Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Current Biology

Review

The Molecular Basis of Human Brain Evolution


Wolfgang Enard
Department of Biology II, Ludwig Maximilian University Munich, Grosshaderner Str. 2, D-82152 Martinsried, Germany
Correspondence: enard@bio.lmu.de
http://dx.doi.org/10.1016/j.cub.2016.09.030

Humans are a remarkable species, especially because of the remarkable properties of their brain. Since the
split from the chimpanzee lineage, the human brain has increased three-fold in size and has acquired abilities
for vocal learning, language and intense cooperation. To better understand the molecular basis of these
changes is of great biological and biomedical interest. However, all the about 16 million fixed genetic changes
that occurred during human evolution are fully correlated with all molecular, cellular, anatomical and behav-
ioral changes that occurred during this time. Hence, as humans and chimpanzees cannot be crossed or
genetically manipulated, no direct evidence for linking particular genetic and molecular changes to human
brain evolution can be obtained. Here, I sketch a framework how indirect evidence can be obtained and re-
view findings related to the molecular basis of human cognition, vocal learning and brain size. In particular, I
discuss how a comprehensive comparative approach, leveraging cellular systems and genomic technolo-
gies, could inform the evolution of our brain in the future.

Introduction Independently of the practical limitations to measure pheno-


The evolution of our own species fascinates many of us and is types, all phenotypic changes and all genetic changes that
relevant to understanding our biological and biomedical condi- occurred before the last common ancestor of currently living hu-
tions [1,2]. As our ecological niche is apparently dependent on mans are fully correlated (Figure 1). As crosses between humans
special cognitive features, the evolution of our brain is of partic- and chimpanzees are not an option (see [4] for a historical ac-
ular relevance. The human brain has a long evolutionary history, count of an attempt in the 1920s), the power of quantitative sys-
but of special interest is the period after we split from our closest tems genetics [5,6] cannot be used to disentangle genotypes and
living relatives, the common chimpanzee (Pan troglodytes) and molecular, cellular and behavioral phenotypes. Furthermore, it is
the bonobo (Pan paniscus). Genetic and phenotypic changes not possible — again for obvious ethical reasons — to test gene
that occurred after this split and that are common to all currently function by genetically manipulating humans or chimpanzees.
living humans can be defined as ‘human-specific’ (Figure 1). In For all of these reasons, direct proof for linking particular ge-
this review, I will discuss how changes in human brain pheno- netic changes to particular phenotypic changes in humans
types after the human–chimpanzee split can be linked to cannot be obtained. In this review, I will sketch a framework of
cellular, molecular and genetic changes. To define human-spe- how plausible indirect evidence can and has been obtained
cific changes, comparisons need to include humans, chimpan- to tackle this question nevertheless. I will then illustrate this
zees and at least one primate outgroup, such as the orangutan approach by discussing recent findings related to the evolution
or the rhesus macaque (Figure 1). Fortunately, the increasing of human social cognition, language and speech as well as brain
availability of genome sequences from humans, chimpanzees size and how the prospective increase of genomic information
and other primates, as well as even our closest extinct relatives, and cellular model systems from many primate and mammalian
the Neanderthals and Denisovans [3], allows us to precisely species should allow a more comprehensive comparative
identify most of the about 16 million genetic changes that approach to human brain evolution.
occurred in the human lineage, i.e. after the split from the chim-
panzee lineage and before the most recent common ancestor of A Framework for Quantitative Systems Genetics without
currently living humans (Figure 1, Box 1). Crossing
In contrast to these fairly well defined, discrete genetic One can view human brain evolution as a problem of doing
changes, the gradual changes in molecular, cellular, anatom- quantitative systems genetics with complete access to geno-
ical and behavioral phenotypes are much more difficult to reli- types, limited access to phenotypes and no possibility to do
ably measure in a systematic manner, particularly as access crosses and genetic manipulations. To still reconstruct plau-
to samples is limited for ethical and practical reasons. Espe- sible scenarios of causal relationships between genotypes
cially for embryos it is practically and ethically impossible to and phenotypes during human brain evolution, one needs to
obtain invasive samples from chimpanzees and other great add information from additional sources. This includes the
apes. Despite these limitations to comprehensively define inference of positive selection on genes or certain regions of
human-specific brain phenotypes across different levels and the genome, genotype–phenotype associations within hu-
during development, substantial progress has been made mans, genotype–phenotype associations within model organ-
in recent years to measure anatomical, cellular and mo- isms and genotype–phenotype associations across species.
lecular phenotypes in humans, chimpanzees and other pri- Such data can — and probably most often must — be com-
mates (Box 2). plemented with experimental approaches that test particular

Current Biology 26, R1109–R1117, October 24, 2016 ª 2016 Elsevier Ltd. R1109
Current Biology

Review
Figure 1. Genetic and phenotypic changes
in the human lineage.
All genetic and phenotypic changes that occurred
Humans
after the split from the chimpanzee lineage and
~14 million substitutions before the most recent common ancestor of
Social cognition currently living humans are fully correlated. The
~2 million indels Vocal learning
~24,000 amino acid changes Neanderthal tree is scaled according to divergence times esti-
Brain size
mated for apes and humans in [111] and [112],
Denisovan respectively. The number of substitutions are
taken from [113], assuming that 80% of all genetic
Chimpanzee changes are fixed in currently living humans [100].
Brain images with permission from [114] and
Bonobo Dean Falk.

Gorilla

Orangutan
In any case, the affected phenotype,
~ 14 million years and ~ 1.5 substitutions/100 bp which is the cause of positive selec-
Current Biology
tion, must be inferred from additional
data. Thus, linking positive selection in
genomic elements to human brain evolu-
genotype–phenotype hypotheses in an experimental system tion requires additional evidence. This is especially true as evi-
(Figure 2). dence for positive selection during human evolution is found in
Genotype–phenotype associations within humans are maybe many protein coding genes but is rarely specific to the human
the most important type of information used to interpret genetic lineage. The reason for this is unclear, but involvement of genes
changes during human evolution. Fortunately, more and more in pathogen interactions [16] or non-adaptive processes, such as
resources are available that annotate genes and regulatory re- compensatory evolution that can lead to positive selection
gions by charting mRNA expression and epigenetic modifica- without changes in function [14,17,18], could be responsible.
tions with high spatial and temporal resolution in the human brain Hence, it is crucial to evaluate positive selection occurring on
[7,8]. Furthermore, common and rare genetic variants within the human lineage in a comparative context of several primate
humans can be linked to human phenotypes, including psychiat- or mammalian lineages. A comparative approach has a long
ric diseases, language related phenotypes or brain structures tradition in evolutionary biology for correlating two or more
[9–11]. For model organisms, such as mice, similar resources phenotypic traits across species [19,20] and should be informa-
exist that complement the human data as genetic and other tive also for correlating genotype–phenotype relationships
experimental manipulations can be done during all stages of across species, as discussed recently in the context of brain
development [12]. size evolution [21].
While this annotation of the human genome allows re- While the above sources of information are needed to
searchers to categorize human-specific genetic changes based constrain hypotheses regarding the genetic and molecular basis
on whether they occur in coding regions, putative regulatory of human brain evolution, they will usually not be sufficient.
regions or putative non-functional regions, the vast amount of Experimental follow-up of hypotheses will be necessary. Which
these changes are thought to be neutral, i.e. having no effect functional readout is informative will obviously depend on the
on fitness or on the brain phenotypes of interest. To identify context and available knowledge. For example, absorption
genes or regulatory elements that could have been affected by spectra of opsin proteins are probably a very informative readout
positive selection and whose alleles could have had a selective to infer color vision of mammals [22], but which readouts might
advantage during human evolution, one uses statistical methods be informative to infer effects on brain size or even cognition is
comparing DNA sequence variation within and between humans much less clear. Investigating evolutionary changes in the rele-
and other species [13]. While this is an elegant way to make use vant cell types is important because it is expected that positively
of the available genomic information, it is important to realize that selected genetic variants are specific to particular cell types and
the power to identify such positively selected elements is low, developmental stages in order to avoid negative side effects
especially for genetic changes that are fixed in humans. This is of the variant in other contexts [23]. Hence, the possibility to
because a signature of positive selection can only be detected generate different cell types and brain organoids from induced
when either several substitutions in the genetic element were pluripotent stem cells is a crucial development to study human
affected by positive selection or when positive selection on a sin- brain evolution, as discussed in more detail below. However,
gle allele was so recent that the so-called ‘hitchhiking’ effect on cellular models are of course limited, especially when complex
the patterns of variation on the linked genomic regions deviates behavioural phenotypes are concerned. Although genetically
significantly from expectations of neutrality. So even when modified primates are now technically feasible [24], they might
considering that better primate genome sequences and align- for ethical and financial reasons not be a major option to study
ments should prevent an excess of false positives that can be human brain evolution. Hence, mice will remain a major tool
an issue with current genome assemblies [14] and that better to investigate particular hypotheses on human brain evolution
statistical models allow consideration of effects such as biased [25]. However, genome sequencing and genome editing greatly
gene conversion [15], the false negative rate will remain high at facilitate experimental investigations also in other organisms and
a given false positive rate. the contribution of ferrets to mechanisms of brain size evolution

R1110 Current Biology 26, R1109–R1117, October 24, 2016


Current Biology

Review

Box1. Genetic changes to the human lineage.

The sequencing of the chimpanzee genome allowed researchers to identify 35 million nucleotide substitutions that differ between
the human reference genome and the chimpanzee reference genome [113]. Half of these occurred in the human lineage and half of
them in the chimpanzee lineage. The sequencing of an outgroup genome, such as that of the orangutan [115] or the rhesus ma-
caque [116] and in particular the resequencing of 79 ape genomes [111], makes it possible to reliably assign single nucleotide
changes to the human lineage. Less than 20% of these are expected to have occurred after the last common ancestor of all
currently living humans [100,113], which means they are still polymorphic in extant humans, resulting in an estimated 14 million
(80% of half of the 35 million nucleotide differences between human and chimpanzee) fixed nucleotide substitutions in the human
lineage (Figure 1). A small proportion of these human-specific changes (31,389 identified according to [3]) occurred after the
split of modern humans from Neanderthals and Denisovans, our closest relatives for which genome sequences are also available,
allowing unique insights into changes specific to modern humans [3]. Over 90% of the genetic changes will not be relevant as less
than 10% of the human genome is evolving under constraint, i.e. is at all functional [99,117]. Changes in the known functional parts
of the genome include about 24,000 fixed encoded amino acid changes as about 60,000 amino acid differences are expected to
occur in 20,000 protein-coding genes between human and chimpanzee [113]. However, protein-coding regions make up only 1%
of the human genome. The remaining less than 9% of functional regions are much more difficult to localize, but they are strongly
enriched in conserved non-coding regions and in regions showing epigenetic signatures of regulatory regions [8,15]. This informa-
tion has also been used to identify functional non-coding regions that are enriched for human-specific substitutions [15], and as the
human genome annotation gets better for non-coding regions, this will allow a more precise annotation of human-specific
changes.
In addition to nucleotide substitutions, humans and chimpanzees differ by about five million insertions and deletions of DNA [113].
While most of them affect just a few base pairs, more than 1000 affect large regions (larger than 2 kilobases) spanning up to
10 million bases [118] totaling about 100 million base pairs of affected sequence between a human and a chimpanzee genome.
Technically, these changes are more difficult to pinpoint, but an analysis of 97 resequenced ape and human genomes allowed
a fairly detailed assessment of rates and events, including about 60 genes that have been affected by deletions and duplications
of exons in the human lineage [118]. While it will be important to increase the quality of primate genomes, especially with respect to
such structurally complex regions [119], genetic information is no longer a limiting factor.
In summary, there are 14 million nucleotide substitutions and about two million insertions and deletions fixed in the human lineage
of which less than 10% will fall into annotated functional regions, including about 24,000 encoded amino acid changes (Figure 1).
Also the vast majority of changes in functional regions are expected to be neutral, i.e. to have no effect on a molecular or organismic
level, and the major problem is to predict which genetic changes might have functional consequences in which phenotypic
context.

and zebra finches to the evolution of vocal learning show the especially in their social cognition [29], allowing intensive coop-
importance of using a wider array of model organisms to under- eration, including morality [30] and in turn cumulative culture [31].
stand human brain evolution. An impressive recent example of comparative psychology in this
When investigating human evolution, one possibility is to start context is a study in which proactive prosociality, a key compo-
from genetic changes and work one’s way up to phenotypes. nent of human cooperation, has been measured in standardized
Such a ‘genotype-first’ approach has been suggested to be behavioural experiments across 15 primate species [32]. It was
feasible for functionally studying the few thousand fixed differ- found that allomaternal care is the best predictor of interspecific
ences that have occurred since modern humans split from Nean- variation, supporting the notion that the emergence of coopera-
derthals and Denisovans [3]. Another approach is to start from a tive breeding is an important life-history trait for human brain
phenotype of interest and work down to the cellular, molecular evolution.
and genetic changes involved. Both approaches have been use- While the genetics of such phenotypes have not been directly
ful and need to merge at some point anyway. For the purpose of explored in humans or other primates, it has been claimed that
this review, I start with phenotypes, such as social cognition, many of these social cognitive skills (including language) are
vocal learning and brain size, and use the framework sketched impaired in children with autism [33]. Hence, genes and path-
above to review relevant recent findings that might help to under- ways associated with autism risk could be candidates for the
stand the molecular basis of their evolution. evolution of human social cognition. While many genetic variants
have been associated with autism spectrum disorders [34], it
Cognitive Traits is unclear whether they have evolved differently on the human
Many cognitive features have been postulated to be unique to lineage. A recent study investigated the protein evolution of
humans, but only more recently have cognitive phenotypes genes associated with different psychiatric disorders across
been quantified in a comparative context [26,27]. This has led primates and mammals [35]. While there was ample evidence
to important insights, for instance that the ability to understand for proteins that evolve under positive selection in the human
others’ inner states and intentions, sometimes referred to as lineage, positive selection or rate of protein evolution was not
‘theory of mind’, is not as unique to humans as initially thought particularly high in the human lineage for either the autism
[28]. Of intense current interest is that humans have changed spectrum disorder-associated genes or any other class of

Current Biology 26, R1109–R1117, October 24, 2016 R1111


Current Biology

Review

Box 2. Phenotypic changes to the human lineage.


and how they could have evolved. On one end of the spectrum is
the view that language and syntax emerged from the evolution of
Naturally, the vast phenotypic space of human evolution has more general cognitive specializations, such as shared intention-
been investigated much less systematically than the genome, ality and sequence learning [40,41]. At the other end of the spec-
but a large number of putative changes on all different pheno- trum is the view that a dedicated language module evolved
typic levels has been collected (https://carta.anthropogeny. rapidly, potentially even caused by a single mutation [42]. Lin-
org/content/about-moca [120]). Of particular relevance are guistic evidence indicates that such a module could have al-
increasing amounts of different neuroimaging data in chim- lowed the hierarchical syntactic structure of language by a single
panzees and rhesus macaques [121,122], neuroanatomical repeatable operation, called ‘merge’, that assembles two syn-
differences in apes and primates [123] and neuron numbers tactic elements a and b to form {a, b} [42]. However, independent
in brains across many species [65]. On the molecular level, of the neurobiological and genetic plausibility of such a module
important datasets include measurements of protein, mRNA emerging suddenly, a phenotype such as the operation ‘merge’
and chromatin signatures in human and chimpanzee lympho- needs to be amenable to empirical investigations at least within
blastoid cell lines [124,125] and in dozens of human, chim- humans before its genetic, molecular or cellular basis can at all
panzee and rhesus brain samples spanning a range of epige- be investigated.
netic features, including profiling of lipids and metabolites In contrast to the cognitive aspects, the evolution of the main
[126–128]. The availability of individuals or tissue samples is modality of language, speech, can be studied much better in
a major limiting factor for broader and faster phenotyping a genetic and comparative framework [38]. Speech is a highly
(although much better than the limits of accessible pheno- elaborate form of vocal production learning, i.e. the learning of
types from fossil data, which is not covered here). This is espe- vocal motor movements based on auditory inputs. In contrast
cially true for obtaining comparable samples during prenatal to communication based on innate vocalizations, vocal learning
development, which are for practical and ethical reasons is at best very limited in non-human primates, including chim-
essentially impossible to obtain from chimpanzees and panzees [43–45]. However, vocal learning is found in several
many other primates [96]. other mammalian lineages, including seals, dolphins, elephants
and bats [44], and also evolved independently in the lineages
leading to songbirds, parrots and hummingbirds [46,47]. Two
disease-associated genes. A similar study recently investigated neuronal pathways, best described for songbirds and parrots,
the protein evolution of ten genes associated with language and are crucial for vocal learning in birds: first, the ‘vocal production
reading impairments in primates, mammals and birds [36]. While pathway’, involving a direct connection from cortical motor areas
they find evidence for positive selection or accelerated evolution to the motor neurons controlling vocalizations; and second, the
in the human lineage, including for the well-studied gene FOXP2 ‘anterior-forebrain pathway’, a cortico-striatal loop specialized
and for the dyslexia associated gene KIAA0319, positive selec- for vocal learning. In humans, a direct projection from the motor
tion is also found outside humans. Another recent study found cortex to the laryngeal motor neurons is necessary to control
that regulatory elements conserved in non-human primates vocal production, while monkeys and apes lack such a direct
and evolving faster in the human lineage are enriched in schizo- projection [48]. Similar evidence for a cortico-striatal specializa-
phrenia-associated genes [37]. This is a promising approach to tion in human speech is lacking, but similarities in transcriptional
link disease-associated loci to human regulatory evolution. How- profiles in birds and humans [49] as well as additional indications
ever, it is currently unclear to what extend this signature is indeed [50,51] suggest that humans might also have evolved a cortico-
human-specific, as accelerated evolution of regulatory elements striatal loop specialized for vocal learning.
in other primates has not been investigated. One can expect that In the search for the genetic basis that gave rise to such adap-
the genetic and molecular mechanisms underlying cognitive tations, FOXP2 is a strong candidate gene. Humans that are
traits and diseases, such as autism, will be much better known heterozygous for a non-functional FOXP2 allele suffer from an
in the near future due to the increasing efforts in human genetics impairment that especially affects their speech and language
and genomics [9]. To identify putative genetic changes that development [52–55]. Analyses of the evolution of FOXP2
could have been relevant for the evolution of human social cogni- revealed two amino acid substitutions, which became fixed in
tion, it will be crucial to use all available information from the human lineage after its separation from the chimpanzee. It
comparative data, including traits such as prosociality. If such was suggested that these two substitutions underwent positive
genotype–phenotype correlations across species reveal prom- selection, potentially due to their effect on speech and language
ising candidates, the validity of phenotypic assays in mice or [56,57]. Subsequent analyses, including several mammalian
cellular systems will probably be the decisive limitation to study species, suggested that two amino acid changes are indeed
further the genetic basis of cognitive evolution. more than expected given the conservation of FOXP2 [36,50].
To test the functional consequences of these two substitutions,
Language and Vocal Learning a mouse model carrying these substitutions in the endogenous
Language is a highly complex behavior that is certainly unique to FOXP2 gene was generated [58]. These mice showed spe-
our species and has emerged after the split from chimpanzees cific alterations in the cortico-basal ganglia circuit, including
[38]. It is undisputed that the evolution of language has a genetic changes in dopamine levels, striatal synaptic plasticity, dendrite
basis and that it is helpful to distinguish different physiological, morphology and gene expression [50,58–60]. Recent evidence
neurological and cognitive aspects of this complex trait [38,39]. indicates that some of these alterations could be caused by an
What is disputed intensively is which of these aspects are crucial increased activity of this humanized FOXP2, leading to increased

R1112 Current Biology 26, R1109–R1117, October 24, 2016


Current Biology

Review
Figure 2. Framework for studying the
molecular basis of human brain evolution.
Studying the molecular basis of human brain
Genome annotations evolution can be regarded as quantitative genetics
Comparative approach without the possibility to cross and genetically
GWAS Genotype–phenotype
Mendelian traits manipulate the investigated species. Hence,
Phenotype –phenotype
additional information is required to obtain plau-
Hypothesis on the genetic sible hypotheses. This includes human variation,
basis of an evolved brain variation in model organisms, variation across
phenoytpe species and evidence for positive selection
or accelerated evolution of genetic elements.
Genome annotations Positive selection Hypotheses usually need to be further tested
QTLs Accelerated evolution in experimental systems amenable to genetic
Experimental testing
knockouts modification such as mice or induced pluripotent
Human allele Ancestral allele stem cells.

across birds and mammals and their —


Current Biology much less measured — connectivity is
a good predictor of ‘intelligence’ [70].
Therefore, the pure size of our brain with
suppression of the synapse suppressor gene Mef2c and hence its 86 billion neurons [67] and probably 100,000 km of connec-
to increased synaptogenesis in the striatum [61]. On a behavioral tions [71] might explain a substantial fraction of our cognitive
level, pups carrying the humanized FOXP2 showed slight but sig- skills. However, neurons and neuronal activity are energetically
nificant differences in the pitch of their ultrasonic vocalisations, very expensive, using up to 60% of the resting metabolic rate
but as this was a transient phenotype not found in adults and in human children, and energy uptake needs to be constant
as mice do not need a cortex to vocalize normally [62], the especially during brain development [72–74]. Hence, explaining
validity of this phenotype for modeling vocal learning evolution how humans can afford such a large brain and still reproduce
is questionable [63]. However, mice carrying the humanized more often than smaller-brained apes, such as orangutans, is
FOXP2 allele learned stimulus-response associations faster, a crucial part of understanding human brain evolution. The costs
showing that the two amino acid changes do affect striatum- and benefits of large brains have only recently been analyzed in a
dependent learning and that such behavioral assays might be larger multivariate comparative framework and the results indi-
more informative for studying the evolution of human vocal cate that an energetic perspective is crucial to understanding
learning [60]. Importantly, the prediction emerging from these brain evolution, as lifespan, diet and neonate mass already
studies — that FOXP2 affects dopamine-dependent learning explain almost 80% of the variation in relative brain size in non-
processes in specific regions of the striatum [60] — was recently human primates [75]. Important aspects as to why humans can
supported by findings that dopaminergic innervation changed afford even larger brains than other primates could be the inten-
specifically in the human lineage and specifically in one striatal sive cooperative breeding found in humans, their higher amount
region [64]. of body fat, cooking and the just recently discovered higher
Additional genes with different degrees of evidence that they amount of total energy spent by humans compared to apes
might be linked to human speech and language evolution have [76–78]. It will be interesting to consider such variables when
been found [36,43,49], but also given its history of intense inves- analysing possible correlations at the molecular and genetic
tigation, FOXP2 is probably still the role model of the possibilities levels, as soon as genome sequences of more primates become
and difficulties of investigating the genetic basis of the evolution available [21].
of a complex human brain phenotype [36]. Independent of such ecological and physiological factors and
their underlying molecular processes that allowed humans to
Brain Size afford large brains, the genetic, molecular and cellular mecha-
Maybe the most obvious, best measurable and most studied nisms that lead to the development of large brains are of intense
aspect of human brain evolution is its increase in size. Only current interest. It is beyond the scope of this review to summa-
recently have data been collected to systematically link brain rize the field, but the crucial point is that the neural progenitor
size to the total number of neurons and non-neuronal cells cells that need to expand in order to build a larger brain have
across primates, mammals and birds [65]. It could be shown recently been much better characterized in mice, ferrets, mon-
that in this respect the human brain, including the cognitively keys and humans [79,80]. For example, recent experimental
important prefrontal cortex [66], has as many cells as expected [81–84] and theoretical [85] evidence now strongly suggests
for a primate brain of this size [67]. The human brain is thus a that an expansion of so-called ‘basal progenitor cells’ is suffi-
scaled-up version of a primate brain. The scaling of brain size cient to cause an increase in cortical size and folding that echoes
and number of neurons can be different in other species such changes seen during evolutionary increases in cortical size and
as rodents and, notably, it was found that elephants possess folding. As the cellular and developmental basis of brain size
fewer cortical neurons than humans [68] and that birds, despite evolution is relatively well understood and as brain size and
their small brain size, have neuron numbers similar to primates correlated variables such as body size, diet and lifespan are fairly
[69]. So it is plausible that the total number of cortical neurons well measured across primates and mammals, the evolution of

Current Biology 26, R1109–R1117, October 24, 2016 R1113


Current Biology

Review

this phenotype is especially amenable to a comparative extent similar to the limitations that exist for studying human phe-
approach [21] and other investigations in the sketched frame- notypes, for which induced pluripotent stem cells (iPSCs) bear
work (Figure 2). This includes experimental approaches to hu- great promise to study developmental processes also in com-
man brain size evolution that have gained momentum recently. plex brain organoids [102]. As iPSCs can be generated from pri-
For example, one study [86] found that the human version of mates, including apes [103–105], it is evident that they could be a
an enhancer regulating FZD8, a receptor of the Wnt pathway, powerful tool to better understand human brain evolution on a
leads to a faster cell cycle in neural progenitors and an increased molecular and cellular level [1,96,97]. Indeed, first comparisons
brain size in transgenic mice, while chimpanzee FZD8 does of differentiated iPSCs from humans and chimpanzees already
not have such an effect. Moreover, another analysis showed indicate that this is the case [92,95]. As molecular phenotypes
that a human-specific gene duplication encoding Rho GTPase such as RNA expression levels can now be measured at the
Activating Protein 11B (ARHGAP11B) increases the proliferation single-cell resolution [106], this will be a powerful system to anal-
of basal progenitors when expressed in mice [87]. While these yse the evolution of molecular phenotypes and differentiation
approaches are crucial first findings, they are certainly not processes in physiologically relevant cell types.
the only genetic changes responsible for the evolution of Furthermore, genetic manipulation of primate iPSCs will allow
human brain size [88]; thousands of enhancers and promoters researchers to probe the genetic basis of brain phenotypes that
have also gained activity during human corticogenesis com- can be studied at the cellular level. Massively parallel reporter as-
pared to rhesus macaques [89]. Furthermore, it is important to says now allow us to screen the activity of thousands of en-
consider that epigenetic signatures and regulatory activity of hancers [107–109] and genome editing technologies [110] will
particular elements can be different, although the expression enable modification of specific genomic regions in different pri-
of the regulated gene or network is unchanged. This turn-over mate iPSCs. Together with better theoretical models [93] and
of regulatory evolution is substantial as comparisons of binding computational tools [94], this will leverage comparative func-
sites between human and mice [90,91] and among mammalian tional genomics to better understand molecular evolution in gen-
livers [18] have shown. Also when comparing enhancer signa- eral and human brain evolution in particular.
tures and expression levels in human and chimpanzee neural While iPSCs hold great promise to study human brain evolu-
crest cells, many genes with an increased enhancer signature tion on a molecular and cellular level, they will not abolish the
in one species are not upregulated on the mRNA level in the need to study genetic changes potentially relevant for human
same species [92]. More theoretical, computational and experi- brain evolution in model organisms such as mice. Genome
mental work, including modeling brain size development in editing will also make this process decisively faster [96]. While
organoids, will be needed to better understand the evolution mouse models will certainly be able to model some aspects of
of gene regulation in general and during development in complex behavioral phenotypes, they will also certainly have
particular [93–97]. limits [25], eventually restricting which aspects of human brain
evolution we will be able to reconstruct.
Prospects for Studying Human Brain Evolution in a
Comparative Framework REFERENCES
As pointed out above, there has been substantial progress in
describing human brain evolution in recent years. Nevertheless, 1. Enard, W. (2012). Functional primate genomics-leveraging the medical
potential. J. Mol. Med. 90, 471–480.
we are still far from understanding how the evolution of some
hallmark changes is rooted in genetic, molecular and cellular 2. Varki, A. (2012). Nothing in medicine makes sense, except in the light of
processes. The framework of how this could be achieved evolution. J. Mol. Med. (Berl) 90, 481–494.
(Figure 2) is also helpful to discuss in which areas substantial €a
3. Pa €bo, S. (2014). The human condition-a molecular approach. Cell 157,
progress could be expected in the near future. This certainly 216–226.
includes increasing knowledge on genotype–phenotype
4. Rossiianov, K. (2002). Beyond species: Il’ya Ivanov and his experiments
associations within humans, including rare mutations, as it is ex- on cross-breeding humans with anthropoid apes. Sci. Context 15,
pected that millions of human genomes will be sequenced in the 277–316.
next ten years [98]. Given a mutation rate of 1–2 3 10–8 per base 5. Mackay, T.F.C., Stone, E.A., and Ayroles, J.F. (2009). The genetics
pair and generation [99], nucleotide substitutions that occurred of quantitative traits: challenges and prospects. Nat. Rev. Genet. 10,
in the human lineage will be mutated back to the chimpanzee 565–577.
state once in every 1–2 3 108 human chromosomes that are 6. Mackay, T.F., Richards, S., Stone, E.A., Barbadilla, A., Ayroles, J.F., Zhu,
born [100]. D., Casillas, S., Han, Y., Magwire, M.M., Cridland, J.M., et al. (2012).
The Drosophila melanogaster genetic reference panel. Nature 482,
Cheaper and better sequencing technologies will also allow us 173–178.
to obtain high-quality genomes for essentially all primates and
mammals [101]. This will be important not only to describe pat- 7. Kang, H.J., Kawasawa, Y.I., Cheng, F., Zhu, Y., Xu, X., Li, M., Sousa,
A.M., Pletikos, M., Meyer, K.A., Sedmak, G., et al. (2011). Spatio-tempo-
terns of positive selection on the human lineage, but also to ral transcriptome of the human brain. Nature 478, 483–489.
analyze these patterns in a comparative framework to identify
8. Roadmap Epigenomics, C., Kundaje, A., Meuleman, W., Ernst, J., Bi-
positive selection that co-varies with brain phenotypes of inter-
lenky, M., Yen, A., Heravi-Moussavi, A., Kheradpour, P., Zhang, Z.,
est, such as brain size or vocal learning. However, to measure Wang, J., et al. (2015). Integrative analysis of 111 reference human epige-
associated molecular and cellular phenotypes in tissues across nomes. Nature 518, 317–330.
many of these species, especially during development, will 9. Geschwind, D.H., and Flint, J. (2015). Genetics and genomics of psychi-
remain practically and ethically very difficult. This is to some atric disease. Science 349, 1489–1494.

R1114 Current Biology 26, R1109–R1117, October 24, 2016


Current Biology

Review
10. Fisher, S.E. (2016). Evolution of language: Lessons from the genome. 33. Tomasello, M., Carpenter, M., Call, J., Behne, T., and Moll, H. (2005).
Psychon. Bull. Rev., epub ahead of print. Understanding and sharing intentions: the origins of cultural cognition.
Behav. Brain Sci. 28, 675–691, discussion 691–735.
11. Hibar, D.P., Stein, J.L., Renteria, M.E., Arias-Vasquez, A., Desrivieres, S.,
Jahanshad, N., Toro, R., Wittfeld, K., Abramovic, L., Andersson, M., et al. 34. Robinson, E.B., St Pourcain, B., Anttila, V., Kosmicki, J.A., Bulik-Sullivan,
(2015). Common genetic variants influence human subcortical brain B., Grove, J., Maller, J., Samocha, K.E., Sanders, S.J., Ripke, S., et al.
structures. Nature 520, 224–229. (2016). Genetic risk for autism spectrum disorders and neuropsychiatric
variation in the general population. Nat. Genet. 48, 552–555.
12. Yue, F., Cheng, Y., Breschi, A., Vierstra, J., Wu, W., Ryba, T., Sandstrom,
R., Ma, Z., Davis, C., Pope, B.D., et al. (2014). A comparative encyclo- 35. Ogawa, L.M., and Vallender, E.J. (2014). Evolutionary conservation
pedia of DNA elements in the mouse genome. Nature 515, 355–364. in genes underlying human psychiatric disorders. Front. Hum. Neurosci.
8, 283.
13. Vitti, J.J., Grossman, S.R., and Sabeti, P.C. (2013). Detecting natural se-
lection in genomic data. Annu. Rev. Genet. 47, 97–120. 36. Mozzi, A., Forni, D., Clerici, M., Pozzoli, U., Mascheretti, S., Guerini, F.R.,
Riva, S., Bresolin, N., Cagliani, R., and Sironi, M. (2016). The evolutionary
14. Mallick, S., Gnerre, S., Muller, P., and Reich, D. (2009). The difficulty of history of genes involved in spoken and written language: beyond
avoiding false positives in genome scans for natural selection. Genome FOXP2. Sci. Rep. 6, 22157.
Res. 19, 922–933.
37. Xu, K., Schadt, E.E., Pollard, K.S., Roussos, P., and Dudley, J.T. (2015).
15. Gittelman, R.M., Hun, E., Ay, F., Madeoy, J., Pennacchio, L., Noble, W.S., Genomic and network patterns of schizophrenia genetic variation in hu-
Hawkins, R.D., and Akey, J.M. (2015). Comprehensive identification and man evolutionary accelerated regions. Mol. Biol. Evol. 32, 1148–1160.
analysis of human accelerated regulatory DNA. Genome Res. 25, 1245–
1255. 38. Fitch, W.T. (2010). The Evolution of Language (Cambridge University
Press).
16. Enard, D., Cai, L., Gwennap, C., and Petrov, D.A. (2016). Viruses are a
dominant driver of protein adaptation in mammals. eLife 5, http://dx. 39. Hauser, M.D., Chomsky, N., and Fitch, W.T. (2002). The faculty of lan-
doi.org/10.7554/eLife.12469. guage: what is it, who has it, and how did it evolve? Science 298,
1569–1579.
17. Ivankov, D.N., Finkelstein, A.V., and Kondrashov, F.A. (2014). A structural
perspective of compensatory evolution. Curr. Opin. Struct. Biol. 26, 40. Tomasello, M. (2003). Constructing a language: A usage-based
104–112. approach to child language acquisition (Cambridge, MA: Harvard Univer-
sity Press).
18. Villar, D., Berthelot, C., Aldridge, S., Rayner, T.F., Lukk, M., Pignatelli, M.,
Park, T.J., Deaville, R., Erichsen, J.T., Jasinska, A.J., et al. (2015). 41. Christiansen, M.H., and Chater, N. (2015). The language faculty that
Enhancer evolution across 20 mammalian species. Cell 160, 554–566. wasn’t: a usage-based account of natural language recursion. Front.
Psychol. 6, 1182.
19. Felsenstein, J. (1985). Phylogenies and the comparative method. Am.
Nat. 125, 1–15. 42. Bolhuis, J.J., Tattersall, I., Chomsky, N., and Berwick, R.C. (2014). How
could language have evolved? PLoS Biol. 12, e1001934.
20. Nunn, C.L. (2011). The Comparative Approach in Evolutionary Anthropol-
ogy and Biology (University of Chicago Press).
43. Konopka, G., and Roberts, T.F. (2016). Insights into the neural and
genetic basis of vocal communication. Cell 164, 1269–1276.
21. Enard, W. (2014). Comparative genomics of brain size evolution. Front.
Hum. Neurosci. 8, 345.
44. Brainard, M.S., and Fitch, W.T. (2014). Editorial overview: communica-
tion and language: animal communication and human language. Curr.
22. Jacobs, G.H. (2009). Evolution of colour vision in mammals. Philos.
Opin. Neurobiol. 28, v–viii.
Trans. R. Soc. Lond. B. Biol. Sci. 364, 2957–2967.

23. Carroll, S.B. (2008). Evo-devo and an expanding evolutionary synthesis: 45. Fischer, J. (2016). Primate vocal production and the riddle of language
a genetic theory of morphological evolution. Cell 134, 25–36. evolution. Psychon. Bull. Rev., epub ahead of print.

24. Sato, K., Oiwa, R., Kumita, W., Henry, R., Sakuma, T., Ito, R., Nozu, R., 46. Jarvis, E.D. (2004). Learned birdsong and the neurobiology of human lan-
Inoue, T., Katano, I., Sato, K., et al. (2016). Generation of a nonhuman guage. Ann. N. Y. Acad. Sci. 1016, 749–777.
primate model of severe combined immunodeficiency using highly effi-
cient genome editing. Cell Stem Cell 19, 127–138. 47. Chakraborty, M., and Jarvis, E.D. (2015). Brain evolution by brain
pathway duplication. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370.
25. Enard, W. (2014). Mouse models of human evolution. Curr. Opin. Genet.
Dev. 29C, 75–80. 48. Simonyan, K. (2014). The laryngeal motor cortex: its organization and
connectivity. Curr. Opin. Neurobiol. 28, 15–21.
26. MacLean, E.L. (2016). Unraveling the evolution of uniquely human cogni-
tion. Proc. Natl. Acad. Sci. USA 113, 6348–6354. 49. Pfenning, A.R., Hara, E., Whitney, O., Rivas, M.V., Wang, R., Roulhac,
P.L., Howard, J.T., Wirthlin, M., Lovell, P.V., Ganapathy, G., et al.
27. Van Schaik, C.P. (2016). The Primate Origins of Human Nature (John Wi- (2014). Convergent transcriptional specializations in the brains of
ley & Sons). humans and song-learning birds. Science 346, 1256846.

28. Call, J., and Tomasello, M. (2008). Does the chimpanzee have a theory of 50. Enard, W. (2011). FOXP2 and the role of cortico-basal ganglia circuits in
mind? 30 years later. Trends Cogn. Sci. 12, 187–192. speech and language evolution. Curr. Opin. Neurobiol. 21, 415–424.

29. Herrmann, E., Call, J., Hernandez-Lloreda, M.V., Hare, B., and Toma- 51. Ackermann, H., Hage, S.R., and Ziegler, W. (2014). Brain mechanisms of
sello, M. (2007). Humans have evolved specialized skills of social cogni- acoustic communication in humans and nonhuman primates: an evolu-
tion: the cultural intelligence hypothesis. Science 317, 1360–1366. tionary perspective. Behav. Brain Sci. 37, 529–546.

30. Tomasello, M., and Vaish, A. (2013). Origins of human cooperation and 52. Lai, C.S., Fisher, S.E., Hurst, J.A., Vargha-Khadem, F., and Monaco, A.P.
morality. Annu. Rev. Psychol. 64, 231–255. (2001). A forkhead-domain gene is mutated in a severe speech and
language disorder. Nature 413, 519–523.
31. Tennie, C., Call, J., and Tomasello, M. (2009). Ratcheting up the ratchet:
on the evolution of cumulative culture. Philos. Trans. R. Soc. Lond. B. 53. Graham, S.A., and Fisher, S.E. (2015). Understanding language from a
Biol. Sci. 364, 2405–2415. genomic perspective. Annu. Rev. Genet. 49, 131–160.

32. Burkart, J.M., Allon, O., Amici, F., Fichtel, C., Finkenwirth, C., Heschl, A., 54. Vargha-Khadem, F., Gadian, D.G., Copp, A., and Mishkin, M. (2005).
Huber, J., Isler, K., Kosonen, Z.K., Martins, E., et al. (2014). The evolu- FOXP2 and the neuroanatomy of speech and language. Nat. Rev. Neuro-
tionary origin of human hyper-cooperation. Nat. Commun. 5, 4747. sci. 6, 131–138.

Current Biology 26, R1109–R1117, October 24, 2016 R1115


Current Biology

Review
55. Watkins, K. (2011). Developmental disorders of speech and language: 74. Lukas, W.D., and Campbell, B.C. (2000). Evolutionary and ecological as-
from genes to brain structure and function. Prog. Brain Res. 189, pects of early brain malnutrition in humans. Human Nat. 11, 1–26.
225–238.
75. Isler, K., and Van Schaik, C.P. (2014). How humans evolved large brains:
56. Enard, W., Przeworski, M., Fisher, S.E., Lai, C.S., Wiebe, V., Kitano, T., comparative evidence. Evol. Anthropol. 23, 65–75.
Monaco, A.P., and Paabo, S. (2002). Molecular evolution of FOXP2, a
gene involved in speech and language. Nature 418, 869–872. 76. Navarrete, A., van Schaik, C.P., and Isler, K. (2011). Energetics and the
evolution of human brain size. Nature 480, 91–93.
57. Zhang, J., Webb, D.M., and Podlaha, O. (2002). Accelerated protein evo-
lution and origins of human-specific features: Foxp2 as an example. Ge- 77. Carmody, R.N., Weintraub, G.S., and Wrangham, R.W. (2011). Energetic
netics 162, 1825–1835. consequences of thermal and nonthermal food processing. Proc. Natl.
Acad. Sci. USA 108, 19199–19203.
58. Enard, W., Gehre, S., Hammerschmidt, K., Holter, S.M., Blass, T., Somel,
M., Bruckner, M.K., Schreiweis, C., Winter, C., Sohr, R., et al. (2009). 78. Pontzer, H., Brown, M.H., Raichlen, D.A., Dunsworth, H., Hare, B.,
A humanized version of Foxp2 affects cortico-basal ganglia circuits in Walker, K., Luke, A., Dugas, L.R., Durazo-Arvizu, R., Schoeller, D.,
mice. Cell 137, 961–971. et al. (2016). Metabolic acceleration and the evolution of human brain
size and life history. Nature 533, 390–392.
59. Reimers-Kipping, S., Hevers, W., Paabo, S., and Enard, W. (2011). Hu-
manized Foxp2 specifically affects cortico-basal ganglia circuits. Neuro- 79. Fernández, V., Llinares-Benadero, C., and Borrell, V. (2016). Cerebral
science 175, 75–84.
cortex expansion and folding: what have we learned? EMBO. J. 35,
1021–1044.
60. Schreiweis, C., Bornschein, U., Burguiere, E., Kerimoglu, C., Schreiter,
S., Dannemann, M., Goyal, S., Rea, E., French, C.A., Puliyadi, R., et al.
80. Dehay, C., Kennedy, H., and Kosik, K.S. (2015). The outer subventric-
(2014). Humanized Foxp2 accelerates learning by enhancing transitions
ular zone and primate-specific cortical complexification. Neuron 85,
from declarative to procedural performance. Proc. Natl. Acad. Sci. USA
683–694.
111, 14253–14258.

61. Chen, Y.C., Kuo, H.Y., Bornschein, U., Takahashi, H., Chen, S.Y., Lu, 81. Nonaka-Kinoshita, M., Reillo, I., Artegiani, B., Martinez-Martinez, M.A.,
K.M., Yang, H.Y., Chen, G.M., Lin, J.R., Lee, Y.H., et al. (2016). Foxp2 Nelson, M., Borrell, V., and Calegari, F. (2013). Regulation of cerebral
controls synaptic wiring of corticostriatal circuits and vocal communica- cortex size and folding by expansion of basal progenitors. EMBO. J.
tion by opposing Mef2c. Nat. Neurosci. http://dx.doi.org/10.1038/nn. 32, 1817–1828.
4380.
82. Stahl, R., Walcher, T., De Juan Romero, C., Pilz, G.A., Cappello, S., Irm-
62. Hammerschmidt, K., Whelan, G., Eichele, G., and Fischer, J. (2015). Mice ler, M., Sanz-Aquela, J.M., Beckers, J., Blum, R., Borrell, V., et al. (2013).
lacking the cerebral cortex develop normal song: insights into the foun- Trnp1 regulates expansion and folding of the mammalian cerebral cortex
dations of vocal learning. Sci. Rep. 5, 8808. by control of radial glial fate. Cell 153, 535–549.

63. Hammerschmidt, K., Schreiweis, C., Minge, C., Paabo, S., Fischer, J., 83. Wong, F.K., Fei, J.F., Mora-Bermudez, F., Taverna, E., Haffner, C., Fu, J.,
and Enard, W. (2015). A humanized version of Foxp2 does not affect ul- Anastassiadis, K., Stewart, A.F., and Huttner, W.B. (2015). Sustained
trasonic vocalization in adult mice. Genes. Brain Behav. 14, 583–590. Pax6 expression generates primate-like basal radial glia in developing
mouse neocortex. PLoS Biol. 13, e1002217.
64. Raghanti, M.A., Edler, M.K., Stephenson, A.R., Wilson, L.J., Hopkins,
W.D., Ely, J.J., Erwin, J.M., Jacobs, B., Hof, P.R., and Sherwood, C.C. 84. Wang, L., Hou, S., and Han, Y.G. (2016). Hedgehog signaling promotes
(2016). Human-specific increase of dopaminergic innervation in a striatal basal progenitor expansion and the growth and folding of the neocortex.
region associated with speech and language: A comparative analysis of Nat. Neurosci. 19, 888–896.
the primate basal ganglia. J. Comp. Neurol. 524, 2117–2129.
85. Lewitus, E., Kelava, I., Kalinka, A.T., Tomancak, P., and Huttner, W.B.
65. Herculano-Houzel, S., Catania, K., Manger, P.R., and Kaas, J.H. (2015). (2014). An adaptive threshold in mammalian neocortical evolution.
Mammalian brains are made of these: a dataset of the numbers and den- PLoS Biol. 12, e1002000.
sities of neuronal and nonneuronal cells in the brain of glires, primates,
scandentia, eulipotyphlans, afrotherians and artiodactyls, and their rela- 86. Boyd, J.L., Skove, S.L., Rouanet, J.P., Pilaz, L.J., Bepler, T., Gordan, R.,
tionship with body mass. Brain Behav. Evol. 86, 145–163. Wray, G.A., and Silver, D.L. (2015). Human-chimpanzee differences in a
FZD8 enhancer alter cell-cycle dynamics in the developing neocortex.
66. Gabi, M., Neves, K., Masseron, C., Ribeiro, P.F., Ventura-Antunes, L., Curr. Biol. 25, 772–779.
Torres, L., Mota, B., Kaas, J.H., and Herculano-Houzel, S. (2016). No
relative expansion of the number of prefrontal neurons in primate and hu- 87. Florio, M., Albert, M., Taverna, E., Namba, T., Brandl, H., Lewitus, E.,
man evolution. Proc. Natl. Acad. Sci. USA 113, 9617–9622. Haffner, C., Sykes, A., Wong, F.K., Peters, J., et al. (2015). Human-spe-
cific gene ARHGAP11B promotes basal progenitor amplification and
67. Herculano-Houzel, S. (2012). The remarkable, yet not extraordinary, hu- neocortex expansion. Science 347, 1465–1470.
man brain as a scaled-up primate brain and its associated cost. Proc.
Natl. Acad. Sci. USA 109 (Suppl 1 ), 10661–10668. 88. Franchini, L.F., and Pollard, K.S. (2015). Can a few non-coding mutations
make a human brain? Bioessays 37, 1054–1061.
68. Herculano-Houzel, S., Avelino-de-Souza, K., Neves, K., Porfirio, J., Mes-
seder, D., Mattos Feijo, L., Maldonado, J., and Manger, P.R. (2014). The 89. Reilly, S.K., Yin, J., Ayoub, A.E., Emera, D., Leng, J., Cotney, J., Sarro, R.,
elephant brain in numbers. Front. Neuroanat. 8, 46. Rakic, P., and Noonan, J.P. (2015). Evolutionary genomics. Evolutionary
an, R.K., Portes, M., Fitch, W.T., Hercu- changes in promoter and enhancer activity during human corticogenesis.
69. Olkowicz, S., Kocourek, M., Luc
mec, P. (2016). Birds have primate-like numbers Science 347, 1155–1159.
lano-Houzel, S., and Ne
of neurons in the forebrain. Proc. Natl. Acad. Sci. USA 113, 7255–7260.
90. Stergachis, A.B., Neph, S., Sandstrom, R., Haugen, E., Reynolds, A.P.,
70. Dicke, U., and Roth, G. (2016). Neuronal factors determining high intelli- Zhang, M., Byron, R., Canfield, T., Stelhing-Sun, S., Lee, K., et al.
gence. Philos. Trans. R. Soc. Lond. B Biol. Sci. 371, 20150180. (2014). Conservation of trans-acting circuitry during mammalian regula-
tory evolution. Nature 515, 365–370.
71. Hofman, M.A. (2014). Evolution of the human brain: when bigger is better.
Front. Neuroanat 8, 15. 91. Vierstra, J., Rynes, E., Sandstrom, R., Zhang, M., Canfield, T., Hansen,
R.S., Stehling-Sun, S., Sabo, P.J., Byron, R., Humbert, R., et al. (2014).
72. Niven, J.E., and Laughlin, S.B. (2008). Energy limitation as a selective pres- Mouse regulatory DNA landscapes reveal global principles of cis-regula-
sure on the evolution of sensory systems. J. Exp. Biol. 211, 1792–1804. tory evolution. Science 346, 1007–1012.

73. Kuzawa, C.W., Chugani, H.T., Grossman, L.I., Lipovich, L., Muzik, O., 92. Prescott, S.L., Srinivasan, R., Marchetto, M.C., Grishina, I., Narvaiza, I.,
Hof, P.R., Wildman, D.E., Sherwood, C.C., Leonard, W.R., and Lange, Selleri, L., Gage, F.H., Swigut, T., and Wysocka, J. (2015). Enhancer
N. (2014). Metabolic costs and evolutionary implications of human brain divergence and cis-regulatory evolution in the human and chimp neural
development. Proc. Natl. Acad. Sci. USA 111, 13010–13015. crest. Cell 163, 68–83.

R1116 Current Biology 26, R1109–R1117, October 24, 2016


Current Biology

Review
93. Tugrul, M., Paixao, T., Barton, N.H., and Tkacik, G. (2015). Dynamics 112. Kuhlwilm, M., Gronau, I., Hubisz, M.J., de Filippo, C., Prado-Martinez, J.,
of transcription factor binding site evolution. PLoS Genet. 11, e1005639. Kircher, M., Fu, Q., Burbano, H.A., Lalueza-Fox, C., de la Rasilla, M., et al.
(2016). Ancient gene flow from early modern humans into Eastern Nean-
94. Thompson, D., Regev, A., and Roy, S. (2015). Comparative analysis of derthals. Nature 530, 429–433.
gene regulatory networks: from network reconstruction to evolution.
Annu. Rev. Cell. Dev. Biol. 31, 399–428. 113. Mikkelsen, T., Hillier, L., Eichler, E., Zody, M., Jaffe, D., Yang, S., Enard,
W., Hellmann, I., Lindblad-Toh, K., Altheide, T., et al. (2005). Initial
95. Otani, T., Marchetto, M.C., Gage, F.H., Simons, B.D., and Livesey, F.J. sequence of the chimpanzee genome and comparison with the human
(2016). 2D and 3D stem cell models of primate cortical development genome. Nature 437, 69–87.
identify species-specific differences in progenitor behavior contributing
to brain size. Cell Stem Cell 18, 467–480.
114. Defelipe, J. (2011). The evolution of the brain, the human nature of cortical
96. Reilly, S.K., and Noonan, J.P. (2016). Evolution of gene regulation in hu- circuits, and intellectual creativity. Front. Neuroanat. 5, 29.
mans. Annu. Rev. Genomics Hum. Genet. 17, 45–67.
115. Locke, D.P., Hillier, L.W., Warren, W.C., Worley, K.C., Nazareth, L.V.,
97. Franchini, L.F., and Pollard, K.S. (2015). Genomic approaches to study- Muzny, D.M., Yang, S.P., Wang, Z., Chinwalla, A.T., Minx, P., et al.
ing human-specific developmental traits. Development 142, 3100–3112. (2011). Comparative and demographic analysis of orang-utan genomes.
Nature 469, 529–533.
98. Stephens, Z.D., Lee, S.Y., Faghri, F., Campbell, R.H., Zhai, C., Efron,
M.J., Iyer, R., Schatz, M.C., Sinha, S., and Robinson, G.E. (2015). Big 116. Rhesus Macaque Genome, S., Analysis, C., Gibbs, R.A., Rogers, J., Ka-
data: astronomical or genomical? PLoS Biol. 13, e1002195. tze, M.G., Bumgarner, R., Weinstock, G.M., Mardis, E.R., Remington,
K.A., Strausberg, R.L., et al. (2007). Evolutionary and biomedical insights
99. Shendure, J., and Akey, J.M. (2015). The origins, determinants, and con- from the rhesus macaque genome. Science 316, 222–234.
sequences of human mutations. Science 349, 1478–1483.
117. Rands, C.M., Meader, S., Ponting, C.P., and Lunter, G. (2014). 8.2% of
100. Enard, W., and Paabo, S. (2004). Comparative primate genomics. Annu. the Human genome is constrained: variation in rates of turnover across
Rev. Genomics. Hum. Genet. 5, 351–378. functional element classes in the human lineage. PLoS Genet. 10,
e1004525.
101. Genome 10K Community of Scientists (2009). Genome 10K: a proposal
to obtain whole-genome sequence for 10,000 vertebrate species.
J. Hered 100, 659–674. 118. Sudmant, P.H., Huddleston, J., Catacchio, C.R., Malig, M., Hillier, L.W.,
Baker, C., Mohajeri, K., Kondova, I., Bontrop, R.E., Persengiev, S., et al.
102. Kelava, I., and Lancaster, M.A. (2016). Stem cell models of human brain (2013). Evolution and diversity of copy number variation in the great ape
development. Cell Stem Cell 18, 736–748. lineage. Genome Res. 23, 1373–1382.

103. Gallego Romero, I., Pavlovic, B.J., Hernando-Herraez, I., Zhou, X., Ward, 119. Huddleston, J., and Eichler, E.E. (2016). An incomplete understanding of
M.C., Banovich, N.E., Kagan, C.L., Burnett, J.E., Huang, C.H., Mitrano, human genetic variation. Genetics 202, 1251–1254.
A., et al. (2015). A panel of induced pluripotent stem cells from chimpan-
zees: a resource for comparative functional genomics. eLife 4, e07103. 120. Varki, A., and Altheide, T.K. (2005). Comparing the human and chim-
panzee genomes: searching for needles in a haystack. Genome Res.
104. Marchetto, M.C., Narvaiza, I., Denli, A.M., Benner, C., Lazzarini, T.A., Na- 15, 1746–1758.
thanson, J.L., Paquola, A.C., Desai, K.N., Herai, R.H., Weitzman, M.D.,
et al. (2013). Differential L1 regulation in pluripotent stem cells of humans 121. Rilling, J.K. (2014). Comparative primate neuroimaging: insights into hu-
and apes. Nature 503, 525–529. man brain evolution. Trends Cogn. Sci. 18, 46–55.
105. Wunderlich, S., Kircher, M., Vieth, B., Haase, A., Merkert, S., Beier, J.,
Gohring, G., Glage, S., Schambach, A., Curnow, E.C., et al. (2014). Pri- 122. Rilling, J.K. (2014). Comparative primate neurobiology and the evolution
mate iPS cells as tools for evolutionary analyses. Stem Cell Res. 12, of brain language systems. Curr. Opin. Neurobiol. 28, 10–14.
622–629.
123. Sherwood, C.C., Bauernfeind, A.L., Bianchi, S., Raghanti, M.A., and Hof,
106. Kolodziejczyk, A.A., Kim, J.K., Svensson, V., Marioni, J.C., and Teich- P.R. (2012). Human brain evolution writ large and small. Prog. Brain Res.
mann, S.A. (2015). The technology and biology of single-cell RNA 195, 237–254.
sequencing. Mol. Cell 58, 610–620.
124. Battle, A., Khan, Z., Wang, S.H., Mitrano, A., Ford, M.J., Pritchard, J.K.,
107. Melnikov, A., Murugan, A., Zhang, X., Tesileanu, T., Wang, L., Rogov, P., and Gilad, Y. (2015). Genomic variation. Impact of regulatory variation
Feizi, S., Gnirke, A., Callan, C.G., Jr., Kinney, J.B., et al. (2012). System- from RNA to protein. Science 347, 664–667.
atic dissection and optimization of inducible enhancers in human cells
using a massively parallel reporter assay. Nat. Biotechnol. 30, 271–277. 125. Zhou, X., Cain, C.E., Myrthil, M., Lewellen, N., Michelini, K., Davenport,
E.R., Stephens, M., Pritchard, J.K., and Gilad, Y. (2014). Epigenetic mod-
108. Shen, S.Q., Myers, C.A., Hughes, A.E., Byrne, L.C., Flannery, J.G., and ifications are associated with inter-species gene expression variation in
Corbo, J.C. (2016). Massively parallel cis-regulatory analysis in the primates. Genome. Biol. 15, 547.
mammalian central nervous system. Genome. Res. 26, 238–255.
126. Bozek, K., Wei, Y., Yan, Z., Liu, X., Xiong, J., Sugimoto, M., Tomita, M.,
109. Tewhey, R., Kotliar, D., Park, D.S., Liu, B., Winnicki, S., Reilly, S.K., An-
Paabo, S., Pieszek, R., Sherwood, C.C., et al. (2014). Exceptional evolu-
dersen, K.G., Mikkelsen, T.S., Lander, E.S., Schaffner, S.F., et al.
tionary divergence of human muscle and brain metabolomes parallels
(2016). Direct identification of hundreds of expression-modulating vari-
human cognitive and physical uniqueness. PLoS Biol. 12, e1001871.
ants using a multiplexed reporter assay. Cell 165, 1519–1529.

110. Cong, L., Ran, F.A., Cox, D., Lin, S., Barretto, R., Habib, N., Hsu, P.D., 127. Bozek, K., Wei, Y., Yan, Z., Liu, X., Xiong, J., Sugimoto, M., Tomita, M.,
Wu, X., Jiang, W., Marraffini, L.A., et al. (2013). Multiplex genome engi- Paabo, S., Sherwood, C.C., Hof, P.R., et al. (2015). Organization and
neering using CRISPR/Cas systems. Science 339, 819–823. evolution of brain lipidome revealed by large-scale analysis of human,
chimpanzee, macaque, and mouse tissues. Neuron 85, 695–702.
111. Prado-Martinez, J., Sudmant, P.H., Kidd, J.M., Li, H., Kelley, J.L., Lor-
ente-Galdos, B., Veeramah, K.R., Woerner, A.E., O’Connor, T.D., Sant- 128. Somel, M., Rohlfs, R., and Liu, X. (2014). Transcriptomic insights into
pere, G., et al. (2013). Great ape genetic diversity and population history. human brain evolution: acceleration, neutrality, heterochrony. Curr.
Nature 499, 471–475. Opin. Genet. Dev. 29, 110–119.

Current Biology 26, R1109–R1117, October 24, 2016 R1117

You might also like