultural variation in sexual customs and laws and the extent of male control, yet nowhere in the world do heterosexual relations begin to approximate those typical of homosexual men This suggests that, in addition to custom and law, heterosexual relations are structured to a substantial degree by the nature and interests of the human female” (p300). 

This conclusion is, of course, diametrically opposite to the feminist contention that it is men who dictate the terms of heterosexual coupling and for whose exclusive benefit such relationships are structured.

It also suggests, again contrary to feminist assumptions of male dominance, that most men are ultimately frustrated in achieving their sexual ambitions to a far greater extent than are most women. 

Thus, Symons concludes: 

The desire for sexual variety dooms most human males to a lifetime of unfulfilled longing” (p228). 

Here, Symons anticipates Camille Paglia who was later to famously observe: 

Men know they are sexual exiles. They wander the earth seeking satisfaction, craving and despising, never content. There is nothing in that anguished motion for women to envy” (Sexual Personae: p19). 

Criticisms of Symons’s Use of Homosexuality as a Test-Case

There is, however, a potential problem with Symons’s use of homosexual behaviour as a window onto the nature of male and female sexuality as they manifest themselves when freed from the conflicting desires of the opposite sex. The whole analysis rests on a questionable premise – namely that homosexuals are, their preference for same-sex partners aside, otherwise similar, if not identical, to heterosexuals of their own sex in their psychology and sexuality.

Symons defends this assumption, arguing: 

There is no reason to suppose that homosexuals differ systematically from heterosexuals in any way other than their sexual object choice” (p292). 

Indeed, in some respects, Symons seems to see even “sexual object choice” as analogous among homosexuals and heterosexuals of the same sex.

For example, he observes that, unlike women, both homosexual and heterosexual men tend to evaluate prospective mates primarily on the basis their physical appearance and youthfulness (p295). 

Thus, in contrast to the failure of periodicals featuring male nudes to attract a substantial female audience (see below), Symons notes the existence of a market for gay pornography parallel in most respects to heterosexual porn – i.e. featuring young, physically attractive models in various states of undress (p301).

This, of course, contradicts the feminist notion that men are led to ‘objectify’ women only due to the sexualized portrayal of the latter in the media.

Instead, Symons concludes: 

That homosexual men are at least as likely as heterosexual men to be interested in pornography, cosmetic qualities and youth seems to me to imply that these interests are no more the result of advertising than adultery and alcohol consumption are the result of country and western music” (p304).[10] 

However, this assumption of the fundamental similarity of heterosexual and homosexual male psychology has been challenged by David Buller in his book, Adapting Minds: Evolutionary Psychology and the Persistent Quest for Human Nature.

Buller cites evidence that male homosexuals are ‘feminized’ in many aspects of their behaviour.

For example, one interesting recent study found that male homosexuals have more female-typical occupation interests than do heterosexual males (Ellis & Ratnasingam 2012).

Moreover, one of the few consistent early correlates of homosexuality is gender non-conformity in childhood and some evidence (e.g. digit ratios, the fraternal birth order effect) has been interpreted to suggest that the level of prenatal exposure to masculinizing androgens (e.g. testosterone) in utero affects sexual orientation (see Born Gay: The Pyschobiology of Sexual Orientation).

Indeed, Symons himself mentions the evidence of an association between homosexuality and levels of masculinizing androgens in utero (albeit in respect of lesbians rather than of male homosexuality) just a few pages before his discussion of the promiscuous behaviours of male homosexuals (p289).

As Buller also notes, although gay men seem, like heterosexual men, to prefer youthful sexual partners, they also appear to prefer sexual partners who are, in other respects highly masculine.[11]

Thus, Buller observes: 

“The males featured in gay men’s magazines embody very masculine, muscular physiques, not pseudo-feminine physiques” (Adapting Minds: p227).

Indeed, the models in such magazines seem in most respects similar in physical appearance to the male models, pop stars, actors and other ‘sex symbols’ and celebrities fantasized about by heterosexual women and girls.

How then are we to resolve this apparent paradox?

One possible explanation that some aspects of the psychology of male homosexuals are feminized but not others – perhaps because different parts of the brain are formed at different stages of prenatal development, at which stages the levels of masculinizing androgens in the womb may vary. 

Indeed, there is even some evidence that homosexual males may be hyper-masculinized in some aspects of their physiology.

For example, it has been found that homosexual males report larger penis-sizes than heterosexual men (Bogaert & Hershberger 1999). 
 
This, researchers Glenn Wilson and Qazi Rahman propose, may be because: 

If it is supposed that the barriers against androgens with respect to certain brain structures (notably those concerned with homosexuality) lead to increased secretion in an effort to break through, or some sort of accumulation elsewhere… then there may be excess testosterone left in other departments” (Born Gay: The Psychobiology of Sex Orientation: p80). 

Another possibility is that male homosexuals actually lie midway between heterosexual men and women in their degree of masculinization.  

On this view, homosexual men come across as relatively feminine only because we naturally tend to compare them to other men (i.e. heterosexual men). However, as compared to women, they may be relatively masculine, as reflected in the male-typical aspects of their sexuality focused upon by Symons.

Interestingly, this latter interpretation suggests the slightly disturbing possibility that, freed from the restraints imposed by women, heterosexual men would be even more indiscriminately promiscuous than their homosexual counterparts.

Evidence consistent with this interpretation is provided by one study from the 1980s which found that, when approached by a female stranger (also a student), on a University campus, with a request to go to bed with them, fully 72% of male students agreed (Clark and Hatfield 1989). 

In contrast, in the same study, not a single one of the 96 females approached by male strangers with the same request on the same university campus agreed to go to bed with the male stranger.

Yet what percentage of the female students subsequently sued the university for sexual harassment was not reported.

Pornography as a “Natural Experiment

For Symons, fantasy represents another window onto sexual and romantic desires. Like homosexuality, fantasy is, by its very nature, unconstrained by the conflicting desires of the opposite sex (or indeed by anything other than the imagination of the fantasist). 

Symons later collaborated in an investigation into sexual fantasy by means of a questionnaire (Ellis and Symons 1990). 

However, in the present work, he investigates fantasy indirectly by focusing on what he calls “the natural experiment of commercial periodical publishing” – i.e. pornographic magazines (p182).

In many respects, this approach is preferable to a survey because, even in an anonymous questionnaire, individuals may be less than honest when dealing with a sensitive topic such as their sexual fantasies. On the other hand, they are unlikely to regularly spend money on a magazine unless they are genuinely attracted by its contents.

Before the internet age, softcore pornographic magazines, largely featuring female nudes, commanded sizeable circulations, despite the not insubstantial stigma attached to their purchase. However, their readership (if indeed ‘readership’ is the right words, since there was typically little reading involved, save of the ‘one-handed’ variety) was almost exclusively male.

In contrast, there was little or no female audience for magazines containing pictures of naked males. Instead, magazines marketed towards women (e.g. fashion magazines) contain, mostly, pictures of other women.

Indeed, when, in the 1970s, attempts were made, in the misguided name of feminism and ‘women’s liberation, to market magazines featuring male nudes to a female readership, one such title, Viva, abandoned publishing male nudes after just a few years due to lack of interest or demand, then subsequently went bust just a few years after that, while the other, Playgirl, although it remained in publication for many years and did not entirely abandon male nudes, was notorious, as a consequence, for attracting a readership composed in large part of homosexual men.

Symons thus concludes forcefully and persuasively: 

The notion must be abandoned that women are simply repressed men waiting to be liberated” (p183). 

Indeed, though it has been loudly and enthusiastically co-opted by feminists, this view of women, and of female sexuality – namely women as “repressed men waiting to be liberated” – represents an obviously quintessentially male persepective. 

Indeed, taken to extremes, it has even been used as a justification for rape.

Thus, the curious, though recurrent, sub-Freudian notion that female rape victims actually secretly enjoy being raped seems to rest ultimately on the assumption that female sexuality is fundamentally the same as that of men (i.e. indiscriminately enjoying of promiscuous sex) and that it is only women’s alleged sexual ‘repression’ that prevents them admitting as much.[12]

Romance Literature 

Unfortunately, however, there is notable omission in Symons’s discussion of pornography as a window into male sexuality – namely, he omits to consider whether there exists any parallel artistic genre that offers equivalent insight into the female psyche.

Later writers on the topic have argued that romance novels (e.g. Mills and Boon, Jane Austin), whose audience is as overwhelmingly female as pornography’s is male, represent the female equivalent of pornography, and that analysis of the the content of such works provides insights into female mate preferences parallel to those provided into male psychology by pornography (e.g. Kruger et al 2003; Salmon 2004; see also Warrior Lovers: Erotic Fiction, Evolution and Female Sexuality, co-authored by Symons himself).

Thus, popular science writer Matt Ridley reports:

Two industries relentlessly exploit the sexual fantasizing of men and women: pornography and the publishing of romance novels: Pornography is aimed almost entirely at men. It varies little from a standard formula all over the world… The romance novel, by contrast, is aimed entirely at a female market. It, too, depicts a fictional world that has changed remarkably little except in adapting to female career ambitions and to a less inhibited attitude toward the description of sex” (The Red Queen: p270-271)

Symons touches upon this analogy only in passing, when he observes that:

Heterosexual men are, of course, aware that the female sexuality portrayed in men’s magazines reflects male fantasy more than female reality, just as heterosexual women are aware that the happy endings of stories in romance magazines exist largely in the realm of fantasy” (p293).

Yet, while feminists perpetually complain about how pornography supposedly creates unrealistic expectations of women and girls and puts undue pressure on women and girls to live up to this male fantasy, few men complain about how the equally unrealistic portrayal of men in romance literature creates unrealistic expectations of boys and men and puts undue pressure on boys and men to live up to a female fantasy.

Female Orgasm as Non-Adaptive

An entire chapter of ‘The Evolution of Human Sexuality’, namely Chapter Three (entitled, “The Female Orgasm: Adaptation or Artefact”), is devoted to rejecting the claim that the female orgasm represents a biological adaptation.

This is perhaps excessive. However, it does at least conveniently contradicts the claim of some critics of evolutionary psychology, and of sociobiology, such as Stephen Jay Gould that the field is ‘ultra-Darwinian’ or ‘hyper-adaptionist’ and committed to the misguided notion that all traits are necessarily adaptive.[13]

In contrast, Symons champions the thesis that the female capacity for orgasm is a simply non-adaptive by-product of the male capacity to orgasm, the latter of which is of course adaptive.

On this view, the female orgasm (and clitoris) is, in effect, the female equivalent of male nipples (only more fun).

Certainly, Symons convincingly critiques the romantic notion, popularized by Desmond Morris among others, that the female orgasm functions as a mechanism designed to enhance ‘pair-bonding between couples.

However, subsequent generations of evolutionary psychologists have developed less naïve models of the adaptive function of female orgasm.

For example, Geoffrey Miller argues that the female orgasm, and clitoris, functions as an adaptation for mate choice (The Mating Mind: p239-241).

Of course, at first glance, experiencing orgasm during coitus may appear to be a bit late for mate choice, since, by the time coitus has occurred, the choice in question has already been made. However, given that, among humans, most sexual intercourse is non-reproductive (i.e. does not result in conception), the theory is not altogether implausible.

On this view, the very factors which Symons views as suggesting female orgasm is non-adaptive – such as the relative difficultly of stimulating female orgasm during ordinary vaginal sex – are positive evidence for its adaptive function in carefully discriminating between suitors/lovers to determine their desirability as father for a woman ’s offspring.

Nevertheless, at least according to the stringent criteria set out by George C Williams in his classic Adaptation and Natural Selection, as well as the more general principle of parsimony (also known as Occam’s Razor), the case for female orgasm as an adaptation remains unproven (see also Sherman 1989; Case Of The Female Orgasm: Bias in the Science of Evolution).

Out-of-Date?

Much of Symons’ work is dedicated to challenging the naïve group-selectionism of Sixties ethologists, especially Desmond Morris. Although scientifically now largely obsolete, Morris’s work still retains a certain popular resonance and therefore this aspect of Symons’s work is not entirely devoid of contemporary relevance.

In place of Morris‘s rather idyllic notion that humans are a naturally monogamous ‘pair-bonding’ species, Symons advocates instead an approach rooted in the individual-level (or even gene-level) selection championed Richard Dawkins in The Selfish Gene (reviewed here).

This leads to some decidedly cynical conclusions regarding the true nature of sexual and romantic relations among humans.

For example, Symons argues that it is adaptive for men to be less sexually attracted to their wives than they are to other women – because they are themselves liable to bear the cost of raising offspring born to their wives but not those born to other women with whom they mate (e.g. those attached to other males).

Another cynical conclusion is that the primary emotion underlying the institution of marriage, both cross-culturally and in our own society, is neither love nor even lust, but rather male sexual jealousy and proprietariness (p123). 

Marriage, then, is an institution borne not of love, but of male sexual jealousy and the behaviour known to biologists as mate-guarding.

Meanwhile, in his excellent chapter on ‘Copulation as a Female Service’ (Chapter Eight), Symons suggests that many aspects of heterosexual romantic relationships may be analogous to prostitution.

As well as its excessive focus on debunking sixties ethologists like Morris, ‘The Evolution of Human Sexuality’ is also out-of-date in a more serious respect Namely, it fails to incorporate the vast amount of empirical research on human sexuality from a sociobiological perspective which has been conducted since the first publication of his work.

For a book first published thirty years ago, this is inevitable – not least because much of this empirical research was inspired by Symons’ own ideas and specifically designed to test theories formulated in this very work.

In addition, potentially important new factors in human reproductive behaviour that even Symons did not foresee have been identified, for example role of levels of fluctuating asymmetry functioning as a criterion for, or at least correlate of, physical attractiveness.

For an updated discussion of the evolutionary psychology of human sexual behaviour, complete with the latest empirical data and research, readers should consult the latest edition of David Buss’s The Evolution Of Desire: Strategies of Human Mating.

In contrast, in support of his theories Symons relies largely on classical literary insight, anecdote and, most importantly, a review of the ethnographic record.

However, this latter focus ensures that, in some respects, the work remains of more than merely of historical interest.

After all, one of the more legitimate criticisms levelled against recent research in evolutionary psychology is that it is insufficiently cross-cultural and, with several notable exceptions (e.g. Buss 1989), relies excessively on research conducted among convenience samples of students at western universities.

Given costs and practicalities, this is inevitable. However, for a field that aspires to understand a human nature presumed to be universal, such a method of sampling is highly problematic, especially given what has recently been revealed about the ‘WEIRD-ness’ of western undergraduate samples.

The Evolution of Human Sexuality’ therefore retains its importance for two reasons. 

First, is it the founding work of modern evolutionary psychological research into human sexual behaviour, and hence of importance as a landmark and classic text in the field, as well as in the history of science more generally. 

Second, it also remains of value to this day for the cross-cultural and ethnographic evidence it marshals in support of its conclusions. 

Endnotes

[1] Actually, the first person to discover this, albeit inadvertently, was the great Victorian polymath, pioneering statistician and infamous eugenicist Francis Galton, who, attempting to discover abnormal facial features possessed by the criminal class, succeeded in morphing the faces of multiple convicted criminals. The result was, presumably to his surprise, an extremely attractive facial composite, since all the various minor deformities of the many convicted criminals whose faces he morphed actually balanced one another out to produce a face with few if any abnormalities or disproportionate features.

[2] More recent research in this area has focused on the related concept of fluctuating asymmetry.

[3] However, recent meta-analyses have called into question the evidence for cyclical fluctuations in female mate preferences (Wood et al 2014; cf. Gildersleeve et al 2014), and it has been suggested that such findings may represent casualties of the so-called replication crisis in psychology. It has also been questioned whether ovulation in humans is indeed concealed, or is actually detectable by subtle cues (e.g. Miller et al 2007), for example, changes in face shape (Oberzaucher et al 2012), breast symmetry (Scutt & Manning 1996) and body scent (Havlicek et al 2006).

[4] Another factor leading recent researchers to overestimate the importance of female choice in human evolution is their feminist orientation, since female choice gives women an important role in human evolution, even, paradoxically, in the evolution of male traits.

[5] Actually, in most cultures, only a girl’s first marriage is arranged on her behalf by her parents. Second- and third-marriages are usually negotiated by the woman herself. However, since female fertility peaks early, it is a girl’s first marriage that is usually of the most reproductive, and hence Darwinian, significance.

[6] Indeed, the human anatomical trait in humans that perhaps shows the most evidence of being a product of intersexual selection is a female one, namely the female breasts, since the latter are, unlike the mammary glands of most other mammals, permanently present from puberty on, not only during lactation, and composed primarily of fatty tissues, not milk (Møller 1995; Manning et al 1997; Havlíček et al 2016). 

[7] Wilson terms his theory “the kin selection theory hypothesis of the origin of homosexuality” (p145). However, a better description might be the ‘helper at the nest theory of homosexuality’, the basic idea being that, like sterile castes in some insects, and like older siblings in some bird species where new nest sites are unavailable, homosexuals, rather than reproducing themselves, direct their energies towards assisting their collateral kin in successfully raising, and provisioning, their own offspring (On Human Nature: p143-7). The main problem with this theory is that there is no evidence that homosexuals do indeed devote any greater energies towards assisting their kin in raising offspring. On the contrary, homosexuals instead seem to devote much of their time and resources towards their own sex life, much as do heterosexuals (Bobrow & Bailey 2001).

[8] As we will see, contrary to the stereotype of evolutionary psychologists as viewing all traits as necessarily adaptive, as they are accused of doing by the likes of Gould, Symons also argued that the female orgasm and menopause are non-adaptive, but rather by-products of other adaptations.

[9] This is not necessarily to say that rampant, indiscriminate promiscuity is a male utopia, or the ideal of any man, be he homosexual or heterosexual. On the contrary, the ideal mating system for any individual male is harem polygyny in which the chastity of his own partners is rigorously policed (see Laura Betzig’s Despotism and Differential Reproduction: which I have reviewed here). However, given an equal sex ratio, this would condemn other males to celibacy and perpetual ‘inceldom. Similarly, Symons reports that “Homosexual men, like most people, usually want to have intimate relationships”. However, he observes:

Such relationships are difficult to maintain, largely owing to the male desire for sexual variety; the unprecedented opportunity to satisfy this desire in a world of men, and the male tendency towards sexual jealousy” (p297).  

It does indeed seem to be true that homosexual relationships, especially those of gay males, are, on average, of shorter duration than are heterosexual relationships. However, Symons’ claim regarding “the male tendency towards sexual jealousy” is questionable.
Actually, subsequent research in evolutionary psychology has suggested that men are no more prone to jealousy than women, but rather that it is sorts of behaviours which most intensely provoke such jealousy that differentiate the sexes (Buss 1992). Moreover, many gay men practice open relationships, which seems to suggest a lack of jealousy – or perhaps this simply reflects a recognition of the difficulty of maintaining relationships given, as Symons puts it, “the male desire for sexual variety [and] the unprecedented opportunity to satisfy this desire in a world of men”. 

[10] Indeed, far from men being led to objectify women due to the portrayal of women in a sexualized manner in the media, Symons suggests:

There may be no positive feedback at all; on the contrary, constant exposure to pictures of nude and nearly nude female bodies may to some extent habituate [i.e. desensitize] men to these stimuli” (p304).

[11] Admittedly, some aspects of body-type typically preferred by gay males (especially the so-called twink ideal) do reflect apparently female traits, especially a relative lack of body-hair. However, lack of body-hair is also obviously indicative of youth. Moreover, a relative lack of body-hair also seems to be a trait favoured in men by heterosexual women. For a discussion of the relative preference on the part of (heterosexual) females for masculine versus feminine physical appearance in male sex partners, see here.

[12] Thus, some men might indeed welcome being ‘raped’, albeit only under highly unusual circumstances – namely by an attractive opposite-sex partner (or, in the case of homosexual men, an attractive same-sex partner) to whom they are sexually attracted. Thus, Kingsley Browne, in his excellent Biology at Work (which I have reviewed here) quotes the perhaps remarkable finding that:

A substantial number of men ‘viewed an advance by a good-looking woman who threatened harm or held a knife as a positive sexual opportunity’” (Biology at Work: p196; quoting Struckman-Johnson & Struckman-Johnson 1994).

Of course, large numbers of women also report rape fantasies (Bivona & Critelli 2009). Yet this does not, of course, mean they would actually welcome real sexual assault, which would almost certainly take a very different form from the fantasy. In practice, therefore, members of neither sex are ever likely to welcome sexual assault in the form which it is actually likely to actually come.

[13] Incidentally, Symons also rejects the theory that the female menopause is adaptive, a theory which has subsequently become known as the grandmother hypothesis (p13). Also, although it does not directly address the issue, Symons’ discussion of human rape (p276-85), has also been interpreted as implicitly favouring the theory that rape is a by-product of the greater male desire for commitment free promiscuous sex, rather than the product of a specific rape adaptation in males (see Palmer 1991; and A Natural History of Rape: reviewed here). 

References 

Bellis & Baker (1990). Do females promote sperm competition?: Data for humans. Animal Behavior, 40: 997-999.
Bivona & Critelli 2009 The nature of women’s rape fantasies: an analysis of prevalence, frequency, and contents. Journal of Sex Research 46(1):33-45
Clark & Hatfield (1989) Gender differences in receptivity to sexual offers Journal of Psychology and Human Sexuality 2(1):39-55
Bobrow & Bailey (2001). Is male homosexuality maintained via kin selection? Evolution and Human Behavior, 22: 361-368.
Bogaert & Hershberger (1999) The relation between sexual orientation and penile size. Archives of Sexual Behavior 1999 Jun;28(3) :213-21. 
Buss (1989). Sex differences in human mate preferences: Evolutionary hypotheses tested in 37 cultures. Behavioral and Brain Sciences 12: 1-49
Ellis & Ratnasingam (2012) Gender, Sexual Orientation, and Occupational Interests: Evidence of Androgen Influences. Mankind Quarterly  53(1): 36–80
Ellis & Symons (1990) Sex differences in sexual fantasy: An evolutionary psychological approach, Journal of Sex Research 27(4): 527-555.
Gildersleeve, Haselton & Fales (2014) Do women’s mate preferences change across the ovulatory cycle? A meta-analytic review. Psychological Bulletin 140(5):1205-59.
Havlíček, Dvořáková, Bartos & Fleg (2006) Non‐Advertized does not Mean Concealed: Body Odour Changes across the Human Menstrual Cycle. Ethology 112(1):81-90.
Havlíček et al (2016) Men’s preferences for women’s breast size and shape in four cultures. Evolution and Human Behavior 38(2): 217–226.
Kenrick & Keefe (1992). Age preferences in mates reflect sex differences in human reproductive strategies. Behavioral and Brain Sciences, 15: 75-133. 
Kruger et al (2003) Proper and Dark Heroes as Dads and Cads. Human Nature 14(3): 305-317.
Manning et al (1997) Breast asymmetry and phenotypic quality in women. Ethology and Sociobiology 18(4): 223–236.
Miller (1998). How mate choice shaped human nature: A review of sexual selection and human evolution. In C. Crawford & D. Krebs (Eds.), Handbook of Evolutionary Psychology: Ideas, Issues, and Applications (pp. 87-129). Mahwah, NJ: Lawrence Erlbaum.
Miller, Tybur & Jordan (2007). Ovulatory cycle effects on tip earnings by lap dancers: economic evidence for human estrous? Evolution and Human Behavior. 28(6):375–381.
Møller et al (1995) Breast asymmetry, sexual selection, and human reproductive success. Ethology and Sociobiology 16(3): 207-219.
Palmer (1991) Human Rape: Adaptation or By-Product? Journal of Sex Research 28(3): 365-386.
Penton-Voak et al (1999) Menstrual cycle alters face preferences, Nature 399 741-2.
Puts (2010) Beauty and the Beast: Mechanisms of Sexual Selection in Humans. Evolution and Human Behavior 31 157-175.
Salmon (2004) The Pornography Debate: What Sex Differences in Erotica Can Tell Us About Human Sexuality. In Evolutionary Psychology, Public Policy and Personal Decisions (London: Lawrence Erlbaum Associates, 2004).
Scutt & Manning (1996) Symmetry and ovulation in women. Human Reproduction 11(11):2477-80.
Sherman (1989) The clitoris debate and levels of analysis, Animal Behaviour, 37: 697-8.
Struckman-Johnson & Struckman-Johnson (1994) Men’s reactions to hypothetical female sexual advances: A beauty bias in response to sexual coercion. Sex Roles 31(7-8): 387–405.
Wood et al (2014). Meta-analysis of menstrual cycle effects on women’s mate preferencesEmotion Review, 6(3), 229–249.

Judith Harris’s ‘The Nurture Assumption’: By Parent or Peers

Judith Harris, The Nurture Assumption: Why Children Turn Out the Way They Do. Free Press, 1998.

Almost all psychological traits on which individual humans differ, from personality and intelligence to mental illness, are now known to be substantially heritable. In other words, individual differences in these traits are, at least in part, a consequence of genetic differences between individuals.

This finding is so robust that it has even been termed by Eric Turkenheimer the First Law of Behviour Genetics and, although once anathema to most psychologists save a marginal fringe of behavioural geneticists, it has now, under the sheer weight of evidence produced by the latter, belatedly become the new orthodoxy. 

On reflection, however, this transformation is not entirely a revelation. 

After all, it was only in the mid-twentieth century that the curious notion that individual differences were entirely the product of environmental differences first arose, and, even then, this delusion was largely restricted to psychologists, sociologists, feminists and other such ‘professional damned fools’, along with those among the semi-educated public who seek to cultivate an air of intellectualism by aping the former’s affections. 

Before then, poets, peasants and laypeople alike had long recognized that ability, insanity, temperament and personality all tended to run in families, just as physical traits like stature, complexion, hair and eye colour also do.[1]

However, while the discovery of a heritable component to character and ability merely confirms the conventional wisdom of an earlier age, another behavioural genetic finding, far more surprising and counterintuitive, has passed relatively unreported. 

This is the discovery that the so-called shared family environment (i.e. the environment shared by siblings, or non-siblings, raised in the same family home) actually has next to no effect on adult personality and behaviour. 

This we know from such classic study designs in behavioural genetics as twin studies, adoption studies and family studies.

In short, individuals of a given degree of relatedness, whether identical twins, fraternal twins, siblings, half-siblings or unrelated adoptees, are, by the time they reach adulthood, no more similar to one another in personality or IQ when they are raised in the same household than when they are raised in entirely different households. 

The Myth of Parental Influence 

Yet parental influence has long loomed large in virtually every psychological theory of child development, from the Freudian Oedipus complex and Bowby’s attachment theory to the whole literary genre of books aimed at instructing anxious parents on how best to raise their children so as to ensure that the latter develop into healthy, functional, successful adults. 

Indeed, not only is the conventional wisdom among psychologists overturned, but so is the conventional wisdom among sociologists – for one aspect of the shared family environment is, of course, household income and social class

Thus, if the family that a person is brought up in has next to no impact on their psychological outcomes as an adult, then this means that the socioeconomic status of the family home in which they are raised also has no effect. 

Poverty, or a deprived upbringing, then, has no effect on IQ, personality or the prevalence of mental illness, at least by the time a person has reached adulthood.[2]

Neither is it only leftist sociologists who have proved mistaken. 

Thus, just as leftists use economic deprivation as an indiscriminate, catch-all excuse for all manner of social pathology (e.g. crime, unemployment, educational underperformance) so conservatives are apt to place the blame on divorce, family breakdown, having children out of wedlock and the consequential increase in the prevalence of single-parent households

However, all these factors are, once again, part of the shared family environment – and according to the findings of behavioural genetics, they have next to no influence on adult personality or intelligence. 

Of course, chaotic or abusive family environments do indeed tend to produce offspring with negative life outcomes. 

However, none of this proves that it was the chaotic or abusive family environment that caused the negative outcomes. 

Rather, another explanation is at hand – perhaps the offspring simply biologically inherit the personality traits of their parents, the very personality traits that caused their family environment to be so chaotic and abusive in the first place.[3] 

For example, parents who divorce or bear offspring out-of-wedlock likely differ in personality from those who first get married then stick together, perhaps being more impulsive or less self-disciplined and conscientious (e.g. less able refrain from having children from a relationship that was destined to be fleeting, or less able to persevere and make the relationship last). 

Their offspring may, then, simply biologically inherit these undesirable personality attributes, which then themselves lead to the negative social outcomes associated with being raised in single-parent households or broken homes. The association between family breakdown and negative outcomes for offspring might, then, reflect simply the biological inheritance of personality. 

Similarly, as leftists are fond of reminding us, children from economically-deprived backgrounds do indeed have lower recorded IQs and educational attainment than those from more privileged family backgrounds, as well as other negative outcomes as adults (e.g. lower earnings, higher rates of unemployment). 

However, this does not prove that coming from a deprived family background necessarily itself depresses your IQ, educational attainment or future salary. 

Rather, an equally plausible possibility is simply that offspring simply biologically inherit the low intelligence of their parents – the very low intelligence which was likely a factor causing the low socioeconomic status of their parents, since intelligence is known to correlate strongly with educational and occupational advancement.[4]

In short, the problem with all of this body of research which purports to demonstrate the influence of parents and family background on psychology and behavioural outcomes for offspring is that they fail to control for the heritability of personality and intelligence, an obvious confounding factor

The Non-Shared Environment

However, not everything is explained by heredity. As a crude but broadly accurate generalization, only about half the variation for most psychological traits is attributable to genes. This leaves about half of the variation in intelligence, personality and mental illness to be explained environmental factors.  

What are these environmental factors if they are not to be sought in the shared family environment

The obvious answer is, of course, the non-shared family environment – i.e. the ways in which even children brought up in the same family-home nevertheless experience different micro-environments, both within the home and, perhaps more importantly, outside it. 

Thus, even the fairest and most even-handed parents inevitably treat their different offspring differently in some ways.  

Indeed, among the principal reasons why parents treat their different offspring differently is precisely because the different offspring themselves differ in their own behaviour quite independently of any parental treatment.

This is well illustrated by the question of the relationship between corporal punishment and behaviour in children.

Corporal punishment 

Rather than differences in the behaviour of different children resulting from differences in how their parents treat them, it may be that differences in how parents treat their children may reflect responses to differences in the behaviour of the children themselves. 

In other words, the psychologists have the direction of causation precisely backwards. 

Take, for example, one particularly controversial issue, namely the physical chastisement of children by their parents as a punishment for bad behaviour (e.g. spanking). 

Some psychologists have sometimes argued that physical chastisement actually causes misbehaviour. 

As evidence, they cite the fact that children who are spanked more often by their parents or caregivers on average actually behave worse than those whose caregivers only rarely or never spank the children entrusted to their care.  

This, they claim, is because, in employing spanking as a form of discipline, caregivers are inadvertently imparting the message that violence is a good way of solving your problems. 

Actually, however, I suspect children are more than capable of working out for themselves that violence is often an effective means of getting your way, at least if you have superior physical strength to your adversary. Unfortunately, this is something that, unlike reading, arithmetic and long division, does not require explicit instruction by teachers or parents. 

Instead, a more obvious explanation for the correlation between spanking and misbehaviour in children is not that spanking causes misbehaviour, but rather that misbehaviour causes spanking. 

Indeed, once you think about it, this is in fact rather obvious: If a child never seriously misbehaves, then a parent likely never has any reason to spank that child, even if the parent is, in principle, a strict disciplinarian; whereas, on the other hand, a highly disobedient child is likely to try the patience of even the most patient caregiver, whatever his or her moral opposition to physical chastisement in principle. 

In other words, causation runs in exactly the opposite direction to that assumed by the naïve psychologists.[5] 

Another factor may also be at play – namely, offspring biologically inherit from their parents the personality traits that cause both the misbehaviour and the punishment. 

In other words, parents with aggressive personalities may be more likely to lose their temper and physically chastise their children, while children who inherit these aggressive personalities are themselves more likely to misbehave, not least by behaving in an aggressive or violent manner. 

However, even if parents treat their different offspring differently owing to the different behaviour of the offspring themselves, this is not the sort of environmental factor capable of explaining the residual non-shared environmental effects on offspring outcomes. 

After all, this merely begs the question as to what caused these differences in offspring behaviour in the first place? 

If the differences in offspring behaviour exist prior to differences in parental responses to this behaviour, then these differences cannot be explained by the differences in parental responses.  

Peer Groups 

This brings us back to the question of the environmental causes of offspring outcomes – namely, if about half the differences among children’s IQs and personalities are attributable to environmental factors, but these environmental factors are not to be found in the shared family environment (i.e. the environment shared by children raised in the same household), then where are these environmental factors to be sought? 

The search for environmental factors affecting personality and intelligence has, thus far, been largely unsuccessful. Indeed, some behavioural geneticists have almost gone as far as conceding scholarly defeat in identifying correlates for the environmental portion of the variance. 

Thus, leading contemporary behavioural geneticist Robert Plomin in his recent book, Blueprint: How DNA Makes Us Who We Are, concludes that those environmental factors that affect cognitive ability, personality, and the development of mental illness are, as he puts it, ‘unsystematic’ in nature. 

In other words, he seems to be saying that they are mere random noise. This is tantamount to accepting that the null hypothesis is true. 

Judith Harris, however, has a quite different take. According to Harris, environmental causes must be sought, not within the family home, but rather outside it – in a person’s interactions with their peer-group and the wider community.[6]

Environment ≠ Nurture 

Thus, Harris argues that the so-called nature-nurture debate is misnamed, since the word ‘nurture’ usually refers to deliberate care and moulding of a child (or of a plant or animal). But many environmental effects are not deliberate. 

Thus, Harris repeatedly references behaviourist John B. Watson’s infamous boast: 

Give me a dozen healthy infants, well-formed, and my own specified world to bring them up in and I’ll guarantee to take any one at random and train him to become any type of specialist I might select—doctor, lawyer, artist, merchant-chief and, yes, even beggar-man and thief, regardless of his talents, penchants, tendencies, abilities, vocations, and race of his ancestors.

Yet what strikes me as particularly preposterous about Watson’s boast is not its radical environmental determinism, nor even its rather convenient unfalsifiability.[7] 

Rather, what most strikes me as most preposterous about Watson’s claim is its frankly breath-taking arrogance. 

Thus, Watson not only insisted that it was environment alone that entirely determined adult personality. In this same quotation, he also proclaimed that he already fully understood the nature of these environmental effects to such an extent that, given omnipotent powers to match his evidently already omniscient understanding of human development, he could produce any outcome he wished. 

Yet, in reality, environmental effects are anything but clear-cut. Pushing a child in a certain direction, or into a certain career, may sometimes have the desired effect, but other times may seemingly have the exact opposite effect to that desired, provoking the child to rebel against parental dictates. 

Thus, even to the extent that environment does determine outcomes, the precise nature of the environmental factors implicated, and their interaction with one another, and with the child’s innate genetic endowment, is surely far more complex than the simple mechanisms proposed by behaviourists like Watson (e.g. reinforcement and punishment). 

Language Acquisition 

The most persuasive evidence for Harris’s theory of the importance of peer groups comes from an interesting and widely documented peculiarity of language acquisition

The children of immigrants, whose parents speak a different language inside the family home, and may even themselves be monolingual, nevertheless typically grow up to speak the language of their host culture rather better than they do the language to which they were first exposed in the family home. 

Indeed, while their parents may never achieve fluency in the language of their host culture, having missed out on the Chomskian critical period for language acquisition, their children often actually lose the ability to speak their parent’s language, often much to the consternation of parents and grandparents. 

Yet, from an sociobiological or evolutionary psychological perspective, such an outcome is obviously adaptive. 

After all, if a child is to succeed in wider society, they must master its language, whereas, if their parent’s first language is not spoken anywhere in their host society except in their family, then it is of limited utility, and, once their parents themselves become proficient in the language of the host culture, it becomes entirely redundant.

As sociologist-turned-sociobiologist Pierre van den Berghe observes in his excellent The Ethnic Phenomenon (reviewed here):

Children quickly discover that their home language is a restricted medium that not useable in most situations outside the family home. When they discover that their parents are bilingual they conclude – rightly for their purposes – that the home language is entirely redundant… Mastery of the new language entails success at school, at work and in ‘the world’… [against which] the smiling approval of a grandmother is but slender counterweight” (The Ethnic Phenomenon: p258). 

Code-Switching 

Harris suggests that the same applies to personality. Just as the child of immigrants switches between one language and another at home and school, so they also adopt different personalities. 

Thus, many parents are surprised to be told by their children’s teachers at parents’ evenings that their offspring is quiet and well-behaved at school, since, they report, he or she isn’t at all like that at home. 

Yet, at home, a child has only, at most, a sibling or two with whom to compete for his parents’ attention. In contrast, at school, he or she has a whole class with whom to compete for their teacher’s attention.

It is therefore unsurprising that most children are less outgoing at school than they are at home with their parents. 

For example, an older sibling might be able push his little brother around at home. But, if he is small for his age, he is unlikely to be able to get away with the same behaviour among his peers at school. 

Children therefore adopt two quite different personalities – one for interactions with family and siblings, and another for among their peers.

This then, for Harris, explains why, perhaps surprisingly, birth-order has generally been found to have little if any effect on personality, at least as personality manifests itself outside the family home. 

An Evolutionary Theory of Socialization? 

Interestingly, even evolutionary psychologists have not been immune from the delusion of parental influence. Thus, in one influential paper, anthropologists Patricia Draper and Henry Harpending argued that offspring calibrate their reproductive strategy by reference to the presence or absence of a father in their household (Draper & Harpending 1982). 

On this view, being raised in a father-absent household is indicative of a social environment where low male parental investment is the norm, and hence offspring adjust their own reproductive strategy accordingly, adopting a promiscuous, low-investment mating strategy characterized by precocious sexual development and an inability to maintain lasting long-term relationships (Draper & Harpending 1982; Belsky et al 1991). 

There is indeed, as these authors amply demonstrate, a consistent correlation between father-absence during development and both earlier sexual development and more frequent partner-switching in later life. 

Yet there is also another, arguably more obvious, explanation readily at hand to explain this association. Perhaps offspring simply inherit biologically the personality traits, including sociosexual orientation, of their parents. 

On this view, offspring raised in single-parent households are more likely to adopt a promiscuous, low-investment mating strategy simply because they biologically inherit the promiscuous sociosexual orientation of their parents, the very promiscuous sociosexual orientation that caused the latter to have children out-of-wedlock or from relationships that were destined to break down and hence caused the father-absent childhood of their offspring. 

Moreover, even on purely a priori theoretical grounds, Draper, Harpending and Belsky’s reasoning is dubious. 

After all, whether you personally were raised in a one- or two-parent family is obviously a very unreliable indicator of the sorts of relationships prevalent in the wider community into which you are born, since it represents a sample size of just one. 

Instead, therefore, it would be far more reliable to calibrate your reproductive strategy in response to the prevalence of one-parent households in the wider community at large, rather than the particular household type into which you happen to have been born.  

This, of course, directly supports Harris’s own theory of ‘peer group socialization’. 

In short, to the extent that children do adapt to the environment and circumstances of their upbringing (and they surely do), they must integrate into, adopt the norms of, and a reproductive strategy to maximize their fitness within, the wider community into which they are born, rather than the possibly quite idiosyncratic circumstances and attitudes of their own family. 

Absent Fathers, from Upper-Class to Under-Class 

Besides language-acquisition among the children of immigrants, another example cited by Harris in support of her theory of ‘peer group socialization’ is the culture, behaviours and upbringing of British upper-class males.

Here, she reports, boys were, and, to some extent, still are, reared primarily, not by their parents, but rather by nannies, governoresses and, more recently, in exclusive fee-paying all-male boarding schools

Yet, despite having next to no contact with their fathers throughout most of their childhood, these boys nevertheless managed somehow to acquire manners, attitudes and accents similar, if not identical, to those of their upper-class fathers, and not at all those of the middle-class nannies, governoresses and masters with whom they spent most of their childhood being raised. 

Yet this phenomenon is by no means restricted to the British upper-classes.

On the contrary, rather than citing the example of the British upper-classes in centuries gone by, Harris might just as well have cited that of contemporary underclass in Britain and America, since what was once true of the British upper-classes, is now equally true of the underclass

Just as the British upper-classes were once raised by governoresses, nannies and in private schools with next to no contact with their fathers, so contemporary underclass males are similarly raised in single-parent households, often to unwed mothers, and typically have little if any contact with their biological fathers. 

Here, as Warren Farrell observes in his seminal The Myth of Male Power (which I have reviewed here, here and here), there is a now a “a new nuclear family: woman, government and child”, what Farrell terms “Government as a Substitute Husband”. 

Yet, once again, these underclass males, raised by single parents with the financial assistance of the taxpayer, typically turn out much like their absent fathers with whom they have had little if any contact, often going on to promiscuously father a succession of offspring themselves, with whom they likewise have next to no contact. 

Abuse 

But what of actual abuse? Surely this has a long-term devastating psychological impact on children. This, at any rate, is the conventional wisdom, and questioning this wisdom, at least with respect to sexual abuse, is tantamount to contemporary heresy, with attendant persecution

Thus, for example, it is claimed that criminals who are abusive towards their children were themselves almost invariably abused, mistreated or neglected as children, which is what has led to their own abusive, behaviour.

A particularly eloquent expression of this theory is found in the novel Clockers, by Richard Price, where one of the lead characters, a police officer, explains how, during his first few years on the job, a senior colleague had restrained him from attacking an abusive mother who had left her infant son handcuffed to a radiator, telling him:

Rocco, that lady you were gonna brain? Twenty years ago when she was a little girl. I arrested her father for beating her baby brother to death. The father was a piece of shit. Now that she’s all grown up? She’s a real piece of shit. That kid you saved today. If he lives that long, if he grows up? He’s gonna be a real piece of shit. It’s the cycle of shit and you can’t do nothing about it” (Clockers: p96).

Take, for example, what is perhaps the form of child abuse that provokes the most outrage and disgust – namely, sexual abuse. Here, it is frequently asserted that paedophiles were almost invariably themselves abused as children, which creates a so-called cycle of abuse

However, there are at least three problems with this claim. 

First, it cannot explain how the first person in this cycle came to be abusive. 

Second, we might doubt whether it is really true that paedophiles are disproportionately likely to have themselves been abused as children. After all, abuse is something that almost invariably happens surreptitiously ‘behind closed doors’ and is therefore difficult to verify or disprove. 

Therefore, even if most paedophiles claim to have been victims of abuse, it is possible that they are simply lying in order to elicit sympathy or excuse or shift culpability for their own offending. 

Finally, and most importantly for present purposes, even if paedophiles can be shown to be disproportionately likely to have themselves been victimized as children, this by no means proves that their past victimization caused their current sexual orientation. 

Rather, since most abuse is perpetrated by parents or other close family members, an alternative possibility is that victims simply biologically inherit the sexual orientation of their abuser.

After all, if homosexuality is partially heritable, as is now widely accepted, then why not paedophilia as well? 

In short, the ‘cycle of shit’ referred to by Price’s fictional police officer may well be real, but mediated by genetics rather than childhood experience.

However, this conclusion is not entirely clear. On the contrary, Harris is at pains to emphasize that the finding that the shared family environment accounts for hardly any of the variance in outcomes among adults does not preclude the possibility that severe abuse may indeed have an adverse effect on adult outcomes. 

After all, adoption studies can only tell us what percent of the variance is caused by heredity or by shared or unshared environments within a specific population as a whole.

Perhaps the shared family environment accounts for so little of the variance precisely because the sort of severe abuse that does indeed have a devastating long-term effect on personality and mental health is, thankfully, so very rare in modern societies. 

Indeed, it may be especially rare within the families sampled in adoption studies precisely because adoptive families are carefully screened for suitability before being allowed to adopt. 

Moreover, Harris emphasizes an important caveat: Even if abuse does not have long-term adverse psychological effects, this does not mean that abuse causes no harm, and nor does it in any way excuse such abuse. 

On the contrary, the primary reason we shouldn’t mistreat children (and should severely punish those who do) is not on account of some putative long-term psychological effect on the adults whom the children subsequently become, but rather because of the very real pain and suffering inflicted on a child at the time the abuse takes place. 

Race Differences in IQ 

Finally, Harris even touches upon that most vexed area of the (so-called) nature-nurture debate – race differences in intelligence

Here, the politically-correct claim that differences in intelligence between racial groups, as recorded in IQ tests, are of purely environmental origin runs into a problem, since the sorts of environmental effects that are usually posited by environmental determinists as accounting for the black-white test score gap in America (e.g. differences in rates of poverty and socioeconomic status) have been shown to be inadequate because, even after controlling for these factors, there remains a still unaccounted for gap in test-scores.[8]

Thus, as Arthur R. Jensen laments: 

This gives rise to the hypothesizing of still other, more subtle environmental factors that either have not been or cannot be measured—a history of slavery, social oppression, and racial discrimination, white racism, the ‘black experience,’ and minority status consciousness [etc]” (Straight Talk About Mental Tests: p223). 

The problem with these explanations, however, is that none of these factors has yet been demonstrated to have any effect on IQ scores. 

Moreover, some of the factors proposed as explanations are formulated in such a vague form (e.g. “white racism, the ‘black experience’”) that it is difficult to conceive of how they could ever be subjected to controlled testing in the first place.[9]

Jensen has termed this mysterious factor the X-factor

In coining this term, Jensen was emphasizing its vague, mysterious and unfalsifiable nature. Jensen did not actually believe that this posited X-factor, whatever it was, really did account for the test-score gap. Rather, he thought heredity explained most, if not all, of the remaining unexplained test-score gap. 

However, Harris takes Jensen at his word and takes the search for the X-factor very seriously. Indeed, she apparently believes she has discovered and identified it. Thus, she announces: 

I believe I know what this X factor is… I can describe it quite clearly. Black kids and white kids identify with different groups that have different norms. The differences are exaggerated by group contrast effects and have consequences that compound themselves over the years. That’s the X factor” (p248-9). 

Unfortunately, Harris does not really develop this fascinating claim. Indeed, she cites no direct evidence in support of this claim, and evidently seems to regard the alternative possibility – namely, that race differences in intelligence are at least partly genetic in origin – as so unpalatable that it can safely ruled out a priori.

In fact, however, although not discussed by Harris, there is at least some evidence in support of her theory. Indeed, her theory potentially reconciles the apparently conflicting findings of two of the most widely-cited studies in this vexed area of research and debate.

First, in the more recent of these two studies, Minnesota Transracial Adoption Study, the same race differences in IQ were observed among black, white and mixed-race children adopted into upper-middle class white families as are found among black, white and mixed-race populations in the community at large (Scarr & Weinberg 1976). 

Moreover, although, when tested during childhood, the children’s adoptive households did seem to have had a positive effect on their IQ scores, in a follow-up study it was found that by the time they reached the cusp of adulthood, the black teenagers who had been adopted into upper-middle-class white homes actually scored no higher in IQ than did blacks in the wider population not raised in upper-middle class white families (Weinberg, Scarr & Waldman 1992). 

Although Scarr, Weinberg and Waldman took pains to present their findings as compatible with a purely environmentalist theory of race differences, this study has, not unreasonably, been widely cited by hereditarians as evidence for the existence of innate racial differences in intelligence (e.g. Levin 1994; Lynn 1994; Whitney 1996).

However, in the light of the findings of the behavioural genetics studies discussed by Harris in ‘The Nurture Assumption’, the fact that white upper-middle-class adoptive homes had no effect on the adult IQs of the black children adopted into them is, in fact, hardly surprising. 

After all, as we have seen, the shared family environment generally has no effect on IQ, at least by the time the person being tested has reached adulthood.[10]

One would therefore not expect adoptive homes, howsoever white and upper-middle-class, to have any effect on adult IQs of the black children adopted into them, or indeed of the white or mixed-race children adopted into them. 

In short, adoptive homes have no effect on adult IQ, whether or not the adoptees, or adoptive families, are black, white, brown, yellow, green or purple! 

But, if race differences in intelligence are indeed entirely environmental in origin, then where are these environmental causes to be found, if not in the family environment? 

Harris has an answer – black culture

According to her, the black adoptees, although raised in white adoptive families, nevertheless still come to identify as ‘black’, and to identify with the wider black culture and social norms. In addition, they may, on account of their racial identification, come to socialize with other blacks in school and elsewhere. 

As a result of this acculturation to African-American norms and culture, they therefore, according to Harris, come to score lower in IQ than their white peers and adoptive siblings. 

But how can we ever test this theory? Is it not untestable, and is this not precisely the problem identified by Jensen with previous positedX-factors.

Actually, however, although not discussed by Harris, there is a way of testing this theory – namely, looking at the IQs of black children raised in white families where there is no wider black culture with which to identify, and few if any black peers with whom to socialize?

This, then, brings us to the second of the two studies which Harris’s theory potentially reconciles, namely the famous Eyferth study.  

Here, it was found that the mixed-race children fathered by black American servicemen who had had sexual relationships with German women during the Allied occupation of Germany after World War Two had almost exactly the same average IQ scores as a control group of offspring fathered by white US servicemen during the same time period (Eyferth 1959). 

The crucial difference from the Minnesota study may be that these children, raised in an almost entirely monoracial, white Germany in the mid-twentieth century, had no wider African-American culture with which to identify or whose norms to adopt, and few if any black or mixed-race peers in their vicinity with whom to socialize. 

This, then, is perhaps the last lifeline for a purely environmentalist theory of race differences in intelligence – namely the theory that African-American culture depresses intelligence. 

Unfortunately, however, this proposition – namely, that African-American culture depresses your IQ – is almost as politically unpalatable and politically-incorrect as is the notion that race differences in intelligence reflect innate genetic differences.[11]

Endnotes

[1] Thus, this ancient wisdom is reflected, for example, in many folk sayings, such as the apple does not fall far from the tree, a chip off the old block and like father, like son, many of which long predate either Darwin’s theory of evolution, and Mendel’s work on heredity, let alone the modern work of behavioural geneticists.

[2] It is important to emphasize here that this applies only to psychological outcomes, and not, for example, economic outcomes. For example, a child raised by wealthy parents is indeed likely to be wealthier than one raised in poverty, if only because s/he is likely to inherit (some of) the wealth of his parents. It is also possible that s/he may, on average, obtain a better job as a consequence of the opportunities opened by his privileged upbringing. However, his IQ will be no higher than had s/he been raised in relative poverty, and neither will s/he be any more or less likely to suffer from a mental illness

[3] Similarly, it is often claimed that children raised in care homes, or in foster care, tend to have negative life-outcomes. However, again, this by no means proves that it is care homes or foster care that causes these negative life-outcomes. On the contrary, since children who end up in foster care are typically either abandoned by their biological parents, or forcibly taken from their parents by social services on account of the inadequate care provided by the latter, or sometimes outright abuse, it is obvious that their parents represent an unrepresentative sample of society as a whole. An obvious alternative explanation, then, is that the children in question simply inherit the dysfunctional personality attributes of their biological parents, namely the very dysfunctional personality attributes that caused the latter to either abandon their children or have them removed by the social services. (In other cases, such children may have been orphaned. However, this is less common today. At any rate, parents who die before their offspring reach maturity are surely also unrepresentative of parents in general. For example, many may live high-risk lifestyles that contribute to their early deaths.)

[4] Likewise, the heritability of such personality traits as conscientiousness and self-discipline, in addition to intelligence, likely also partly account for the association between parental income and academic attainment among their offspring, since both academic attainment, and occupational success, require the self-discipline to work hard to achieve success. These factors, again in addition to intelligence, likely also contribute to the association between parental income and the income and socioeconomic status ultimately attained by their offspring.

[5] This possibility could, at least in theory, be ruled out by longitudinal studies, which could investigate whether the spanking preceded the misbehaviour, or vice versa. However, this is easier said than done, since, unless relying on the reports by caregivers or children themselves, which depends on both the memory and honesty of the caregivers and children themselves, it would have to involve intensive, long-term, and continued observation in order to establish which came first, namely the pattern of misbehaviour, or the adoption of physical chastisement as a method of discipline. This would, presumably, require continuous observation from birth onwards, so as to ensure that the very first instance of spanking or excessive misbehaviour were recorded. Such a study would seem all but impossible and certainly, to my knowledge, has yet to be conducted.

[6] The fact that the relevant environmental variables must be sought outside the family home is one reason why the terms ‘between-family environment’ and ‘within-family environment’, sometimes used as synonyms or alternatives for ‘shared’ and ‘non-shared family environment’ respectively, are potentially misleading. Thus, the ‘within-family environment’ refers to those aspects of the environment that differ for different siblings even within a single family. However, these factors may differ within a single family precisely because they occur outside, not within, the family itself. The terms ‘shared’ and ‘non-shared family environment’ are therefore to be preferred, so as to avoid any potential confusion these alternative terms could cause.

[7] Both practical and ethical considerations, of course, prevent Watson from actually creating his “own specified world” in which to bring up his “dozen healthy infants”. Therefore, no one is able to put his claim to the test. It is therefore unfalsifiable and Watson is therefore free to make such boasts, safe in the knowledge that there is no danger of his actually being called to make good on his claims and thereby proven wrong.

[8] Actually, even if race differences in IQ are found to disappear after controlling for socioeconomic status, it would be a fallacy to conclude that this means that the differences in IQ are entirely a result of differences in social class and that there is no innate difference in intelligence between the races. After all, differences in socioeconomic status are in large part a consequence of differences in cognitive ability, as more intelligent people perform better at school, and at work, and hence rise in socioeconomic status. Therefore, in controlling for socioeconomic status, one is, in effect, also controlling for differences in intelligence, since the two are so strongly correlated. The contrary assumption has been termed by Jensenthe sociologist’s fallacy’.
This fallacy involves the assumption that it is differences in socioeconomic status that cause differences in IQ, rather than differences in intelligence that cause differences in socioeconomic status. As Arthur Jensen explains it:

If SES [i.e. socioeconomic status] were the cause of IQ, the correlation between adults’ IQ and their attained SES would not be markedly higher than the correlation between children’s IQ and their parents’ SES. Further, the IQs of adolescents adopted in infancy are not correlated with the SES of their adoptive parents. Adults’ attained SES (and hence their SES as parents) itself has a large genetic component, so there is a genetic correlation between SES and IQ, and this is so within both the white and the black populations. Consequently, if black and white groups are specially selected so as to be matched or statistically equated on SES, they are thereby also equated to some degree on the genetic component of IQ” (The g Factor: p491).

[9] Actually, at least some of these theories are indeed testable and potentially falsifiable. With regard to the factors quoted by Jensen (namely, “a history of slavery, social oppression, and racial discrimination, white racism… and minority status consciousness”), one way of testing these theories is to look at test scores in those countries where there is no such history. For example, in sub-Saharan Africa, as well as in Haiti and Jamaica, blacks are in the majority, and are moreover in control of the government. Yet the IQ scores of the indigenous populations of sub-Saharan Africa are actually even lower than among blacks in the USA (see Richard Lynn’s Race Differences in Intelligence: reviewed here). True, most such countries still have a history of racial oppression and discrimination, albeit in the form of European colonialism rather than racial slavery or segregation in the American sense. However, in those few sub-Saharan African countries that were not colonized by western powers, or only briefly colonized (e.g. Ethiopia, Liberia), scores are not any higher. Also, other minority groups ostensibly or historically subject to racial oppression and discrimination (e.g. Ashkenazi Jews, Overseas Chinese) actually score higher in IQ than the host populations that ostensibly oppress them. As for “the ‘black experience’”, this meanly begs the question as to why the ‘black experience’ has been so similar, and resulted in the same low IQs in so many different parts of the world, something implausible unless unless the ‘black experience’ itself reflects innate aspects of black African psychology. 

[10] The fact that the heritability of intelligence is higher in adulthood than during childhood, and the influence of the shared family environment correspondingly decreases, has been interpreted as reflecting the fact that, during childhood, our environments are shaped, to a considerable extent, by our parents. For example, some parents may encourage activities that may conceivably enhance intelligence, such as reading books and visiting museums. In contrast, as we enter adulthood, we begin to have freedom to choose and shape our own environments, in accordance with our interests, which may be partly a reflection of our heredity.
Interestingly, this theory suggests that what is biologically inherited is not necessarily intelligence itself, but rather a tendency to seek out intelligence-enhancing environments, i.e. intellectual curiosity rather than intelligence as such. In fact, it is probably a mixture of both factors. Moreover, intellectual curiosity is surely strongly correlated with intelligence, if only because it requires a certain level of intelligence to appreciate intellectual pursuits, since, if one lacks the ability to learn or understand complex concepts, then intellectual pursuits are necessarily unrewarding.

[11] Thus, ironically, the recently deceased James Flynn, though always careful, throughout his career, to remain on the politically-correct radical environmentalist side of the debate with regard to the causes of race differences in intelligence, nevertheless recently found himself taken to task by the leftist, politically-correct British Guardian newspaper for a sentence in his recent book, Does Your Family Make You Smarter, where he described American blacks as coming from a “from a cognitively restricted subculture” (Wilby 2016). Thus, whether one attributes lower black IQs to biology or to culture, either answer is certain offend leftists, and the power of political correctness can, it seems, never be appeased.

References 

Belsky, Steinberg & Draper (1991) Childhood Experience, Interpersonal Development, and Reproductive Strategy: An Evolutionary Theory of Socialization Child Development 62(4): 647-670 

Draper & Harpending (1982) Father Absence and Reproductive Strategy: An Evolutionary Perspective Journal of Anthropological Research 38:3: 255-273 

Eyferth (1959) Eine Untersuchung der Neger-Mischlingskinder in Westdeutschland. Vita Humana, 2, 102–114

Levin (1994) Comment on Minnesota Transracial Adoption Study. Intelligence. 19: 13–20

Lynn, R (1994) Some reinterpretations of the Minnesota Transracial Adoption Study. Intelligence. 19: 21–27

Scarr & Weinberg (1976) IQ test performance of black children adopted by White families. American Psychologist 31(10):726–739 

Weinberg, Scarr & Waldman, (1992) The Minnesota Transracial Adoption Study: A follow-up of IQ test performance at adolescence Intelligence 16:117–135 

Whitney (1996) Shockley’s experiment. Mankind Quarterly 37(1): 41-60

Wilby (2006) Beyond the Flynn effect: New myths about race, family and IQ? Guardian, September 27.

Richard Lynn’s ‘Race Differences in Intelligence’: Useful as a Reference Work, But Biased as a Book

Race Differences in Intelligence: An Evolutionary Analysis, by Richard Lynn (Augusta, GA: Washington Summit, 2006)

Richard Lynn’s ‘Race Differences in Intelligence’ is structured around his massive database of IQ studies conducted among different populations. This collection seems to be largely recycled from his earlier IQ and the Wealth of Nations, and subsequently expanded, revised and reused again in IQ and Global Inequality, The Global Bell Curve, and The Intelligence of Nations (as well as a newer edition of Race Differences in Intelligence, published in 2015). 

Thus, despite its subtitle, “An Evolutionary Analysis”, the focus is very much on documenting the existence of race differences in intelligence, not explaining how or why they evolved. The “Evolutionary Analysis” promised in the subtitle is actually almost entirely confined to the last three chapters. 

The choice of this as a subtitle is therefore misleading and presumably represents an attempt to cash in on the recent rise in, and popularity of, evolutionary psychology and other sociobiological explanations for human behaviours. 

However, whatever the inadequacies of Lynn’s theory of how and why race differences in intelligence evolved (discussed below), his documentation of the existence of these differences is indeed persuasive. The sheer number of studies and the relative consistency over time and place suggests that the differences are indeed real and there is therefore something to be explained in the first place. 

In this respect, it aims to do something similar to what was achieved by Audrey Shuey’s The Testing of Negro Intelligence, first published in 1958, which brought together a huge number of studies, and a huge amount of data, regarding the black-white test score gap in the US. 

However, whereas Shuey focused almost exclusively on the black-white test score gap in North America, Lynn’s ambition is much broader and more ambitious – namely, to review data relating to the intelligences of all racial groups everywhere across the earth. 

Thus, Lynn declares that: 

The objective of this book [is] to broaden the debate from the local problem of the genetic and environmental contributions to the difference between whites and blacks in the United States to the much larger problem of the determinants of the global differences between the ten races whose IQs are summarised” (p182). 

Therefore, his book purports to be: 

The first fully comprehensive review… of the evidence on race differences in intelligence worldwide” (p2). 

Racial Taxonomy

Consistent with this, Lynn includes in his analysis data for many racial groups that rarely receive much if any coverage in previous works on the topic of race differences in intelligence. 

Relying on both morphological criteria and genetic data gathered by Cavalli-Sforz et al in The History and Geography of Human Genes, Lynn identifies ten separate human races. These are: 

1) “Europeans”; 
2) “Africans”; 
3) “Bushmen and Pygmies”; 
4) “South Asians and North Africans”; 
5) “Southeast Asians”; 
6) “Australian Aborigines”; 
7) “Pacific Islanders”; 
8) “East Asians”; 
9) “Artic Peoples”; and 
10) “Native Americans”.

Each of these racial groups receives a chapter of their own, and, in each of the respective chapters, Lynn reviews published (and occasionally unpublished) studies that provide data on each group’s: 

  1. IQs
  2. Reaction times when performing elementary cognitive tasks; and
  3. Brain size

Average IQs 

The average IQs reported by Lynn are, he informs us, corrected for the Flynn Effect – i.e. the rise in IQs over the last century (p5-6).  

However, the Flynn Effect has occurred at different rates in different regions of the world. Likewise, the various environmental factors that have been proposed as possible explanations for the phenomenon (e.g. improved nutrition and health as well as increases in test familiarity, and exposure to visual media) have varied in the extent to which they are present in different places. Correcting for the Flynn Effect is therefore surely easier said than done. 

IQs of “Hybrid populations

Lynn also discusses the average IQs of racially-mixed populations, which are, Lynn reports, consistently intermediate between the average IQs of the two (or more) parent populations.

However, one exception not discussed by Lynn is that recent African immigrants to the US outperform African-Americans both academically and economically, even though, as discussed by African businessman Chanda Chisala, African immigrants tend to be of unadulterated sub-Saharan African ancestry whereas African-Americans are actually a mixed-race population with considerable European ancestry (Chisala 2015a; 2015c; Anderson 2015; see below).

Moreover, both, on the one hand, hybrid vigour or heterosis and, on the other, hybrid incompatibility or outbreeding depression could potentially complicate the assumption that racial hybrids should have average IQs intermediate between the average IQs of the two (or more) parent populations. 

However, Lynn only alludes to the possible effect of hybrid vigour in relation to biracial people in Hawaii, not in relation to other hybrid populations whose IQs he discusses, and never discusses the possible effect of hybrid incompatibility or outbreeding depression at all. 

Genotypic IQs 

Finally, Lynn also purports to estimate what he calls the “genotypic IQ” of at least some of the races discussed. This is a measure of genetic potential, distinguished from their actual realized phenotypic IQ. 

He defines the “genotypic IQ” of a population as the average score of a population if they were raised in environments identical to those of the group with whom they are being compared. 

Thus, he writes: 

The genotypic African IQ… is the IQ that Africans would have if they were raised in the same environment as Europeans” (p69). 

The fact that lower-IQ groups generally provide their offspring with inferior environmental conditions precisely because of their lower intelligence is therefore irrelevant for determining their “genotypic IQ”. However, as Lynn himself later acknowledges: 

It is problematical whether the poor nutrition and health that impair the intelligence of many third world peoples should be regarded as a purely environmental effect or as to some degree a genetic effect arising from the low intelligence of the populations that makes them unable to provide good nutrition and health for their children” (p193). 

Also, Lynn does not explain why he uses Europeans as his comparison group – i.e. why the African genotypic IQ is “the IQ that Africans would have if they were raised in the same environment as Europeans”, as opposed to, say, if they were raised in the same environments East Asians, Middle Eastern populations or indeed their own environments. 

Presumably this reflects historical factors – namely, Europeans were the first racial group to have their IQs systematically measured – the same reason that European IQs are arbitrarily assigned an average score of 100. 

Reaction Times 

Reaction times refer to the time taken to perform so-called elementary cognitive tasks. These are tests where everyone can easily work out the right answer, but where the speed with which different people get there correlates with IQ. 

Arthur Jensen has championed reaction time as a (relatively more) direct measure of one key cognitive process underlying IQ, namely speed of mental processing. 

Yet individuals with quicker reaction times would presumably have an advantage in sports, since reacting to, say, the speed and trajectory of a ball in order to strike or catch it is analogous to an elementary cognitive task. 

However, despite lower IQs, African-Americans, and blacks resident in other western economies, are vastly overrepresented among elite athletes. 
 
To explain this paradox, Lynn distinguishes “reaction time proper” – i.e. when one begins to move one’s hand towards the correct button to press – from “movement time” – how long one’s hand takes to get there. 

Whereas whites generally react faster, Lynn reports that blacks have faster movement times (p58-9).[1] Thus, Lynn concludes: 

The faster movement times of Africans may be a factor in the fast sprinting speed of Africans shown in Olympic records” (p58). 

However, psychologist Richard Nisbett reports that: 

Across a host of studies, movement times are just as highly correlated with IQ as reaction times” (Intelligence and How to Get It: p222). 

Brain Size

Lynn also reviews data regarding the brain-size of different groups. 

The correlation between brain-size and IQ as between individuals is well-established (Rushton and Ankney 2009). 
 
As between species, brain-size is also thought to correlate with intelligence, at least after controlling for body-size. Thus, species whose behaviours suggest high intelligence (e.g. dolphins, chimpanzees, corvids, some parrots) also tend to have large brains as compared to other species of similar body-size.

Indeed, since brain tissue is highly metabolically expensive, increases in brain-size would surely never have evolved without conferring some countervailing selective advantage such as increased intelligence. 

Thus, in the late-1960s, biologist HJ Jerison developed an equation to estimate an animal’s intelligence from its brain- and body-size alone. This is called the animal’s encephalization quotient
 
However, comparing the intelligence of different species poses great difficulties. In short, if you think a culture fair’ intelligence test is an impossibility, then try designing a ‘species fair’ test![2]

Moreover, dwarves have smaller absolute brain-sizes but usually larger brains relative to body-size, but usually have IQs within the normal range.

This is probably because dwarfism is an abnormal and pathological condition – a malfunction in growth and development. Therefore, the increased brain volume relative to body-size associated with disproportionate dwarfism did not evolve through natural selection. Despite its metabolic cost, the additional brain tissue may then indeed confer no adaptive advantage. 

Similarly, some forms of macrocephaly (i.e. abnormally large head and brain size) actually seem to be associated with impaired cognitive ability, probably because the condition reflects a malfunction in brain growth, such that the additional brain tissue may again be without adaptive function.

Sex differences in IQ, meanwhile, are smaller than those between races even though differences in brain-size are greater, at least before one introduces controls for body-size.

Also, Neanderthals had larger brains than modern humans, despite a shorter, albeit more robust, stature.

One theory has it that population differences in brain-size reflect a climatic adaptation that evolved in order to regulate temperature, in accordance with Bermann’s Rule. This seems to be the dominant view among contemporary biological anthropologists, at least those who deign (or dare) to even discuss this politically charged topic.[3] 

Thus, in one recent undergraduate textbook in biological anthropology, authors Mielke, Konigsberg and Relethford contend: 

Larger and relatively broader skulls lose less heat and are adaptive in cold climates; small and relatively narrower skulls lose more heat and are adaptive in hot climates” (Human Biological Variation: p285). 

On this view, head size and shape represents a means of regulating the relative ratio of surface-area-to-volume, since this determines the proportion of a body that is directly exposed to the elements.

Thus, Stephen Molnar, the author of another competing undergraduate textbook in biological anthropology, observes

The closer a structure approaches a spherical shape, the lower will be the surface-to-volume ratio. The reverse is true as elongation occurs—a greater surface area to volume is formed, which results in more surface to dissipate heat generated within a given volume. Since up to 80 percent of our body heat may be lost through our heads on cold days, one can appreciate the significance of shape” (Human Variation: Races, Types and Ethnic Groups, 5th Ed: p188).

This then might explain why, despite the relatively primitive state of their pre-contact civilization and not especially high IQ scores (see below), those whom Lynn terms “Artic Peoples” (i.e. Eskimos) have, according to Lynn’s data, the largest brains of any of the racial groups whom he discusses.[4]

The BermannAllen rules likely also explain at least some of the variation in body-size and stature as between racial groups. 

For example, Eskimos tend to be short and stocky, with short arms and legs and flat faces. This minimizes the ratio of surface-area-to-volume, ensures only a minimal proportion of the body is directly exposed to the elements, and also minimizes the extent of extremities (e.g. arms, legs, noses), which are especially vulnerable to the cold and frostbite. 

In contrast, populations from tropical climates, such as African blacks and Australian Aboriginals, tend to have relatively long arms and legs as compared to trunk size, a factor which likely contributes towards their success in some athletic events. 

Yet, interestingly, Beals et al report that:

Braincase volume is more highly correlated with climate than any of the summative measures of body-size” (Beals et al 1984: p305).

In other words, brain-size is more strongly correlated with the climate in which one’s ancestors evolved than is overall body-size or other bodily dimensions.

Why this is so is not clear. One might perhaps infer it is because head-size and shape is especially important in the regulation of temperature.

In fact, however, contrary to popular wisdom, humans do not lose an especially high proportion of our body heat through our heads, certainly not “up to 80 percent of our body heat”, as claimed in Stephen Molnar’s anthropology textbook as quoted above, a preposterous figure given that the head comprises only about 10% of the body’s overall surface area.

Indeed, the amount of heat lost through our head is relatively higher than that lost through other parts of the body only because other parts of the body are typically covered by clothing.

At any rate, it is surely implausible that an increase in brain tissue, which is metabolically highly expensive, would have evolved solely for the purpose of regulating temperature, when the same result could surely have been achieved by modifying only the external shape of the skull.

Conversely, even if race differences in brain-size did evolve purely for temperature regulation, differences in intelligence could still have emerged as a by-product of such selection.

In other words, if larger brains did evolve among populations inhabiting colder latitudes solely for the purposes of temperature regulation, the extra brain tissue that resulted may still have resulted in greater levels of cognitive ability among these populations, even if there was no direct selection for increased cognitive ability itself.

Europeans

The first racial group discussed by Lynn are those he terms “Europeans” (i.e. white Caucasians). He reviews data on IQ both in Europe and among diaspora populations elsewhere in the world (e.g. North America, Australia). 

The results are consistent, almost always giving an average IQ of about 100 – though this figure is, of course, arbitrary and reflects the fact that IQ tests were first normed by reference to European populations. This is what James Thompson refers to as the ‘Greenwich mean IQ’ and the IQs of all other populations in Lynn’s book are calculated by reference to this figure. 
 
Southeast Europeans, however, score slightly lower. This, Lynn argues, is because: 

Balkan peoples are a hybrid population or cline, comprising a genetic mix between the Europeans and South Asians in Turkey” (p18). 

Therefore, as a hybrid population, their IQs are intermediate between those of the two parent populations, and, according to Lynn, South Asians score somewhat lower in IQ than do white European populations (see below). Similarly, the Turkish people, being intermixed with Europeans, score slightly higher than other Middle-Eastern populations (p80).

In the newer 2015 edition, Lynn argues that IQs are somewhat lower elsewhere in southern Europe, namely southern Spain and Italy, for much the same reason, namely because: 

The populations of these regions are a genetic mix of European people with those from the Near East and North Africa, with the result that their IQs are intermediate between the parent populations” (Preface, 2015 Edition).[5]

An alternative explanation is that these regions (e.g. Balkan countries, Southern Italy) have lower living-standards. 

However, instead of viewing differences in living standards as causing differences in recorded IQs as between populations, Lynn argues that differences in innate ability themselves cause differences in living standards, because, according to Lynn, more intelligent populations are better able to achieve high levels of economic development (see IQ and the Wealth of Nations).[6]

Moreover, Lynn observes that in Eastern Europe, living standards are substantially below elsewhere in Europe as a consequence of the legacy of communism. However, populations from Eastern Europe score only slightly below those from elsewhere in Europe, suggesting that even substantial differences in living-standards may have only a minor impact on IQ (p20). 

Portuguese 

The Portuguese also, Lynn claims, score lower than elsewhere in Europe. 

However, he cites just two studies. These give average IQs of 101 and 88 respectively, which Lynn averages to give an average of 94.5 (p19). 

Yet these two results are actually highly divergent, the former actually being slightly higher than the average for north-west Europe. This suggests an inadequate basis on which to posit a genetic difference in ability. 

However, from this meagre data set, Lynn does not hesitate to provocatively conclude: 

Intelligence in Portugal has been depressed by the admixture of sub-Saharan Africans. Portugal was the only European country to import black slaves from the fifteenth century onwards” (p19). 

This echoes nineteenth century French racialist Arthur De Gobineau’s infamous theory that empires decline because, through their empires, they conquer large numbers of inferior peoples, who then inevitably interbreed with their conquerors, which, according to De Gobineau, results in the dilution the very qualities that permitted their imperial glories in the first place. 

In support of Lynn’s theory, mitochondrial DNA studies have indeed found higher frequency of sub-Saharan African Haplogroup L in Portugal than elsewhere in Europe (e.g. Pereira et al 2005). 

Ireland and ‘Selective Migration

IQs are also, Lynn reports, somewhat lower than elsewhere in Europe in Ireland. 

Lynn cites four studies of Irish IQs which give average scores of 87, 97, 93 and 91 respectively. Again, these are rather divergent but nevertheless consistently below the European average, all but one substantially so. 
 
Of course, in England, in less politically correct times, the supposed stupidity of the Irish was once a staple of popular humour, Irish jokes being the English equivalent of Polish jokes in America.[7]
 
However, the low IQ scores reported for Ireland seem anomalous given the higher average IQs recorded elsewhere in North-West Europe, especially the UK, Ireland’s next-door neighbour, whose populations are closely related to those in Ireland.

Also, in relation to Lynn’s Cold Winters Theory (see below), the climate in Ireland is quite cold.

Moreover, although head size is obviously a crude, indirect measure of brain size, it is perhaps worth observing that Carleton Coon reported in 1939 that Ireland has “the largest heads of any country excepting Belgium”, head-size being especially large in the southwestern half of Ireland (The Races of Europe: p265). Thus, Coon reports that overall:

Ireland consistently has the largest head size of any equal land area in Europe” (The Races of Europe: p377).

Of course, historically Ireland was, until relatively recently, quite poor by European standards. 

It is also sparsely populated and a relatively high proportion of the population live in rural areas, and there is some evidence that people from rural areas have lower average IQs than those from urban areas

However, economic deprivation cannot explain the disparity. Today, despite the 2008 economic crash, and inevitable British bailout, Ireland enjoys, according to the UN, a higher Human Development Index than does the UK, and has done for some time. Indeed, by this measure, Ireland enjoys one of the highest standards of living in the world

Moreover, although formerly Ireland was much poorer, the studies cited by Lynn were published from 1973 to 1993, yet show no obvious increase over time.[8] 
 
Lynn himself attributes the depressed Irish IQ to what he calls ‘selective migration’, claiming: 

There has been some tendency for the more intelligent to migrate, leaving less intelligent behind” (p19). 

Of course, this would suggest, not only that the remaining Irish would have lower average IQs, but also that the descendants of Irish émigrés in Britain, Australia, America and other diaspora communities would have relatively higher IQs than other white people. 

In support of this, Americans reporting Irish ancestry do indeed enjoy higher relative incomes as compared to most other white American ethnicities. 

Interestingly, Lynn also invokes “selective migration” to explain the divergences in East Asian IQs. Here, however, it was supposedly the less intelligent who chose to migrate (p136; p138; p169).[9]

Meanwhile, other hereditarians have sought to explain away the impressive academic performance of recent African immigrants to the West (see below), and their offspring, by reference to selective immigration of high IQ Africans, an explanation which is wholly inadequate on mathematical grounds alone (see Chisala 2015b; 2019).

It certainly seems plausible that migrants differ in personality from those who choose to remain at home. It is likely that they are braver, have greater determination, drive and willpower than those who choose to stay behind. They may also perhaps be less ethnocentric, and more tolerant of foreign cultures.[10]

However, I see no obvious reason they would differ in intelligence.

As African businessman Chanda Chisala writes:

Realizing that life is better in a very rich country than in your poor country is never exactly the most g-loaded epiphany among Africans” (Chisala 2015b).

Likewise, it likely didn’t take much brain-power for Irish people to realize during the Irish Potato Famine that they were less likely to starve to death if they emigrated abroad.

Of course, wealth is correlated with intelligence and may affect the decision to migrate.

The rich usually have little economic incentive to migrate, while the poor may be unable to afford the often-substantial costs of migration (e.g. transportation).

However, without actual historical data showing certain socioeconomic classes or intellectual ability groups were more likely to migrate than others, Lynn’s claims regarding ‘selective migration’ represent little more than a post-hoc rationalization for IQ differences that are otherwise anomalous and not easily explicable in terms of heredity

Ireland, Catholicism and Celibacy

Interestingly, in the 2015 edition of ‘Race Differences in Intelligence’, Lynn also proposes, in addition, a further explanation for the low IQs supposedly found in Ireland, namely the clerical celibacy demanded under Catholicism. Thus, Lynn argues:

There is a dysgenic effect of Roman Catholicism, in which clerical celibacy has reduced the fertility of some of the most intelligent, who have become priests and nuns” (2015 Edition; see also Lynn 2015). 

Of course, this theory presupposes that it was indeed the most intelligent among the Irish people who became priests. However, this is a questionable assumption, especially given the well-established inverse correlation between intelligence and religiosity (Zuckerman et al 2013).

However, it is perhaps arguable that, in an earlier age, when religious dogmas were relentlessly enforced, religious scholarship may have been the only form of intellectual endeavour that it was safe for intellectually-minded people to engage in.

Anyone investigating more substantial matters, such as whether the earth revolved around the sun or vice versa, was liable to be burnt at the stake if he reached the wrong (i.e. the right) conclusion.

However, such an effect would surely also apply in other historically Catholic countries as well.

Yet there is little if any evidence of depressed IQs in, say, France or Austria, although the populaions of both these countries were, until recently, like that of Ireland, predominantly Catholic.[11]

Africans 

The next chapter is titled “Africans”. However, Lynn uses this term to refer specifically to black Africans – i.e. those formerly termed ‘Negroes’. He therefore excludes from this chapter, not only the predominantly ‘Caucasoid’ populations of North Africa, but also African Pygmies and the Khoisan of southern Africa, who are considered separately in a chapter of their own. 

Lynn’s previous estimate of the average sub-Saharan African IQ as just 70 provoked widespread incredulity and much criticism. However, undeterred, Lynn now goes even further, estimating the average African IQ even lower, at just 67.[12]

Curiously, according to Lynn’s data, populations from the Horn of Africa (e.g. Ethiopia and Somalia) have IQs no higher than populations elsewhere in sub-Saharan Africa.[13]

Yet populations from the Horn of Africa are known to be partly, if not predominantly, Caucasoid in ancestry, having substantial genetic affinities with populations from the Middle East.[14].

Therefore, just as populations from Southern Europe have lower average IQs than other Europeans because, according to Lynn, they are genetically intermediate between Europeans and Middle Eastern populations, so populations from the Horn of Africa should score higher than those from elsewhere in sub-Saharan Africa because of intermixture with Middle Eastern populations.

However, Lynn’s data gives average IQs for Ethiopia and Somalia of just 68 and 69 respectively – no higher than elsewhere in sub-Saharan Africa (The Intelligence of Nations: p87; p141-2).

On the other hand, blacks resident in western economies score rather higher, with average IQs around 85. 

The only exception, strangely, are the Beta Israel, who also hail from the Horn of Africa, but are now mostly resident in Israel, yet who score no higher than those blacks still resident in Africa. From this, Lynn concludes:

These results suggest that education in western schools does not benefit the African IQ” (p53). 

However, why then do blacks resident in other western economies score higher? Are blacks in Israel somehow treated differently than those resident in the UK, USA or France? 

For his part, Lynn attributes the higher scores of blacks resident in these other Western economies to both superior economic conditions and, more controversially, to racial admixture. 

Thus, African-Americans in particular are known to be a racially-mixed population, with substantial European ancestry (usually estimated at around 20%) in addition to their African ancestry.[15]

Therefore, Lynn argues that the higher IQs of African-Americans reflect, in part, the effect of the European portion of their ancestry. 

However, this explanation is difficult to square with the observation that, as documented by African businessman Chanda Chisala among others, recent African immigrants to the US, themselves presumably largely of unmixed sub-Saharan African descent, actually consistently outperform African-Americans (and sometimes whites as well!) both academically and  economically (Chisala 2015a2015cAnderson 2015).[16]

Musical Ability” 

Lynn also reviews the evidence pertaining to one class of specific mental ability not covered in most previous reviews on the subject – namely, race differences in musical ability. 

The accomplishments of African-Americans in twentieth century jazz and popular music are, of course, much celebrated. To Lynn, however, this represents a paradox, since musical abilities are known to correlate with general intelligence and African-Americans generally have low IQs. 
 
In addressing this perceived paradox, Lynn reviews the results of various psychometric measures of musical ability. These tests include: 

  • Recognizing a change in pitch; 
  • Remembering a tune; 
  • Identifying the constituent notes in a chord; and 
  • Recognizing whether different songs have similar rhythm (p55). 

In relation to these sorts of tests, Lynn reports that African-Americans actually score somewhat lower in most elements of musical intelligence than do whites, and their musical ability is indeed generally commensurate with their general low IQs. 

The only exception is for rhythmical ability. 

This is, of course, congruent with the familiar observation that black musical styles place great emphasis on rhythm

However, even with respect to rhythmical ability, blacks score no higher than whites. Instead, blacks’ scores on measures of rhythmical ability are exceptional only in that this is the only form of musical ability on which blacks score equal to, but no higher than, whites (p56). 

For Lynn, the low scores of African-Americans in psychometric tests of musical ability are, on further reflection, little surprise. 

The low musical abilities of Africans… are consistent with their generally poor achievements in classical music. There are no African composers, conductors, or instrumentalists of the first rank and it is rare to see African players in the leading symphony orchestras” (p57). 

However, who qualifies as a composer, conductor or instrumentalist “of the first rank” is, ultimately, unlike the results of psychometric testing, a subjective assessment, as are all artistic judgements. 

Moreover, why is achievement in classical music, an obviously distinctly western genre of music, to be taken as the sole measure of musical accomplishment? 

Even if we concede that the ability required to compose and perform classical music is greater than that required for other genres (e.g. jazz and popular music), musical intelligence surely facilitates composition and performance in other genres too – and, given the financial rewards offered by popular music often dwarf those enjoyed by players and composers of classical music, the more musically-gifted race would have every incentive to dominate this field too. 

Perhaps, then, these psychometric measures fail to capture some key element of musical ability relevant to musical accomplishment, especially in genres other than classical. 

In this context, it is notable that no lesser champion of standardized testing than Arthur Jensen has himself acknowledged that intelligence tests are incapable of measuring creativity (Langan & LoSasso 2002: p24-5). 

In particular, one feature common to many African-American musical styles, from rap freestyling to jazz, is improvisation.  

Thus, Dinesh D’Souza speculates tentatively that: 

Blacks have certain inherited abilities, such as improvisational decision making, that could explain why they predominate in… jazz, rap and basketball” (The End of Racism: p440-1). 

Steve Sailer rather less tentatively expands upon this theme, positing an African advantage in: 

Creative improvisation and on-the-fly interpersonal decision-making” (Sailer 1996). 

On this basis, Sailer concludes that: 

Beyond basketball, these black cerebral superiorities in ‘real time’ responsiveness also contribute to black dominance in jazz, running with the football, rap, dance, trash talking, preaching, and oratory” (Sailer 1996). 

Bushmen and Pygmies” 

Grouped together as the subjects of the next chapter are black Africans’ sub-Saharan African neighbours, namely San Bushmen and Pygmies

Quite why these two populations are grouped together by Lynn in a single chapter is unclear. 

He cites Cavalli-Sforza et al in The History and Geography of Human Genes as providing evidence that: 

These two peoples have distinctive but closely related genetic characteristics and form two related clusters” (p73). 

However, although both groups are obviously indigenous to sub-Saharan Africa and quite morphologically distinct from the other black African populations who today represent the great majority of the population of sub-Saharan Africa, they share no especial morphological similarity to one another.[17]

Moreover, since Lynn acknowledges that they have “distinctive… genetic characteristics and form two… clusters”, they presumably should each of merited chapters of their own.[18]

One therefore suspects that they are lumped together more for convenience than on legitimate taxonomic grounds. 

In short, both are marginal groups of hunter-gatherers, now few in number, few if any of whom have been exposed to the sort of standardized testing necessary to provide a useful estimate of their average IQs. Therefore, since his data on neither group alone is really sufficient to justify its own chapter, he groups them together in a single chapter.  

However, the lack of data on IQ for either group means that even this combined chapter remains one of the shorter chapters in Lynn’s book, and, as we will see, the paucity of reliable data on the cognitive ability of either group leads one to suspect that Lynn might have been better omitting both groups from his survey of race differences in cognitive ability altogether, just as he omitted at least one other phenotypically quite distinct racial group for whom presumably there is again little data on IQs, namely the Negrito populations of South and South-East Asia. 

San Bushmen

It may be some meagre consolation to African blacks that, at least in Lynn’s telling, they no longer qualify as the lowest scoring racial group when it comes to IQ. Instead, this dubious honour is now accorded their sub-Saharan African neighbours, San Bushmen
 
In Race: The Reality of Human Differences (which I have reviewed here and here), authors Vincent Sarich and Frank Miele quote anthropologist and geneticist Henry Harpending as observing: 

All of us have the impression that Bushmen are really quick and clever and are quite different from their [black Bantu] neighbors… Bushmen don’t look like their black African neighbors either. I expect that there will soon be real data from the Namibian school system about the relative performance of Bushmen… and Bantu kids – or more likely, they will suppress it” (Race: The Reality of Human Differences (reviewed here): p227). 

Today, however, some fifteen or so years after Sarich and Miele published this quotation, the only such data I am aware of is that reported by Lynn in this book, which suggests, at least according to Lynn, a level of intelligence even lower than that of other sub-Saharan Africans. 

Unfortunately, however, the data in question is very limited and, in my view, inadequate to support Lynn’s controversial conclusions regarding Bushmen ability.  

It also consists of just three studies, none of which remotely resemble a full IQ test (p74-5). 

Yet, from this meagre dataset, Lynn does not hesitate to attribute to Bushmen an average IQ of just 52. 

If Lynn’s estimate of the average sub-Saharan African IQ at around 70 provoked widespread incredulity, then his much lower estimate for Bushmen is unlikely to fare better. 

Lynn anticipates such a reaction, and responds by pointing out:  

An IQ of 54 represents the mental age of the average European 8-year-old, and the average European 8-year-old can read, write, and do arithmetic and would have no difficulty in learning and performing the activities of gathering foods and hunting carried out by the San Bushmen. An average 8-year-old can easily be taught to pick berries put them in a container and carry them home, collect ostrich eggs and use the shells for storing water and learn how to use a bow and arrow” (p76). 

Indeed, Lynn continues, other non-human animals survive in difficult, challenging environments with even lower levels of intelligence:  

Apes with mental abilities about the same as those of human 4-year olds survive quite well as gatherers and occasional hunters and so also did early hominids with IQs around 40 and brain sizes much smaller than those of modern Bushmen. For these reasons there is nothing puzzling about contemporary Bushmen with average IQs of about 54” (p77). 

Here, Lynn makes an important point. Many non-human animals survive and prosper in ecologically challenging environments with levels of intelligence much lower than that of any hominid, let alone any extant human race. 

On the other hand, however, I suspect Lynn would not last long in Kalahari Desert – the home environment of most contemporary Bushmen.

Pygmies 

Lynn’s data on the IQs of Pygmies is even more inadequate than his data for Bushmen. Indeed, it amounts to just one study, which again fell far short of a full IQ test. 

Moreover, the author of the study, Lynn reports, did not quantify his results, reporting only that Pygmies scored much “much worse” than other populations tested using the same test (p78). 

However, while the other populations tested using the same test and outperforming Pygmies included “Eskimos, Native American and Filipinos”, Lynn conspicuously does not mention that they included other black Africans, or indeed other very low IQ groups such as Australian Aboriginals (p78). 

Thus, Lynn’s assumption that Pygmies are lower in cognitive ability than other black Africans is not supported even by the single study that he cites. 

Lynn also infers a low level of intelligence for Pygmies from their lifestyle and mode of sustenance: 

Most of them still retain a primitive hunter-gatherer existence while many of the Negroid Africans became farmers over the last few hundred years” (p78). 

Thus, Lynn assumes that whether a population has successfully transitioned to agriculture is largely a product of their intelligence (p191). 

In contrast, most historians and anthropologists would emphasize the importance of environmental factors in explaining whether a group transitions to agriculture.[19]

Finally, Lynn also infers a low IQ from the widespread enslavement of Pygmies by neighbouring Bantus: 

The enslavement of Pygmies by Negroid Africans is consistent with the general principle that the more intelligent races generally defeat and enslave the less intelligent, just as Europeans and South Asians have frequently enslaved Africans but not vice versa” (p78). 

However, while it may be a “general principle that the more intelligent races typically defeat and enslave the less intelligent”, if only because, being, on average, superior in military technology, the former are better able to conquer the latter than vice versa, this is hardly a rigid rule. 

After all, Middle Eastern and North African Muslims sometimes enslaved Europeans.[20] Yet, according to Lynn, the Arabs belong to a rather less intelligent race than do the Europeans whom they so often enslaved

Interestingly, it is notable that Pygmies are the only racial group whom Lynn includes in his survey for whom he does not provide an actual figure as an estimate their average IQ, which presumably reflects a tacit admission of the inadequacy of the available data.[21] 

Curiously, unlike for all the other racial groups discussed, Lynn also fails to provide any data on Pygmy brain-size. 

Presumably, Pygmies have small brains as compared to other races, if only on account of their smaller body-size – but what about their brain-size relative to body-size? Is there simply no data available?

Australian Aborigines 

Another group who are barely mentioned at all in most previous discussions of the topic of race differences in intelligence are Australian Aborigines. Here, however, unlike for Bushmen and Pygmies, data from Australian schools are actually surprisingly abundant. 

These give, Lynn reports, an average Aboriginal IQ of just 62 (p104). 

Unlike his estimates for Bushmen and Pygmies, this figure seems to be reliable, given the number of studies cited and the consistency of their results. One might say, then, that Australian Aboriginals have the lowest recorded IQs of any human race for whom reliable data is available. 

Interestingly, in addition to his data on IQ, Lynn also reports the results of Piagetian measures of development conducted among Aboriginals. He reports, rather remarkably, that a large minority of Aboriginal adults fail to reach what Piaget called the concrete operational stage of development – or, more specifically, fail to recognize a substance, transferred to a new container, necessarily remains of the same quantity (p105-7). 

Perhaps even more remarkable, however, are reports of Aborigine spatial memory (p107-8). This refers to the ability to remember the location of objects, and their locations relative to one another. 

Thus, he reports, one study found that, despite their low general cognitive ability, Aborigines nevertheless score much higher than Europeans in tests of spatial memory (Kearins 1981).  

Another study found no difference in the performance of whites and Aborigines (Drinkwater 1975). However, since Aborigines have much lower IQs overall, even equal performance on spatial memory as against Europeans is still out of sync with the performance of whites and Aborigines on other types of intelligence test (p108). 

Lynn speculates that Aboriginal spatial memory may represent an adaptation to facilitate navigation in a desert environment with few available landmarks.[22]

The difference, Lynn argues, seems to be innate, since it was found even among Aborigines who had been living in an urban environment (i.e. not a desert) for several generations (p108; but see Kearins 1986). 

Two other studies reported lower scores than for Europeans. However, one was an unpublished dissertation and hence must be treated with caution, while the and the other (Knapp & Seagrim 1981) “did not present his data in such a way that the magnitude of the white advantage can be calculated” (p108). 

Intriguingly, Lynn reports that this ability even appears to be reflected in neuroanatomy. Thus, despite smaller brains overall, Aborigines’ right visual cortex, implicated in spatial ability, is relatively larger than in Europeans (Klekamp et al 1987; p108-9).

New Guineans and Jared Diamond 

In his celebrated Guns, Germs and Steel, Jared Diamond famously claimed: 

In mental ability New Guineans are probably genetically superior to Westerners, and they surely are superior in escaping the devastating developmental disadvantages under which most children in industrialized societies grow up” (Guns, Germs and Steel: p21). 

Diamond bases this claim on the fact that, in the West, survival, throughout most of our recent history, depended on who was struck down by disease, which was largely random. 

In contrast, in New Guinea, he argues, people had to survive on their wits, with survival depending on one’s ability to procure food and avoid homicide, activities in which intelligence was likely to be at a premium (Guns, Germs and Steel: p20-21). 

He also argues that the intelligence of western children is likely reduced because they spend too much time watching television and movies (Guns, Germs and Steel: p21). 

However, there is no evidence television has a negative impact on children’s cognitive development. Indeed, given the rise in IQs over the twentieth century has been concomitant with increases in television viewing, it has even been speculated that increasingly stimulating visual media may have contributed to rising IQs. 

On the basis of two IQ studies, plus three studies of Piagetian development, Lynn concludes that the average IQ of indigenous New Guineans is just 62 (p112-3). 

This is, of course, exactly the same as his estimate for the average IQ of Australian Aboriginals.

It is therefore consistent with Lynn’s racial taxonomy, since, citing Cavalli-Sforza et al, he classes New Guineans as in the same genetic cluster, and hence as part of the same race as Australian Aboriginals (p101). 

Pacific Islanders 

Other Pacific Islanders, however, including Polynesians, Micronesians, Melanesians and Hawiians, are grouped separately and hence receive a chapter of their own. 

They also, Lynn reports, score rather higher in IQ, with most such populations having average IQs of about 85 (p117). However, the Māoris of New Zealand score rather higher, with an average IQ of about 90 (p116). 

Hawaiians and Hybrid Vigor 

For the descendants of the inhabitants of one particular Pacific Island, namely Hawaii, Lynn also reports data regarding the IQs of racially-mixed individuals, both those of part-Native-Hawiian and part-East Asian ancestry, and those of part-Native-Hawiian and part-European ancestry. 

These racial hybrids, as expected, score on average between the average scores for the two parent populations. However, Lynn reports: 

The IQs of the two hybrid groups are slightly higher than the average of the two parent races. The average IQ of the Europeans and Hawaiians is 90.5, while the IQ of the children is 93. Similarly, the average IQ of the Chinese and Hawaiians is 90, while the IQ of the children is 91. The slightly higher than expected IQs of the children of the mixed race parents may be a hybrid vigor or heterosis effect” (p118). 

Actually, the difference between the “expected IQs” and the IQs actually recorded for the hybrid groups is so small (only one point for the Chinese-Hawaiians), that it could easily be dismissed as mere noise, and I doubt it would reach statistical significance. 

Nevertheless, Lynn’s discussion begs the question as to why hybrid vigor has not similarly elevated the IQs of the other hybrid, or racially-mixed, populations discussed in other chapters, and why Lynn has not discussed this issue when reporting the average IQs of other racially-mixed populations in other chapters. 

Of course, while hybrid vigor is a real phenomenon, so is outbreeding depression and hybrid incompatibilities

Presumably then, which of these countervailing effects outweighs the other for different types of hybrid depends on the degree of genetic distance between the two parent populations. This, of course, varies for different races. 

It is therefore possible that some racial mixes may tend to elevate intelligence, whereas others, especially between more distantly-related populations, may tend, on average, to depress intelligence. 

For what it’s worth, Pacific Islanders, including Hawiians, are thought to be genetically closer to East Asians than to Europeans. 

South Asians and North Africans

Another group rarely treated separately in earlier works are those whom Lynn terms “South Asians and North Africans”, though this group also includes populations from the Middle East. 

Physical anthropologists often lumped these peoples together with Europeans as collectively “Caucasian” or “Caucasoid”. However, while acknowledging that they are “closely related to the Europeans”, Lynn cites Cavalli-Sforza et al as showing they form “a distinctive genetic cluster” (p79). 

He also reports that they score substantially lower in IQ than do Europeans. Their average IQ in their native homelands is just 84 (p80), while South Asians resident in the UK score only slightly higher with an average IQ of just 89 (p82-4). 

This conclusion is surely surprising and should, in my opinion, be treated with caution. 

For one thing, all of the earliest known human civilizations – namely,Mesopotamia, Egypt and the Indus Valley civilization – surely emerged among these peoples, or at least in regions today inhabited primarily by people of this race.[23]

Moreover, people of Indian ancestry in particular are today regarded as a model majority in both Britain and America, whose overrepresentation in the professions, especially medicine, is widely commented upon.[24]

Indeed, according to some measures, British-Indians are now the highest earning ethnicity in Britain, or the second-highest earning after the Chinese, and Indians are also the highest earners in the USA.[25]

Interestingly, in this light, one study cited by Lynn showed a massive gain of 14-points for children from India who had been resident in the UK for more than four years as compared to those who had been resident for less than four years, the former scoring almost as high in IQ as the indigenous British, with an average IQ of 97 (p83-4; Mackintosh & Mascie-Taylor 1985).[26]

In the light of this study, it would be interesting to measure the IQs of a sample composed exclusively of people who traced their ancestry to India but who had been resident in the UK for the entirety of their lives (or even whose ancestors had been resident in the UK for successive generations), since all of the other studies cited by Lynn of the IQs of Indian children in the UK presumably include both recent arrivals and long-term residents grouped together, and many British-Indians have now been resident in the UK for multiple generations.

In fact, however, the co-author of this paper, Nicholas Mackintosh, claims, in his own review of Lynn’s book, that the results of his study are misreported by Lynn. In fact, he asserts, the study in question (which I personally have not read) reported an average IQ of 97 for ten-year old children of Indian ancestry resident in Britain but only of 93 for children of Pakistani background (Mackintosh 2007).

In the same book review, Mackintosh also claims that another paper which he co-authored and which is cited by Lynn regarding the IQs of South Asian children resident in the UK is also misreported, and again in fact recorded significantly higher IQs for children of Indian ancestry than for those of Pakistani origin, the former averaging 91 and the latter only 85 (West et al 1992).

Thus, Mackintosh reports:

In fact, three British studies have given the same IQ tests to Indian and Pakistani children, and in all three, Indian children have outscored the Pakistanis by 4–6 IQ points” (Mackintosh 2007: 94).

In this light, it is interesting to observe that there is also a large difference in socio-economic status and average earnings as between, on the one hand, British-Indians, and, on the other, both British-Pakistanis and Bangladeshis . Indeed, the same data suggesting that British-Indians are the highest earning ethnicity in Britain also show that British-Pakistanis and Bangladeshis are among the lowest earners in the UK

The primary divide between these three countries is, of course, not so much racial as religious. This suggests a religion as a causal factor in the difference.[27]

Indeed, data on average earnings by religion rather than national origin show a similar pattern with Hindus having the highest average salaries of any religious community in the UK excepting Jews, and Muslims having the lowest.

A similar pattern is apparent in the USA, where Hindus, again, come second to Jews, with Muslims among the lower earning groups.

Similarly, one study found that predominantly Muslim countries tend to have lower average IQs than do predominantly non-Muslim countries (Templer 2010; see also Dutton 2020). 

Perhaps, then, cultural practices in Muslim countries are responsible for reducing IQs. 

For example, the prevalence of consanguineous (i.e. incestuous) marriage, especially cross-cousin marriage, in many Islamic cultures may have an adverse impact on intelligence levels due to the effects of inbreeding depression (Woodley 2009). 

Another cultural practice that could affect intelligence in Muslim countries is the practice of even pregnant women, though supposedly exempt from requirement to fast during daylieght hours during Ramadan, nevertheless still choosing to do so (cf. Aziz et al 2004). 

However, Lynn’s own data show little difference between IQs in India and those in Pakistan and Bangladesh, nor indeed between IQs in India and those in Muslim countries in the Middle East or North Africa. Nor, according to Lynn’s data, do people of Indian ancestry resident in the UK score noticeably higher in IQ than do people who trace their ancestry to Bagladeshi and Pakistani (cf. Mackintosh 2007). 

An alternative suggestion is that Middle-Eastern and North African IQs have been depressed as a result of interbreeding with sub-Saharan Africans, perhaps as a result of the Islamic slave trade.[28]

This is possible because, although male slaves in the Islamic world were routinely castrated and hence incapable of procreation, female slaves outnumbered males and were often employed as concubines, a practice which, unlike in puritanical North America, was regarded as perfectly socially acceptable on the part of slave owners

This would be consistent with the finding that Arab populations from the Middle East show some evidence of sub-Saharan African ancestry in their mitochondrial DNA, which is passed down the female line, but not in their Y-chromosome ancestry, passed down the male line (Richards et al 2003). 

In contrast, in the United States, the use of female slaves for sexual purposes, although it certainly occurred, was, in the prevailing puritanical Christian morality of the American South, in theory very much frowned upon.

In addition, in North America, due to the one-drop rule, all mixed-race descendants of slaves with any detectable degree of black African ancestry were classed as black. Therefore, at least in theory, the white bloodline would have remained ‘pure’, though some mixed-race individuals may have been able to pass

Therefore, sub-Saharan African genes may have entered the Middle Eastern, and North African, gene-pools in a way they were not able to do so among whites in North America. 

This might explain why genotypic intelligence among North African and Middle Eastern populations may have declined in the period since the great civilizations of Mesopotamia and ancient Egypt and even since the Golden Age of Islam, when the intellectual achievements of Middle Eastern and North African peoples seemed so much more impressive.

This would again be redolent of Arthur De Gobineau’s infamous theory that empires decline because, through their empires, they conquer large numbers of inferior peoples, who then inevitably interbreed with their conquerors, which, according to De Gobineau, diluted the very qualities that permitted their imperial glories in the first place.

However, it is difficult to see how this could have had a significant effect on the genetics, or the IQs, of Muslim people from as far away as South Asia, where any sub-Saharan African genetic input must have been minimal.

Jews

Besides Indians, another economically and intellectually overachieving model minority who derive, at least in part, from the race whom Lynn classes as “South Asians and North Africans” are Jews

Lynn has recently written a whole book on the topic of Jewish intelligence and achievement, titled The Chosen People: A Study of Jewish Intelligence and Achievement (review forthcoming). 

However, in ‘Race Differences in Intelligence’, Jews do not even warrant a chapter of their own. Instead, they are discussed only at the end of the chapter on “South Asians and North Africans”, although Ashkenazi Jews, and Sephardi Jews (but not Mizrahi Jews), also have substantial European ancestry. 

The decision not to devote an entire chapter to the Jewish people is surely correct, because, although even widely disparate groups (e.g. AshkenazimSephardic and Mizrahim, even the Lemba) do indeed share genetic affinities, Jews are not racially distinct (i.e. reliably physically distinguishable on phenotypic criteria) from other peoples. 

However, the decision to include them in the chapter on “South Asians and North Africans” is potentially controversial, since, as Lynn readily acknowledges, Ashkenazi Jews, who today constitute the majority of Jews, have substantial European as well as Middle Eastern ancestry, as indeed do Sephardi Jews (but not Mizrahi Jews). 

Lynn claims British and US Jews have average IQs of around 108 (p68). His data for Israel are not broken down by ethnicity, but give an average IQ for Israel as a whole of 95, which Lynn, rather conjecturally, infers scores of 103 for Ashkenazi Jews, 91 for Mizrahi Jews and 86 for Palestinian-Arabs (p94). 

Lynn’s explanations for Ashkenazi intelligence, however, are wholly unpersuasive. 

First, he observes that, despite Biblical and Talmudic admonitions against miscegenation with Gentiles, Jews inevitably interbred to some extent with the host populations alongside whom they lived. From this, Lynn infers that: 

Ashkenazim Jews in Europe will have absorbed a significant proportion of the genes for higher intelligence possessed by… Europeans” (p95). 

It is indeed true that, if, as Lynn claims, Europeans are indeed a more intelligent race than are populations from the Middle East, then interbreeding with Europeans may indeed explain how Ashkenazim came to score higher in IQ than do other populations tracing their ancestry to the Middle East. 

However, interbreeding with Europeans can hardly explain how Ashkenazi Jews came to outscore, and outperform academically and economically, even the very Europeans with whom they are said to have interbred! 

This explanation therefore fails to explain why Ashkenazim have higher IQs than do Europeans. 

Lynn’s second explanation for high Ashkenazi Jewish IQs is equally unpersuasive. He suggests that: 

The second factor that has probably operated to increase the intelligence of Ashkenazim Jews in Europe and the United States as compared with Oriental Jews is that the Ashkenazim Jews have been more subject to persecution… Oriental Jews experienced some persecution sufficient to raise their IQ of 91, as compared with 84 among other South Asians and North Africans, but not so much as that experienced by Ashkenazim Jews in Europe.” (p95).[29]

On purely theoretical grounds, the idea that persecution selects for intelligence may seem plausible, if hardly compelling.

For example, one might speculate that only the relatively smarter Jews were able to anticipate looming pogroms and hence escape – or, alternatively, since wealth is correlated with intelligence, perhaps only the relatively richer, and hence generally smarter, Jews could afford the costs of migration, including bribes to officials, in order to escape such looming pogroms.[30] 

These are, however, obviously speculative, post-hoc ‘just-so stories’ (in the negative Gouldian sense), and, in the absence of hard data, I put little stock in them. 

There is in fact no evidence that persecution generally acts to increase a group’s intelligence. On the contrary, other groups who have been subject to persecution throughout much of their histories – e.g. the Roma (i.e. Gypsies) and African-Americans – are generally found to have relatively low IQs. 

East and South-East Asians

Excepting Jews, the highest average IQs are found among East Asians, who have, according to Lynn’s data, an average IQ of 105, somewhat higher than that of Europeans (p121-48). 

However, whereas Jews score relatively higher in verbal intelligence than spatio-visual ability, East Asians show the opposite pattern, with relatively higher scores for spatio-visual ability.[31]

However, it is important to emphasize that this relatively high figure applies only to East Asians – i.e. Chinese, Japanese Koreans, Taiwanese etc. – though it has been suggested that the results for China may reflect the oversampling of western diaspora populations and populations from technologically and economically advanced urban areas of China, as opposed to relatively more backward rural regions where IQs seem to be much lower.

Moreover, these high average IQ scores do not apply to the related populations of Southeast Asia (i.e. Thais, Filipinos, Vietnamese, Malaysians, Cambodians, Indonesians etc.), who actually score much lower in IQ, with average scores of only around 87 in their indigenous homelands, but rising to 93 among those resident in the US. 

Thus, Lynn distinguishes the East Asians from Southeast Asians as a separate race, on the grounds that the latter, despite “some genetic affinity with East Asians” form a distinct genetic cluster in data gathered and analyzed by Cavalli-Sforza et al, and also have distinct morphological features, with “the flattened nose and epicanthic eye-fold… [being] less prominent” than among East Asians (p97). 

This is an important point, since many previous writers on the topic have implied that the higher average IQs of East Asians applied to all ‘Asians’ or ‘Mongoloids’, which would presumably include South-East Asians.[32]

Yet, in Lynn’s opinion, it is just as misleading to group all these groups together as ‘Mongoloid’ or ‘Asian’ as it was to group “Europeans” and “South Asians and North Africans” together as ‘Caucasian’ or ‘Caucasoid’. 

However, whether low scores throughout South-East Asia are entirely genetic in origin is unclear. Thus, Vietnamese resident in the West have sometimes, but not always, scored considerably higher, and Jason Malloy suggests that Lynn exaggerates the overrepresentation of ethnic Chinese among Vietnamese immigrants to the West so as attribute such results to East Asians rather than South-East Asians (Malloy 2014).[33]

Moreover, in relation to Lynn’s Cold Winters Theory (discussed below), whereby it is claimed that populations were exposed to colder temperatures during their evolution evolved higher levels of intelligence in order to cope with the adaptive challenges that surviving cold temperatures posed, it is notable that climate varies greatly across China, reflecting the geographic size of the country, with Southern China having a subtropical climate with mild winters.

However, perhaps East Asians, like the Han Chinese, are to be regarded as only relatively recent arrivals in what is now Southern China. This would be consistent with claim of some physical anthropologists that the some aspects of the morphology of East Asians reflects adaptation to the extreme cold of Siberia and the Steppe, and also with the historical expansion of the Han Chinese.

Even more problematic for Cold Winters Theory is the fact that, although Lynn classifies them as East Asian (p121), the higher average IQ scores of East Asians (as compared to whites), does not even extend to the people after whom the Mongoloid race was named – namely the Mongols themselves.

According to Lynn, Mongolians score only around the same as whites, with an average IQ of only 101 (Lynn 2007).

This report is based on just two studies. Moreover, it had not been published at the time the first edition of ‘Race Differences in Intelligence’ came off the presses.

However, Lynn infers a lower IQ for Mongolians from their lower level of cultural, technological and economic development (p240).

Yet, inhabiting the Mongolian-Manchurian grassland Steppe and Gobi Desert, Mongolians were surely subjected to an environment even colder and more austere than that of other East Asians.

On the one hand, this might explain their lower levels of cultural, technological and economic development. On the other, according to Lynn’s Cold Winters Theory, it ought presumably to have resulted in their evolving, if anything, even higher levels of intelligence than other East Asians.

Lynn’s explanation for this anomaly is that the low population-size of the Mongols, and their isolation from other populations, meant that the necessary mutations for higher IQ never arose (p240).[34]

This is the same explanation that Lynn provides for the related anomaly of why Eskimos (“Arctic Peoples”), to whom Mongolians share some genetic affinity, also score low in IQ, an explanation that is discussed in the final part of this review.

Native Americans

Another group sometimes subsumed with Asian populations as “Mongoloids” are the indigenous populations of the American continent, namely “Native Americans”. 

However, on the basis of both genetic data from Cavalli-Sforza et al and morphological differences (“darker and sometimes reddish skin, hooked or straight nose, and lack of the complete East Asian epicanthic fold”), Lynn classifies them as a separate race and hence accords them a chapter of their own. 

His data suggest average IQs of about 86, for both Native Americans resident in Latin America, and also for those resident in North America, despite the substantially higher living standards of the latter (p158; 162-3; p166). 

Mestizo populations, however, have somewhat higher scores, with average IQs intermediate between those of the parent populations (p160).[35]

Like the Asian populations with whom they share their ancestry, Native Americans score rather higher on spatio-visual intelligence than on verbal intelligence (p156). 

In particular, they also have especially high visual memory (p159-60). 

As he did for African-Americans, Lynn also discusses the musical abilities of Native Americans. Interestingly, psychometrical testing shows that their musical ability is rather higher than their general cognitive ability, giving a MQ (Musical Quotient) of approximately 92 (p160). 

They also show the same pattern of musical abilities as do African-Americans, with higher scores for rhythmical ability than for other forms of musical ability (p160). 

However, whereas blacks, as we have seen, only score as high as Europeans for rhythmical ability, but no higher, Native Americans, because of higher IQs (and MQs) overall, actually outscore both Europeans and African-Americans when it comes to rhythmical ability. 

These results are curious. Unlike African-Americans, Native Americans are not, to my knowledge, known for their contribution to any genres of western music, and neither are their indigenous musical traditions especially celebrated. 

Artic Peoples” (i.e. Eskimos)

Distinguished from other Native Americans are the inhabitants of the far north of the American landmass. These, together with other indigenous populations from the area around the Bering straight, namely those from Greenland, the Aleutian Islands, and the far north-east of Siberia, together form the racial group whom Lynn refers to as “Arctic Peoples”, though the more familiar, if less politically correct, term would be ‘Eskimos’.[36]

As well as forming a distinctive genetic cluster per Cavalli-Sforza et al, they are also morphologically distinct, not least in their extreme adaptation to the cold, with, Lynn reports: 

Shorter legs and arms and a thick trunk to conserve heat, a more pronounced epicanthic eye-fold, and a nose well flattened into the face to reduce the risk of frostbite” (p149). 

As we will see, Lynn is a champion of what is sometimes called Cold Winters Theory – namely the theory that the greater environmental challenges, and hence cognitive demands, associated with living in colder climates selected for increased intelligence among those races inhabiting higher latitudes. 

Therefore, on the basis of this theory, one might imagine that Eskimos, who surely evolved in one of the most difficult, and certainly in the coldest, environment of any human group, would also have the highest IQs. 

This conclusion would also be supported by the observation that, according to the data cited by Lynn himself, Eskimos also have the largest average brain-size of any race (p153). 

Interestingly, some early reports did indeed suggest that Eskimos had high levels of cognitive ability as compared to whites.[37] However, Lynn now reports that Eskimos actually have rather lower IQ scores than do whites and East Asians, with results from 15 different studies giving an average IQ of around 90. 

Actually, however, viewed in global perspective, this average IQ of 90 for Eskimos is not that low. Indeed, of the ten major races surveyed by Lynn, only Europeans and East Asians score higher.[38]

It is an especially high score for a population who, until recently, lived exclusively as hunter-gatherers. Other foraging groups, or descendants of peoples who, until recently, subsisted as foragers, tend, according to Lynn’s data, to have low IQs (e.g. Australian Aboriginals, San Bushmen, Pygmies). 

One obvious explanation for the relatively low IQs of Eskimos as compared to Europeans and East Asians would be their deprived living conditions

However, Lynn is skeptical of the claim that environmental factors are entirely to blame for the difference in IQ between Eskimos and whites, since he observes: 

The IQ of the Arctic Peoples has not shown any increase relative to that of Europeans since the early 1930s, although their environment has improved in so far as in the second half of the twentieth century they received improved welfare payments and education. If the intelligence of the Arctic Peoples had been impaired by adverse environmental conditions in the 1930s it should have increased by the early 1980s” (p153-4). 

He also notes that all the children tested in the studies he cites were enrolled in schools (since this was where the testing took place), and hence were presumably reasonably familiar with the procedure of test-taking (p154).

Lynn’s explanation for the relatively low scores of Eskimos is discussed below in the final part of this review.

Visual Memory, Spatial Memory and Hunter-Gathering 

Eskimos also score especially high on tests of visual memory, something not usually measured in standard IQ tests (p152-3). 

This is a proficiency they share in common with Native Americans (p159-60), to whom they are obviously closely related. 

However, as we have seen, Australian Aboriginals, who are not closely related to either group, also seem to possess a similar ability, though Lynn refers to this as “spatial Memory” rather than “visual Memory” (p107-8). 

These are, strictly speaking, somewhat different abilities, although they may not be entirely separate either, and may also be difficult to distinguish between in tests. 

If Aboriginals score high on spatial memory, they may then also score high on visual memory, and vice versa for Eskimos and Native Americans. However, since Lynn does not provide comparative data on visual memory among Aboriginals, or on spatial memory among Eskimos or Native Americans, this is not certain. 

Interestingly, one thing all these three groups share in common is a recent history of subsisting, at least in part, as hunter-gatherers.[39]

One is tempted, then, to attribute this ability to the demands of a hunter-gatherer lifestyle, perhaps reflecting the need to remember the location of plant foods which appear only seasonally, or to find one’s way home after a long hunting expedition.[40] 

It would therefore be interesting to test the visual and spatial memories of other groups who either continue to subsist as hunter-gatherers or only recently transitioned to agriculture or urban life, such as Pygmies and San Bushmen. However, since tests of spatial and visual memory are not included in most IQ tests, the data is probably not yet available.  

For his part, Lynn attributes Eskimo visual memory to the need to “find their way home after going out on long hunting expeditions” (p152-3). 

Thus, just as the desert environment of Australian Aboriginals provides few landmarks, so: 

The landscape of the frozen tundra [of the Eskimos] provides few distinctive cues, so hunters would need to note and remember such few features as do exist” (p153). 

Proximate Causes: Heredity or Environment?

Chapter fourteen discusses the proximate causes of race differences in intelligence and the extent to which the differences observed can be attributed to either heredity or environmental factors, and, if partly the latter, which environmental factors are most important.  

Lynn declares at the beginning of the chapter that the objective of his book is “to broaden the debate” from an exclusive focus on the black-white test score gap in the US, to instead looking at IQ differences among all ten racial groups across the world for whom data on IQ or intelligence is presented in Lynn’s book (p182). 

Actually, however, in this chapter alone, Lynn does indeed focus primarily on black-white differences, if only because it is in relation to this difference that most research has been conducted, and hence to this difference that most available evidence relates. 

Downplaying the effect of schooling, Lynn identifies malnutrition as the major environmental influence on IQ (p182-7). 

However, he rejects malnutrition as an explanation for the low scores of American blacks, noting there is no evidence of short stature in black Americans and nor have surveys have found a greater prevalence of malnutrition (p185). 

As to global differences, he concludes that: 

The effect of malnourishment on Africans in sub-Saharan Africa and the Caribbean probably explains about half of the low IQs, leaving the remaining half to genetic factors” (p185). 

However, it is unclear what is meant by “half of the low scores” as he has identified no comparison group.[41] 

He also argues that the study of racially mixed individuals further suggests a genetic component to observed IQ differences. Thus, he claims: 

There is a statistically significant association between light skin and intelligence” (p190). 

As evidence he cites his own study (Lynn 2002) to claim: 

When the amount of European ancestry in American blacks is assessed by skin color, dark-skinned blacks have an IQ of 85 and light-skinned blacks have an IQ of 92” (p190). 

However, he fails to explain how he managed to divide American blacks into two discrete groups by reference to a trait that obviously varies continuously. 

More importantly, he neglects to mention altogether two other studies that also investigated the relationship between IQ and degree of racial admixture among African-Americans, but used blood-groups rather than skin tone to assess ancestry (Loehlin et al 1973; Scarr et al 1977). 

This is surely a more reliable measure of ancestry than is skin tone, since the latter is affected by environmental factors (e.g. exposure to the sun darkens the skin), and could conceivably have an indirect psychological effect.[42]

However, both these studies found no association between ancestry and IQ (Loehlin et al 1973; Scarr et al 1977).[43] 

Meanwhile, Lynn mentions the Eyferth study (1961) of the IQs of German children fathered by black and white US servicemen in the period after World War II, only to report, “the IQ of African-Europeans [i.e. those fathered by the black US servicemen] was 94 in relation to 100 for European women” (p63). 

However, he fails to mention that the IQ of those German children fathered by black US servicemen (i.e. those of mixed race) was actually almost identical to that of those fathered by white US servicemen (who, with German mothers, were wholly white). This finding is, of course, evidence against the hereditarian hypothesis with respect to race differences. 

Yet Lynn can hardly claim to be unaware of this finding, or its implications with respect to race differences, since this is actually among the studies most frequently cited by opponents of the hereditarian hypothesis with respect to the black-white test score gap for precisely this reason. 

Lynn’s presentation of the evidence regarding the relative contributions of heredity and environment to race differences in IQ is therefore highly selective and biased. 

An Evolutionary Analysis 

Only in the last three chapters does Lynn provide the belated “Evolutionary Analysis” promised in his subtitle. 

Lynn’s analysis is evolutionary in two senses. 

First, he presents both a functionalist explanation of why race differences in intelligence (supposedly) evolved (Chapter 16). This is the sort of ultimate evolutionary explanation with which evolutionary psychologists and sociobiogists are usually concerned. 

However, in addition, Lynn also traces evolution of intelligence over evolutionary history, both in humans of different races (Chapter 17) and among our non-humans and our pre-human ancestors (Chapter 15). 

In other words, he addresses the questions of both adaptation and phylogeny, two of Niko Tinbergen’s famous Four Questions

In discussing the former of these two questions (namely, why race differences in intelligence evolved: Chapter 16), Lynn identifies climate as the ultimate environmental factor responsible for the evolution of race differences in intelligence. 

Thus, he claims that, as humans spread out beyond Africa towards regions further from the equator and hence generally with colder temperatures, especially during winters, the colder climates that these pioneers encountered posed greater challenges for the humans who encountered them in terms of feeding themselves and obtaining shelter etc., and that different human races evolved different levels of intelligence in response to the adaptive challenges posed by such difficulties. 

Hunting vs. Gathering 

The greater problems supposedly posed by colder climates included not just difficulties of keeping warm (i.e. the need for clothing, fires, insulated homes), but also the difficulties of keeping fed. 

Thus, Lynn emphasizes the dietary differences between foragers inhabiting different regions of the world: 

Among contemporary hunter-gatherers the proportions of foods obtained by hunting and gathering varies by hunting and by gathering varies according to latitude. Peoples in tropical and subtropical latitudes are largely gatherers, while peoples in temperate environments rely more on hunting, and peoples in arctic and sub-arctic environments rely almost exclusively on hunting and fishing and have to do so because plant foods are unavailable except for berries and nuts in the summer and autumn” (p227). 

I must confess that I was previously unaware of this dietary difference. However, in my defence, this is perhaps because many anthropologists seem all too ready to overgeneralize from the lifestyles of the most intensively studied tropical groups (e.g. the San of Southern Africa) to imply that what is true of these groups is true of all foragers, and was moreover necessarily also true of all our hunter-gatherer ancestors before they transitioned to agriculture. 

Thus, for example, feminist anthropologists seemingly never tire of claiming that it is female gatherers, not male hunters, who provide most of the caloric demands of foraging peoples. 

Actually, however, this is true only for groups inhabiting tropical climes, where plant foods are easily obtainable all year round, not of hunter-gatherers in general (Ember 1978). 

It is certainly not true, for example, of Eskimos, among whom females are almost entirely reliant on male hunters to provision them for most of the year, since plant foods are hardly available at all except for during a few summer months. 

Similarly, radical-leftist anthropologist Marshall Sahlins famously characterized hunter-gatherer peoples as “The Original Affluent Society”, because, according to his data, they do not want for food and actually have more available leisure-time than do most agriculturalists, and even most modern westerners. 

Unfortunately, however, he relied primarily on data from tropical peoples such as the !Kung San to arrive at his estimates, and these findings do not necessarily generalize to other groups such as the Inuit or other Eskimos

The idea that it was our ancestor’s transition to a primarily carnivorous diet that led to increases in hominid brain-size and intelligence was once a popular theory in paleoanthropology. 

However, it has now fallen into disfavour, if only because it put accorded male hunters the starring role in hominid evolution, with female gatherers relegated to a supporting role, and hence offended the sensibilities of feminists, who have become increasingly influential in academia, even in science. 

Nevertheless, it is seems to be true that, across taxa, carnivores tend to have larger brains than herbivores. 

Of course, non-human carnivores did not evolve the exceptional intelligence of humans.  

However, Desmond Morris in The Naked Ape (reviewed here) argued that, because our hominid ancestors only adopted a primarily carnivorous diet relatively late in their evolution, they were unable to compete with such specialized hunters as lions and tigers in terms of their fangs and claws. They therefore had to adopt a different approach, using intelligence instead or claws and fangs, hence inventing handheld weapons and cooperative group hunting. 

Lynn’s argument, however, is somewhat different to the traditional version of the so-called hunting ape hypothesis, as championed by popularizers like Desmond Morris and Robert Ardley

Thus, in the traditional version, it is the intelligence of early hominids, the descendants all populations of contemporary humans, that increased as a result of the increasing cognitive demands that hunting placed upon us. 

However, Lynn argues that it is only certain races that were subject to such selection, as their dependence on hunting increased as they populated colder regions of the globe. 

Indeed, Lynn’s arguments actually cast some doubt on the traditional version of the hunting ape theory

After all, anatomically modern humans are thought to have first evolved in Africa. Yet if African foragers actually subsisted primarily on a diet of wild plant foods, and only occasionally hunted or scavenged meat to supplement this primarily herbivorous diet, then the supposed cognitive demands of hunting can hardly be invoked to explain the massive increase in hominid brain-size that occurred during the period before our ancestors left Africa to colonize the remainder of the world.[44]

Indeed, Lynn is seemingly clear that he rejects the Hunting Ape Hypothesis, writing that the increases in hominid brain-size after our ancestors “entered a new niche of the open savannah in which survival was more cognitively demanding” occurred, not because of the cognitive demands of hunting, but rather that: 

The cognitive demands of the new niche would have consisted principally of finding a variety of different kinds of foods and protecting themselves from predators” (p202)[45]

Cold Winters Theory’ 

It may indeed be true that surviving in the extreme cold is more difficult than surviving the sometimes extreme heat of tropical climate. After all, around the world, many more people die annually from the extreme cold than from extreme heat (Zhau et al 2021).

Indeed, cold weather may not just be challenging for humans, but rather inimicable to life itself. Thus, the coldest regions of Euasia are invariably arid tundra, whereas, in contrast, tropical rainforests are positively teeming with life.

However, there are several problems with so-called ‘Cold Winters Theory’ as an explanation for the race differences in IQ reported by Lynn. 

For one thing, other species have evidently adapted themselves to colder climates without evolving a level of intelligence as high as human population, let alone that of Europeans and East Asians. 

Indeed, I am not aware of any studies even suggesting a relationship between brain-size or intelligence and the temperature or latitude of their species-ranges among non-human species. However, one might expect to find an association between temperature and brain-size, if only because of Bergmann’s rule

Similarly, Neanderthals were ultimately displaced and driven to extinction throughout Eurasia by anatomically-modern humans, who, at least according to the conventional account, outcompeted Neanderthals due to their superior intelligence and tool-making ability. 

However, whereas anatomically modern humans are thought to have evolved in tropical Africa before spreading outwards to Eurasia, the Neanderthals were a cold-adapted species of hominid who had evolved and thrived in Eurasia during the last Ice age. Therefore, if anatomically-modern humans indeed outcompeted Neanderthals because they were smarter, it was certainly not because they evolved in a colder climate.

At any rate, even if the conditions were indeed less demanding in tropical Africa than in temperate or polar latitudes, then, according to basic Darwinian (and Malthusian) theory, in the absence of some other factor limiting population growth (e.g. warfare, predation, homicide, disease), this would presumably mean that humans would respond to greater resource abundance in the tropics by reproducing until they reached the greater carrying capacity of that environment.   

By the time the carrying capacity of the environment was reached, however, the environment would no longer be so resource-abundant given the greater number of humans competing for its resources. 

This leads me to believe that the key factors selecting for increases in the intelligence of hominids were not ecological but rather social – i.e. not access to food and shelter etc., but rather competition with other humans. 

Also, I remain unconvinced that the environments inhabited by the two races that have, according to Lynn, the lowest average IQs, namely, San Bushmen and Australian Aborigines, are cognitively undemanding. 

These are, of course, the Kalahari Desert and Australian outback (also composed, in large part, of deserts) respectively, two notoriously barren and arid environments.[46]

Meanwhile, the Eskimos occupy what is certainly the coldest, and also undoubtedly one of the most demanding, environments anywhere in the world, and also have, according to Lynn’s own data, the largest brains.

However, according to Lynn’s data, their average IQ is only about 90, high for a foraging group, but well below that of Europeans and East Asians.[47] 

For his part, Lynn attempts to explain away this anomaly by arguing that Arctic Populations were precluded from evolving higher IQs by small and dispersed populations, reflecting of the harshness of the environment. This meant the necessary mutations either never arose or never spread through the population (p153; p239-40; p221).[48]
 
On the other hand, he explains their large brains as reflecting visual memory rather than general intelligence, as well as a lack of mutations for neural efficiency (p153; p240).

However, these seem like post-hoc rationalizations.

After all, if conditions were harsher in Eurasia than in Africa, then this would presumably also have resulted in smaller and more dispersed populations in Eurasia than in Africa. However, this evidently did not prevent mutations for higher IQ spreading among Eurasians. 

Why then, when the environment becomes even harsher, and the population even more dispersed, would this pattern suddenly reverse itself? 
 
Likewise, if whole-brain-size is related to general intelligence, it is inconsistent to invoke specific abilities to explain Inuit brains. 

Thus, according to Lynn, Australian Aborigines have high spatial memory, which is closely related to visual memory. However, also according to Lynn, only their right visual cortex is enlarged (p108-9) and they have small overall brain-size (p108-9; p210; p212). 

Endnotes

[1] Curiously, Lynn reports, this black advantage for movement-time does not appear in the simplest form of elementary task (simple reaction time), where the subject simply has to press a button on the lighting of a light, rather than hitting a specific button, rather than alternative buttons, on the lighting of a particular light rather than other lights (p58). These latter forms of elementary cognitive test presumably involve some greater degree of cognitive processing. 

[2] First, there are the practical difficulties. Obviously, non-human animals cannot use written tests, or an interview format. Designing a maze for laboratory mice may be relatively straightforward, but building a comparable maze for elephants is rather more challenging. Second, and more important, different species likely have evolved different specialized abilities for dealing with specific adaptive problems. For example, migratory birds may have evolved specific spatio-visual abilities for navigation. However, this is not necessarily reflective of high general intelligence, and to assess their intelligence solely on the basis of their migratory ability, or even their general spatio-visual ability, would likely overestimate their general level of cognitive ability. In other words, it reflects a modulardomain-specific adaptation.
Admittedly, the same is true to some extent for human races. Thus, some races score relatively higher on certain types of intellectual ability. For example, East Asians tend to score higher on spatio-visual ability than on verbal ability; Ashkenazi Jews show the opposite pattern, scoring higher in verbal intelligence than in spatio-visual ability; while American blacks score relatively higher in tests involving rote memory than in those requiring abstract reasoning ability. Similarly, as discussed by Lynn, some races seem to have certain quite specific abilities not commensurate to their general intelligence (e.g. Aborigine visual memory). However, in general, both between and within races, most variation in human intelligence loads onto the ‘g-factor’ of general intelligence.

[3] American anthropologist Carleton Coon is credited as the first to first to propose that population differences in skull size reflect a thermoregulatory adaptation to climatic differences (Coon 1955). An alternative theory, less supported, is that it was differing levels of ambient light that resulted in differences in brain-size as between different populations tracing their ancestry to different parts of the globe (Pearce & Dunbar 2011). On this view, the larger brains of populations who trace their descent to areas of greater latitude presumably reflect only the demands of the visual system, rather than any differences in general intelligence. Yet another theory, less politically-correct than these, is so-called Cold Winters Theory, which posits that colder climates placed a greater premium on intelligence, which caused populations inhabiting colder regions of the globe to evolve larger brains and higher levels of intelligence. This is, of course, the theory championed by Lynn himself, and I discuss the problems with this theory the final part of this review.

[4] Curiously however, although, as reported by Lynn, the cold-adapted Eskimos indeed have the largest brains of any human poulation, the same does not seem to be true of another arctic population, namely the reindeer-herding Sámi (or Lapps) of Scandinavia and the Kola Penninsula. On the contrary, anthropologist Carleton Coon reports that the Sámi actually “have very small heads” (The Races of Europe: p266). This would seem to be contrary to  Bermann’s Rule. However, this may be accounted for by the diminutive stature of Sámi. Thus, head-size (and brain-size) also correlates with overall body-size, and Coon also reports that, although small in absolute size, Sámi heads are actually “large in proportion to body size” (The Races of Europe: p303).

[5] Lynn has recently published research regarding differences in IQ across different regions of Italy (Lynn 2010).

[6] Actually, Lynn acknowledges causation in both directions, possibly creating a feedback loop. He also acknowledges other factors in contributing to differences in economic development and prosperity, including the effects of the economic system adopted. For example, countries that adopted communism tend to be poorer than comparable countries that have capitalist economies (e.g. Eastern Europe is poorer than Western Europe, and North Korea poorer than South Korea).  

[7] Incidentally, Lynn cites two studies of Polish IQ, whose results are even more divergent than those of Portugal or Ireland, giving average IQs of 106 and 91 respectively. One of these scores is substantially below the European average, while the other the substantially above. 

[8] Essayist Ron Unz has argued that IQs in Ireland have risen in concert with living standards in Ireland (Unz 2012a; Unz 2012b). However, judging from dates when the studies cited by Lynn in ‘Race Differences in Intelligence’ were published, there is no obvious increase over time. True the earliest study, an MA thesis, published in 1973 gives the lowest figure, with an average IQ of just 87 (Gill and Byrt 1973). This rises to 97 in a study published in 1981 that provided little details on its methodology (Buj 1981). However, it declines again for in the latest study cited by Lynn on Irish IQs, which was published in 1993 but gives average IQs of just 93 and 91 for two separate samples (Carr 1993). In the more recent 2015 edition, Lynn cites a few extra studies, eleven in total. Again, however, there is no obvious increase over time, the latest study cited by Lynn, which was published in 2012, giving an average IQ of just 92 (2015 edition).

[9] While this claim is made in reference to immigrants to America and the West, it is perhaps worth noting that East Asians in South-East Asia, namely the Overseas Chinese, largely dominate the economies of South-East Asia, and are therefore on average much wealthier than the average Chinese person still residing in China (see World on Fire by Amy Chua). Given the association of intelligence with wealth, this would suggest that Chinese immigrants to South-East Asia are not substantially less intelligent than those who remained in China. Did the more intelligent Chinese migrate to South-East Asia, while the less intelligent migrated to America? If so, why would this be?

[10] According to Daniel Nettle in Personality: What Makes You the Way You Are, in the framework of the five-factor model of personality, a liking for travel is associated primarily with extraversion. One study found that an intention to migrate was positively associated with both extraversion and openness to experience, but negatively associated with agreeableness, conscientiousness, and neuroticism (Fouarge et al 2019). A study of migration within the United States found a rather more complex set of relationships between migration and each of the big five personality traits (Jokela 2009).

[11] Other Catholic countries, namely those in Southern Europe, such as Italy and Spain, may indeed have slightly lower IQs, at least in the far south of these countries. However, as we have seen, Lynn explains this in terms of racial admixture from Middle-Eastern and North African populations. Therefore, there is no need to invoke priestly celibacy in order to explain it. The crucial test case, then, is Catholic countries other than Ireland from Northern Europe, such as Austria and France.

[12] In the 2015 edition, he returns to a slightly higher figure of 71.

[13] In the 2006 edition, Lynn cites no studies from the Horn of Africa. However, in the 2015 edition, he cites five studies from Ethiopia, and, in The Intelligence of Nations, he and co-author David Becker also cite a study on Somalian IQs.

[14] Indeed, physical anthropologist John Baker, in his excellent Race (which I have reviewed here, here and here) argues that:

The ‘Aethiopid’ race of Ethiopia and Somaliland are an essentially Europid subrace with some Negrid admixture” (Race: p225).

This may be an exaggeration. However, recent genetic studies indeed show affinities between populations from the Horn of Africa and those from the Middle East (e.g. Ali et al 2020; Khan 2011a; Khan 2011b; Hodgson 2014).

[15] However, it is not at all clear that the same is true for black African minorities resident in other western polities, whose IQs are also, according to Lynn’s data, also considerably above those for indigenous Africans. Here, I suspect black populations are more diverse.
For example, in Britain, Afro-Caribbean people, who emigrated to Britain by way of the West Indies, are probably mostly mixed-race, like African-Americans, since both descend from white-owned slave populations. However, Britain also plays host to many immigrants direct from Africa, most of whom are, I suspect, of relatively unmixed sub-Saharan African descent. Yet, despite having greater levels of sub-Saharan African DNA, African immigrants to the UK outperform Afro-Caribbeans in UK schools, just as they do African-Americans in the US (Chisala 2015a).

[16] Blogger John ‘Chuck’ Fuerst suggests, the higher scores for Somali immigrants might reflect the fact that peoples from the Horn of Africa actually, as we have seen, have genetic affinities with North African and Middle Eastern populations (Fuerst 2015). However, the problem with attributing the relatively high scores of Somali refugees and immigrants to Caucasoid-admixture is that, as we have seen, according to the data collected by Lynn, IQs are no higher in the Horn of Africa than elsewhere in sub-Saharan Africa.

[17] If anything, “Bushmen” should presumably be grouped, not with Pygmies, with rather the distinct but related Khoikhoi pastoralists. However, the latter are now all but extinct as an independent people and are not mentioned by Lynn.

[18] For example, Lynn also acknowledges that those whom he terms “South Asians and North Africans” are “closely related to the Europeans” (p79). However, they nevertheless merit a chapter of their own. Likewise, he acknowledges that “South-East Asians” share “some genetic affinity with East Asians with whom they are to some degree interbred” (p97). Nevertheless, he justifies considering these two ostensible races in separate chapters, partly on the basis that “the flattened nose and epicanthic eye-fold are less prominent” among the former (p97). Yet the morphological differences between Pygmies and Khoisan are even greater, but they are lumped together in the same chapter.

[19] There is indeed, as Lynn notes, a correlation between a group’s IQ and their lifestyle (i.e. whether they are foragers or agriculturalists). However, the direction of causation is unclear. Does high intelligence allow a group to transition to agriculture, or does an agriculturalist lifestyle somehow increase a group’s average IQ? And, if the latter, is this a genetic or a purely environmental effect?

[20] Indeed, the very word slave is thought to derive from the ethnonym Slav, because of the frequency with which Slavic peoples were enslaved during the Middle Ages. Often they were enslaved by Muslims, the Ottoman Turks having conquered much of Southeast Europe. Other times they were enslaved by Europeans and thence often sold on to the Ottoman Turks. As the last peoples in Europe to be Christianized, Slavs were long vulnerable to enslavement by both Muslims and Christians since, just as Islamic law forbade the enslavement of fellow Muslims, so Papal degree long prohibited the capture and enslavement of other Christians. Indeed, it is claimed that non-Slavic captives from elsewhere in Europe were often falsely described as Slavs in order to justify their enslavement.

[21] Indeed, Lynn could hardly have arrived at an actual figure for the average Pygmy IQ, since, as we have seen, he reports the results of only a single actual study of Pygmy intelligence, the author of which did not present his results in a quantitative format.

[22] Thus, he suggests that the lower performance of the Aboriginals tested by Drinkwater (1975), as compared to those tested by Kearins (1981), may reflect the fact that the latter were the descendants of coastal populations of Aborigines, for whom the need to navigate in deserts without landmarks would have been less important. 

[23] The fact that the earliest civilization emerged among Middle Eastern, North African and South Asian populations is attributed by Lynn to the sort of environmental factors of the sort that, elsewhere in his book, he largely discounts or downplays. Thus, Lynn writes: 

“[Europeans] were not able to develop early civilizations like those built by the South Asians and North Africans because Europe was still cold, was covered with forest, and had heavy soils that were difficult to plough unlike the light soils on which the early civilizations were built, and there were no river flood plains to provide annual highly fertile alluvial deposits from which agricultural surpluses could be obtained to support an urban civilization and an intellectual class” (p237).

[24] An interesting question is whether there exist differences in IQ as between different caste groups within the Indian subcontinent, since, at least in theory, these represented endogamous breeding populations between whom strict separation was maintained. Thus, it would be interesting to know the average IQ of Brahmins or of the high-achieving Parsi people (though the latter are not strictly a caste, since they are not Hindu).
In general, cross-culturally, people of higher socio-economic status have, on average, higher IQs, presumably because higher intelligence facilitates upward social mobility whereas low intelligence is associated with downward movility. However, since, at least in theory, Indian caste status was determined by birth, this, at least in theory, almost totally precluded even the very possibility of upward or downward mobility between castes. Therefore, the general finding that higher socio-economic status is associated with higher intelligence may not hold for Indian castes, or, more likely, the association betweeen intelligence and caste may be much weaker than for other societies.

[25] However, all of these comparisons, in both Britain and America, omit to include Jewish people as a separate ethnicity, instead grouping them with other whites. Jews earn more, on average, than members of any other religious group in Britain and America, including Hindus.

[26] I assume that this is the study that Lynn is citing, since this is the only matching study included in his references. However, curiously, Lynn refers to this study here as “Mackintosh et al 1985” (p83-4), despite their being only two authors listed in his references, such that “Mackintosh & Mascie-Taylor 1985” would be the more usual citation. Indeed, Lynn uses this latter form of citation (i.e. “Mackintosh & Mascie-Taylor 1985”) elsewhere when citing what seems to be the same paper in his earlier chapter on Africans (p47; p49).

[27] In order to determine whether religion or national origin is the key determining factor, it would be interesting to have data on the incomes (and IQs) of Pakistani Hindus, Bangladeshi Hindus and Muslim Indians resident in the West. However, I have not been able to find any such data.
Interestingly, author and physician AS Amin, in a goodreads book review, suggests that a difference in ability between South Asian Muslims and Hindus may reflect the fact that it was disproportionately lower caste Indians who converted to Islam so to escape caste discrimination, just as it is also documented that many lower caste Indians also converted to Buddhism and Christianity for the same reason.
On the other hand, however, geography seems also to have been a factor, with certain regions of the sub-continent converting en masse, which would have diluted any such effect.

[28] An alternative possibility is that it was the spread of Arab genes, as a result of the Arab conquests, and resulting spread of Islam, that depressed IQs in the Middle-East and North Africa, since Arabs were, prior to the rise of Islam, a relatively backward group of desert nomads, whose intellectual achievements were minimal compared to those of many of the groups whom they conquered (e.g. Persians, Mesopotamians, Assyrians, and Egyptians). Indeed, even the achievements of Muslim civilization during the Islamic Golden Age seem to have been disproportionately those of Persian converts, not the Arabs themselves. 

[29] One might, incidentally, question Lynn’s assumption that Oriental Jews were less subject to persecution than were the Ashkenazim in Europe. This is, of course, the politically correct view, which sees Islamic civilization as, prior to recent times, more tolerant than Christendom. On this view, anti-Jewish sentiment only emerged in the Middle East as a consequence of Zionism and the establishment of the Jewish state in what was formerly Palestine. However, for alternative views, see The Myth of the Andalusian Paradise. See also Robert Spencer’s The Truth About Muhammad (which I have reviewed here), in which he argues that Islam is inherently antisemitic (i.e. anti-Jewish).
Interestingly, Kevin Macdonald, in A People That Shall Dwell Alone (which I have reviewed here) makes almost the opposite argument to that of Lynn. Thus, he argues that it was precisely because Jews were so discriminated against in the Muslim world that their culture, and ultimately their IQs, were to decline, as they were, according to Macdonald, largely excluded from high-status and cognitively-demanding occupations, which were reserved for Muslims (p301-4). Thus, Macdonald concludes: 

The pattern of lower verbal intelligence, relatively high fertility, and low-investment parenting among Jews in the Muslim world is linked ultimately to anti-Semitism” (A People That Shall Dwell Alone (reviewed here): p304). 

[30] Lynn, for his part, does not explain why he believes persecution supposedly select for higher intelligence, simply assuming that it is logial that it would.

[31] This pattern among East Asians of lower scores on the verbal component of IQ tests was initially attributed to a lack of fluency in the language of the test, since the first East Asians to be tested were among diaspora populations resident in the West. However, the same pattern has now been found even among East Asians tested in their first language, in both the West and East Asia.

[32] For example, Sarich and Miele, in Race: The Reality of Human Differences (which I have reviewed here and here) write that “Asians have a slightly higher IQ than do whites” (Race: The Reality of Human Differences: p196). However, in fact, this applies only to East Asians, not to South-East Asians (nor to South Asians and West Asians, who are “Asian” in at least the strict geographical, and the British-English, sense.) Similarly, in his own oversimplified tripartite racial taxonomy in Race, Evolution and Behavior (which I have reviewed here), Philippe Rushton seems to imply that the traits he attributes to Mongoloids, including high IQs and large brain-size, apply to all members of this race, including South-East Asians and even Native Americans.

[33] Ethnic Chinese were overrepresented among Vietnamese boat people, though less so among later waves of immigrants. However, perhaps a greater problem is that they were disproportionately middle-class and drawn from the business elite, and hence unrepresentative of the Vietnamese as a whole, and likely of disproportionately high cognitive ability, since higher social classes tend to have higher average IQs.

[34] In his paper on Mongolian IQs, Lynn also suggests that Mongolians have lower IQs than other East Asians because they are genetically intermediate between East Asians and Eskimos (“Arctic Peoples”), who themselves have lower IQs (Lynn 2007). However, this merely begs the question as to why Eskimos themselves have lower IQs than East Asians, another anomaly with respect to Cold Winters Theory, which is discussed in the final part of this review.

[35] With regard to the population of Colombia, Lynn writes: 

The population of Colombia is 75 percent Native American and Mestizo, 20 percent European, and 5 percent African. It is reasonable to assume that the higher IQ of the Europeans and the lower IQ of the Africans will approximately balance out and that the IQ of 84 represents the intelligence of the Native Americans” (p58). 

However, this assumption that the African and European genetic contributions will balance out seems dubious since, by Lynn’s own reckoning, the European contribution to the Colombian gene-pool is three times greater than that of Africans.

[36] The currently-preferred term Inuit is not sufficiently inclusive, because it applies only to those Eskimos indigenous to the North American continent, not the related but culturally distinct populations inhabiting Siberia or the Aleutian Islands. I continue to use the term Eskimos, because it is more accurate, not obviously pejorative, probably more widely understood, and also because I deplore the euphemism treadmill. Elsewhere, I have generally deferred to Lynn’s own usage, for example mostly using ‘Aborigine’, rather than the now preferred ‘Aboriginal’, a particularly preposterous example of the euphemism treadmill since the terms are so similar, comparable to how, today, it is acceptable to say ‘people of colour’, but not ‘coloured people’.

[37] For example, Hans Eysenck made various references in his writings to the fact that Eskimo children performed as well as European children in IQ tests as evidence for his claim that economic deprivation did not necessarily reduce IQ scores (e.g. The Structure and Measurement of Intelligence: p23). See also discussion in: Jason Malloy, A World of Difference: Richard Lynn Maps World Intelligence (Malloy 2016).

[38] Certain specific subpopulations also score higher (e.g. Ashkenazim and Māoris, though the latter only barely). However, these are subpopulations within the major ten races that Lynn identifies, not races in and of themselves.

[39] Actually, by the time Columbus landed in the Americas, many Native Americans had already partly transitioned to agriculture. However, not least because of a lack of domesticated animals that they could use as a meat source, most supplemented this with hunting and sometimes gathering too.

[40] However, Lynn reports that Japanese also score high on tests of visual memory (p143). However, excepting perhaps the Ainu, the Japanese do not have a recent history of subsisting as foragers. This suggests that foraging is not the only possible cause of high visual memory in a population.

[41] Presumably the comparison group Lynn has in mind are Europeans, since, as we have seen it is European living standards that he takes as his baseline for the purposes of estimating a group’s ”genotypic IQ” (p69), and, in a sense, all the IQ scores that he reports are measured against a European standard in so far as they are calculated by reference to an arbitrarily assigned average of 100 for European populations.

[42] Thus, it is at least theoretically possible that a relatively darker-skinned African-American child might be treated differently than a lighter-skinned child, especially one whose race is relatively indeterminate, by others (e.g. teachers) in a way that could conceivably affect their cognitive development and IQ. In addition, a darker skinned African-American child might, as a consequence of their darker complexion, come to identify as an African American to a greater extent than a lighter skinned child, which might affect who they socialize with, which celebrities they identify with and the extent to which they identify with broader black culture, all of which could conceivably have an effect on IQ. I do not contend that these effects are likely or even plausible, but they are at least theoretically possible. Using blood group to assess ancestry, especially if one actually introduces controls for skin tone (since this may be associated with blood-group, since both are presumed to be markers of degree of African ancestry), obviously eliminates this possibility. Today, this can also be done by looking at subjects’ actual DNA, which obviously has the potential to provide a more accurate measure of ancestry than either skin-tone or blood-group (e.g. Lasker et al 2019).

[43] More recently, a better study has been published regarding the association between European admixture and intelligence among African-Americans, which used genetic data to assess ancestry, and actually sought to control for the possible confounding effect of skin-colour and appearance (Lasker et al 2019). Unlike the blood-group studies, this largely supports the hereditarian hypothesis. However, this was not available at the time Lynn authored his book. Also, it ought to be noted that it was published in a controversial pay-to-publish academic journal, and therefore the quality of peer review to which the paper was subjected may be open to question. No doubt in the future, with the reduced costs of genetic testing, more studies using a similar methodology will be conducted, finally resolving the question of the relative contributions of heredity and environment to the black-white test score gap in America, and perhaps disparities between other ethnic groups too.

[44] It is a fallacy, however, to assume that what is true for those foraging peoples that have managed to survive as foragers in modern times and hence come to be studied by anthropologists was necessarily also true of all foraging groups before the transition to agriculture. On the contrary, those foraging groups that have survived into modern times, tend to have done so only in the ecologically most marginal and barren environments (e.g. the Kalahari Desert occupied by the San), since these areas are of least use to agriculturalists, and therefore represent the only regions where more technologically and socially advanced agriculturalists have yet to displace them (see Ember 1978). However, this would seem to suggest that African hunter-gatherers, prior to the expansion of Bantu agriculturalists, would have occupied more fertile areas, and therefore might have had even less need to rely on hunting than do contemporary hunter-gatherers such as the San, who are today largely restricted to the Kalahari Desert.

[45] Here, interestingly, Lynn departs from the theory of fellow race realist, and fellow exponent of ‘Cold Winters Theory’, Philippe Rushton. The latter, in his book, Race, Evolution and Behavior (which I have reviewed here), argues that: 

Hunting in the open grasslands of northern Europe was more difficult than hunting in the woodlands of the tropics and subtropics where there is plenty of cover for hunters to hide in” (Race, Evolution and Behavior: p228). 

In contrast, Lynn argues “open grasslands”, albeit on the African Savannah rather than in Northern Europe, actually made things harder, not for predators, but rather for prey – or at least arboreal primate prey. Thus, Lynn writes: 

The other principle problem of the hominids living in open grasslands would have been to protect themselves against lions, cheetahs and leopards. Apes and monkeys escape from the big cats by climbing into trees and swinging or jumping form one tree to another. For the Autralopithecines and the later hominids in open grasslands this was no longer possible” (p203). 

[46] To clarify, this is not to say that either San Bushmen or Australian Aborigines evolved primarily in these desert environments. On the contrary, many of them formerly occupied more fertile areas, before being displaced by more advanced neighbours, Bantu agriculturalists in the case of Khoisan, and European (more specifically British) colonizers, in the case of Aborigines. However, that they are nevertheless capable of surviving in these demanding desert environments suggests either:

(1) They are more intelligent than Lynn concludes; or
(2) That surviving in challenging environments does not require the level of intelligence that Lynn’s Cold Winters Theory supposes.

[47] Besides Eskimos, another potential test case for ‘Cold Winters Theory’ are the Sámi (or Lapps) of Northern Scandinavia. Like Eskimos, they have inhabited an extremely cold, northern environment for many generations and are genetically, and morphologically, quite distinct from other populations. Also, again like Eskimos, they maintained a foraging lifestyle until modern times. However, unlike other cold-adapted populations, the Sámi have, according to Carleton Coon, “very small heads” and hence presumably not especially large brains, though he also reports that their head-size is actually large in proportion to body-size. (The Races of Europe: p266; p303). According to Armstrong et al (2014), the only study of Sámi cognitive ability of which I am aware, the average IQ of the Sámi is almost identical to that of neighbouring populations of Finns (about 101).

[48] Lynn gives the same explanation for the relatively lower recorded IQs of Mongolians, as compared to other East Asians (p240).

References

Ali et al (2020) Genome-wide analyses disclose the distinctive HLA architecture and the pharmacogenetic landscape of the Somali population. Science Reports 10:5652.

Anderson M (2015) Chapter 1: Statistical Portrait of the U.S. Black Immigrant Population. In A Rising Share of the U.S. Black Population Is Foreign Born. Pew Research Center: Social & Demographic Trends, April 9, 2015. 

Armstrong et al (2014) Cognitive abilities amongst the Sámi population. Intelligence 46: 35-39.

Aziz et al (2004) Intellectual development of children born of mothers who fasted in Ramadan during pregnancy International Journal for Vitamin and Nutrition Research (2004), 74, pp. 374-380.

Beals et al (1984) Brain Size, Cranial Morphology, Climate, and Time Machines. Current Anthropology 25(3), 301–330.

Buj (1981) Average IQ values in various European countries Personality and Individual Differences 2(2): 168-9.

Carr (1993) Twenty Years a Growing: A Research Note on Gains in the Intelligence Test Scores of Irish Children over Two Decades Irish Journal of Psychology 14(4): 576-582.

Chisala (2015a) The IQ Gap Is No Longer a Black and White IssueUnz Review, 25 June. 

Chisala (2015b) Closing the Black-White IQ Gap Debate, Part I, Unz Review, 5 October.

Chisala (2015c) Closing the Black-White IQ Gap Debate, Part 2Unz Review, 22 October. 

Chisala (2019) Why Do Blacks Outperform Whites in UK Schools? Unz Review, November 29

Coon (1955) Some Problems of Human Variability and Natural Selection in Climate and Culture. American Naturalist 89(848): 257-279

Drinkwater (1975) Visual memory skills of medium contact aboriginal childrenAustralian Journal of Psychology 28(1): 37-43. 

Dutton (2020) Why Islam Makes You Stupid . . . But Also Means You’ll Conquer The World (Whitefish, MT: Washington Summit, 2020).

Ember (1978) Myths about Hunter-Gatherers Ethnology 17(4): 439-448 

Eyferth (1959) Eine Untersuchung der Neger-Mischlingskinder in Westdeutschland. Vita Humana, 2:102–114. 

Fouarge et al (2019) Personality traits, migration intentions, and cultural distance. Papers in Regional Science 98(6): 2425-2454

Fuerst (2015) The Measured proficiency of Somali Americans, HumaVarieties.org

Gill & Byrt (1973). The Standardization of Raven’s Progressive Matrices and the Mill Hill Vocabulary Scale for Irish School Children Aged 6–12 Years. University College, Cork: MA Thesis.

Hodgeson et al (2014) Early Back-to-Africa Migration into the Horn of Africa. PLoS Genetics 10(6): e1004393.

Jokela (2009) Personality predicts migration within and between U.S. states Journal of Research in Personality 43(1): 79-83.

Kearins (1986) Visual spatial memory in aboriginal and white Australian childrenAustralian Journal of Psychology 38(3): 203-214. 

Kearins (1981) Visual spatial memory in Australian Aboriginal children of desert regions Cognitive Psychology 13(3): 434-460. 

Khan (2011a) The genetic affinities of Ethiopians. Discover Magazine, January 10.

Khan (2011b) A genomic sketch of the Horn of Africa. Discover Magazine, June 10

Klekamp et al (1987) A quantitative study of Australian aboriginal and Caucasian brains. Journal of Anatomy 150: 191–210.

Knapp & Seagrim (1981) Visual memory Australian aboriginal children and children of European descent International Journal of Psychology 16(1-4): 213-231. 

Langan & LoSasso (2002) Discussions on Genius and Intelligence: Mega Foundation Interview with Arthur Jensen‘ (Eastport, New York: MegaPress) 

Lasker et al (2019) Global ancestry and cognitive abilityPsych 1(1), 431-459 

Loehlin et al (1973) Blood group genes and negro-white ability differences. Behavior Genetics 3(3): 263-270 

Lynn (2002) Skin Color and Intelligence in African-Americans. Population & Environment 23:201-207 

Lynn (2007) IQ of Mongolians. Mankind Quarterly 47(3).

Lynn (2010) In Italy, north–south differences in IQ predict differences in income, education, infant mortality, stature, and literacy. Intelligence, 38, 93-100. 

Lynn (2015) Selective Emigration, Roman Catholicism and the Decline of Intelligence in the Republic of Ireland. Mankind Quarterly 55(3): 242-253.

Mackintosh (2007) Review of Race differences in intelligence: An Evolutionary Hypothesis [sic], by Richard Lynn, Intelligence 35,(1): 94-96

Mackintosh & Mascie-Taylor (1985). The IQ question. In Education for All. Cmnd paper 4453. London: HMSO. 

Malloy (2014) HVGIQ: VietnamHumanvarieties.org, June 19. 

Malloy (2006) A World of Difference: Richard Lynn Maps World Intelligence. Gnxp.com, February 01. 

Pearce & Dunbar (2011) Latitudinal variation in light levels drives human visual system size, Biology Letters, 8(1): 90–93. 

Pereira et al (2005). African female heritage in Iberia: a reassessment of mtDNA lineage distribution in present timesHuman Biology77 (2): 213–29. 

Richards et al (2003) Extensive Female-Mediated Gene Flow from Sub-Saharan Africa into Near Eastern Arab PopulationsAmerican Journal of Human Genetics 72(4):1058–1064.

Rushton, J. P., & Ankney, C. D. (2009). Whole brain size and general mental ability: A reviewInternational Journal of Neuroscience119, 691-731

Sailer (1996) Great Black HopesNational Review, August 12

Scarr et al (1977) Absence of a relationship between degree of white ancestry and intellectual skills within a black population. Human Genetics 39(1):69-86 

Templer (2010) The Comparison of Mean IQ in Muslim and Non-Muslim CountriesMankind Quarterly 50(3):188-209 

Torrence (1983) Time budgeting and hunter-gatherer technology. In G. Bailey (Ed.). Hunter-Gatherer Economy in Prehistory: A European Perspective. Cambridge, Cambridge University Press.

West et al (1992) Cognitive and educational attainment in different ethnic groups, Journal of Biosocial Science 24(4): 539-554.

Woodley (2009) Inbreeding depression and IQ in a study of 72 countries Intelligence 37(3): 268-276

Zhau et al (2021) Global, regional, and national burden of mortality associated with non-optimal ambient temperatures from 2000 to 2019: a three-stage modelling study, Lancet 5(7): E415-E425