Ponting Annurev Constraint
Ponting Annurev Constraint
ANNUAL
REVIEWS Further No Gene in the Genome Makes
Click here for quick links to
Annual Reviews content online,
including:
Sense Except in the Light
• Other articles in this volume
of Evolution
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
71
GG15CH04-Ponting ARI 15 July 2014 11:5
1. INTRODUCTION
Nothing makes sense in biology except in the light of evolution.
effectively distinguish between functional sites and inconsequential sites, such as those involved
in low-occupancy transcription factor interactions (39), and it conflated epiphenomena (such as
random transcription events) with primary causes (biological function). Current experimental
approaches thus are unable to predict comprehensively or accurately how important particular
regulatory elements are to organismal biology. Experimental targeting of mutations in vivo can
reveal deleterious effects, but formally it is still necessary to demonstrate that these have effects in
natural settings, something that current protocols rarely cover.
An alternative approach to predict the biological importance and locations of noncoding func-
tional elements is to recognize the telltale signatures written over time into genomes by the sieving
of mutations by natural selection. This approach has the advantages of being comprehensive (be-
cause the scrutiny of selection extends to any element that is functional in any cell type and at
any developmental stage) and inexpensive (because of the availability of genome sequences and of
cheap sequencing technologies). Selection also will not have regarded as functional any sequence
that, despite being bound by a factor or being transcribed, remains of no biological consequence.
Finally, unlike biochemical assays, evolutionary analyses can predict an element’s biological im-
portance by estimating the strength of selection that acted on its sequence. This approach can
indicate the extent to which fitness has been affected by ancient mutations that landed in single
elements, and more generally in classes of functional elements, and thus can rank elements by
their proposed biological importance.
Here, we review the relative contributions to human biology of different classes of functional el-
ements inferred from comparisons of recently sequenced mammalian and human genomes. Owing
to space limitations, we mostly restrict the discussion to classes of functional elements rather than
individual examples, and we focus on conservation and constraint as opposed to positive selection
and adaptation, which have been reviewed elsewhere (41). As a consequence, the review focuses
on the approximately 10% of the human genome that is under significant selective constraints and
not on the vast majority that evolves neutrally (101, 103). The review is structured around three
counterpoints (Figure 1): (a) selection that acted on either non-protein-coding or protein-coding
sequence; (b) selection that acted on either the genomes of humans (or other mammals) or those
of other, more distantly related metazoans, such as Drosophila (fruit flies); and (c) selection that
was either ancient or more contemporary.
72 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
Weak S E LE C T IO N Strong
Non-protein-
coding genes
a
Protein-
coding genes
Fruit fly
Protein-
b
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
coding Human
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
genes
Ancient
human
c
Contemporary
human
Figure 1
Schematic illustration of the differing extents to which selection has acted on exonic sequences of genes. These genes have been
grouped according to the threefold organizational principle of the review: Comparisons are made between (a) selection acting on
non-protein-coding versus protein-coding genes; (b) selection acting on protein-coding genes in species such as Drosophila (fruit flies)
versus those in humans; and (c) selection acting on human protein-coding genes during ancient versus more contemporary evolution.
On average, selection has been weakest on transcribed non-protein-coding sequence and strongest on protein-coding sequence from
species (such as fruit flies) that have high effective population sizes.
was functional, we would observe it to have changed only very rarely. This is because change
would be mostly deleterious and thus would be negatively selected and less likely to have been
propagated to subsequent generations. For example, three-fourths of the coding bases of human
histone H4, a protein of crucial importance in packaging DNA, have remained unaltered since
we last shared an ancestor with plants. However, if the scrutinized nucleotide was not functional,
changes would not have been selected against and thus would have occurred more frequently. At
the end of these very long time periods, a neutrally evolving sequence would have experienced so
many changes that little trace of its ancestral sequence would remain in extant species. After this
amount of divergence time, alignment of DNA sequence has often become so inaccurate as to be
uninformative.
Comparisons of sequences of extant species whose common ancestors lived in this long-ago
time period often reveal short regions that are alignable and well conserved, separated by long
stretches that are unalignable and poorly conserved. For example, only 2.5% of nucleotides align
in the genomes of chickens and humans, species that last shared a common ancestor a little
over 300 million years ago (52). When an accurate alignment can be obtained, it permits an
evolutionary change to be assigned, usually unambiguously, to a lineage in a phylogeny. When
two such alignments are compared, different numbers of changes assigned to this lineage can
indicate differences in these two regions’ evolutionary rates. A region whose evolutionary rate
significantly exceeds that of another region, chosen specifically because it is believed to have
evolved neutrally (Table 1), is likely to have experienced episodes of positive selection, perhaps
owing to adaptive evolution. A region whose evolutionary rate is significantly lower than that of
the putatively neutrally evolved region has been subject to constraint—in other words, negative
Purifying selection: (or purifying) selection. Conserved sequence is not necessarily constrained: Human and great ape
the process by which genomes are highly conserved because of their relatively recent common ancestry, yet most of their
deleterious mutations sequence is not constrained. Conversely, constrained sequence is not necessarily conserved if its
are preferentially
function has arisen only recently and thus is not shared with the other species under consideration.
removed from the
population Metazoan genomes differ substantially in their fractions of constrained sequence, depending on
their gene content and the size and extent of their regulatory sequence (103) (Figure 2).
Constrained
Genome Protein-coding
Coding Noncoding Unconstrained size (Mb) genes
Figure 2
Proportion of the genome identified as being under constraint (purifying selection) when using the neutral indel model (79, 84).
Alignments between Homo sapiens and Macaca mulatta, Mus musculus and Rattus norvegicus, Gallus gallus and Taeniopygia guttata, Takifugu
rubripes and Tetraodon nigroviridis, and Drosophila melanogaster and Drosophila simulans were used to produce the fraction of the genome
under constraint. The proportions of sequence under constraint depend on the species pair analyzed and the alignment-quality
processing. Data from Meader et al. (84).
74 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
Detecting species-specific functional sequence that has only recently become constrained re-
quires comparisons not between species whose lineages have long been separated but rather within
a population—for example, that of modern humans. Over the several hundred thousand years since
dN : the number of
the ancestral population of modern humans arose, mutations have risen and fallen in population nonsynonymous
frequency: Those that are advantageous tend to rise in frequency because of positive selection, substitutions per
whereas those that are substantially deleterious tend toward extinction because of negative selec- nonsynonymous site
tion. Approaches to identify the imprint of selection thus compare the population frequencies of dS : the number of
alleles in the regions of interest against those in putatively neutrally evolved sequence: Higher synonymous
frequencies might reflect episodes of positive selection, whereas lower frequencies might reflect substitutions per
synonymous site
purifying selection. Importantly, even if a coding sequence region shows a signature of constraint,
it is not possible to infer definitively whether this discriminative sieve of selection has acted on
DNA or RNA or on protein molecular functions.
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
3. BETWEEN-SPECIES COMPARISONS
0.5
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
0.4
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
0.2
Average phastCons score
0.9
D. melanogaster
0.8
0.8 0.7
0.6
0.5
0.6
0.4
0.2
Figure 3
Average nucleotide conservation (phastCons) score sampled across 5 untranslated region (UTR), coding DNA sequence (CDS), and
3 UTR exons and introns in Homo sapiens and Drosophila melanogaster at the first (orange), second (dark red ), and third (dark blue) codon
positions and fourfold degenerate sites (dark aqua). Only exons and introns that were at least 200 nucleotides long and did not overlap
other annotated features were used. The high conservation values observed for the intermediate 3 UTR exons in D. melanogaster are
likely stochastic noise, a consequence of their low number (20 exons larger than 200 nucleotides). The red and light blue shaded areas
represent the range of nucleotide conservation across long noncoding RNA gene models in H. sapiens and D. melanogaster, respectively.
76 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
BACKGROUND SELECTION
The local environment (background) of a variant can affect whether it is purged from a population. For example,
neutral variation in a UTR will be reduced because of its linkage with genomically proximal coding sequence that is
under negative selection. This is due to neutral changes propagating in a population for an extended period of time
only when gametes are free of deleterious alleles (19). Recombination disrupts genetic linkage and thus diminishes
background selection.
greater than 1 when positive selection has acted to preferentially fix amino acid changes in the
population (56). Selective pressure on synonymous sites, however, is not negligible. This is be-
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
cause for many species there is selection on codon usage and for the preservation of splicing
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
regulatory elements (18, 37). In mammals, approximately 20% of 4D sites are subject to a non-
negligible degree of selective constraint (33). Consequently, the use of 4D sites to model neu-
tral evolution in tests for selection can lead to significant biases (Table 1). Splicing regulatory
elements tend to lie near intron–exon boundaries (within 50 base pairs), and it is their conserva-
tion that appears to underlie the peak of conservation scores close to these boundaries (95, 125)
(Figure 3). The heterogeneity of conservation within and between coding exons (Figure 3) in-
dicates that a substantial amount of selection occurs not because of changes in protein function
but because of RNA and/or DNA function. This is important because ignoring selection on nu-
cleotide function in coding sequence leads to underestimation of neutral rates, which then leads
to higher rates of false predictions of positive selection and more false negatives for purifying
selection.
As we have observed, noncoding exonic bases of UTRs are better conserved than noncoding
intronic bases (Figure 3). This higher degree of conservation is due to its content of various types of
functional elements, such as G-quadruplex sequences, internal ribosome entry sites, and upstream
open reading frames (7, 33, 54, 94, 109); conservation may also reflect background selection (see
sidebar, Background Selection). Many mammalian 5 UTRs also contain spliced exons whose
intronic boundaries, similarly to those in coding sequence, show elevated nucleotide conservation
(107) (Figure 3), presumably reflecting the high density of splicing regulatory elements lying
near exon junctions. The increased sequence conservation of 3 UTRs over introns is attributable
at least in part to purifying selection acting to preserve or avoid microRNA (miRNA) binding
sites, also called miRNA response elements (MREs) (20, 112, 129). Mammalian miRNA seed
sequences bind via base-pairing with imperfectly complementary MREs within mRNAs, leading
to translational repression or target degradation (8). Based on the conservation of at least 45,000
MRE: microRNA
miRNA target sites within human 3 UTRs, the majority of human genes appear to have been response element
under selective pressure to maintain pairing to miRNAs (40). Mutations that create MREs can
Neutral proxy:
also be under negative selection because of their deleterious effect on gene expression regulation sequence regions that
(20, 112). are considered to be
Some investigators consider intronic sequences to have evolved neutrally and thus to provide a neutrally evolving and
useful neutral proxy against which selection on coding sequences can be assessed (Table 1). How- that are used as the
null expectation when
ever, in addition to sequences flanking splice acceptor and donor sites and other splicing regulatory
testing for selection
motifs that occur within 200 base pairs of exons, introns contain diverse and numerous functional that acted on
elements that all contribute to lower evolutionary rates (17, 81). Kim & Pritchard (63) reported sequences of interest
that approximately 37,000 conserved noncoding elements fall within introns, representing ap-
proximately 4.6 Mb, although a more recent estimate of intronic functional elements numbers
these at nearly 1.5 million by counting all DNase I–hypersensitive sites (118). Evolutionary rates
also vary depending on the position of an intron within a gene (45). For example, relative to other
introns within a gene, first introns tend to be better conserved, are on average twice as long, and
differ in their nucleotide composition (17, 45).
lncRNA: long
noncoding RNA
(Figures 3 and 4). Nevertheless, across mammalian evolution the strength of this selection on
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
mature lncRNA sequences has been only modest, and sequence conservation only marginally
exceeds that seen for intronic and untranscribed intergenic sequences (Figure 4). Indeed, based
on evidence for purifying selection on inserted or deleted sequences, only 5% of all bases in mouse
RNA transcripts are estimated to be functional (100).
This low level of conservation might imply that such lncRNAs are rarely functional. Instead,
they could be associated with transcriptional noise (114), perhaps emanating from transcription at
a neighboring locus (27). If biologically relevant, the lack of sequence conservation of transcribed
intergenic noncoding loci is likely to reflect the presence in RNA transcripts of only short patches of
functional sequence involved in base-pairing or protein interactions, similar to the limited MREs in
3 UTRs of protein-coding genes (13, 16). The observed low nucleotide conservation of lncRNAs
could also be associated with functional redundancy (24, 108). In contrast, rapid divergence could
result from compensatory mutations that are rapidly fixed when they mitigate the deleterious effect
of a second mutation at a functional site. This mechanism has been proposed to explain both the
accumulation of potentially deleterious mutations within protein-coding sequences or regulatory
elements (67, 130) and the maintenance of RNA secondary structures (98).
78 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
5' 3'
0.18 H. sapiens
0.16
0.14
0.12
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
0.10
D. melanogaster
0.5
0.4
0.3
0.2
0.1
Figure 4
Average nucleotide conservation (phastCons) scores sampled across intergenic long noncoding RNA loci in Homo sapiens and Drosophila
melanogaster.
genes that are frequently transcribed in the germline also likely explains the otherwise curious
observation that promoters of non-protein-coding RNA genes tend to be better conserved than
promoters of protein-coding genes (3, 82, 100). At the 3 termini of genes, sequence in the vicinity
of the polyadenylation site (motif AWUAAA) also exhibits a trend for more stringent purifying
selection, presumably owing to its requirement by cleavage and polyadenylation specificity factors
(23, 90).
The effects of mutations in noncoding sequence can sometimes be predicted only by the
scrutiny of purifying selection and are not corroborated by the scrutiny of laboratory experiments.
For example, the deletion of long intergenic sequences containing more than 1,000 noncoding
elements, each conserved between human and rodents (>70% sequence identity over 100 base
pairs), in laboratory mice has failed to reveal phenotypic effects with respect to litter size, body
weight, or longevity (3). Similarly, the deletion of elements that are completely conserved in
Effective population
size (Ne ): the number sequence between humans, mice, and rats or the deletion of a conserved enhancer region involved
of individuals in an in limb patterning leads to viable and fertile mice lacking overt phenotypes (3, 23). Knockout
idealized population mutant mice for three intergenic lncRNA loci (Hotair, Malat1, and Neat1) also show no overt
that show a level of phenotype, even though these loci are among the most highly expressed and most highly conserved
loss of heterozygosity
of all intergenic lncRNAs (29, 89, 90, 108, 132).
due to genetic drift
equivalent to that of How can the absence of overt phenotypes be reconciled with the deep phylogenetic conser-
the actual population vation of these elements? Different hypotheses have been proposed to resolve this conundrum.
TFBS: transcription First, phenotypes not observable under laboratory conditions may be revealed under more natural
factor binding site conditions. Knockout mutants for the BC1 locus, for example, showed no phenotypic changes
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
except when assayed in natural conditions (75). Second, changes may be missed owing to shallow
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
phenotyping. Fewer than one-fifth of genes in either Caenorhabditis elegans or Saccharomyces cere-
visiae initially yielded phenotypic effects when disrupted, yet upon more extensive investigation,
larger proportions (42–60% and 97% of genes, respectively) were found to yield significant effects
(55, 105). Natural selection is thus better able to discriminate deleterious from neutral variants
than are laboratory experiments.
80 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
118 nucleotides since humans’ last common ancestor with chimpanzees. These substitutions are
very strongly biased toward G or C bases, indicating that, rather than being adaptive changes, they
are the consequences of a mutational process, specifically recombination-associated GC-biased
gene conversion (44).
sites appear to have been subject to selective constraints (22, 74); and sequences upstream of their
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
transcription start sites are diverged, a likely consequence of a lower degree of selection relative
to transcribed regions (80). Additionally, TFBSs, insulator sites, and RNA polymerase II binding
sites all exhibit moderately strong conservation patterns across these drosophilids, as (to a lesser
extent) do UTRs (86).
Nonetheless, several important differences are apparent. In contrast to mammalian coding se-
quence, codon usage is under selection in fruit flies, which strongly modulates protein sequence
evolution (53, 110). Moreover, introns in Drosophila species are subject to a substantial degree of
selection (46), excepting those that are short (<86 nucleotides) (22, 74) (Table 1). Perhaps the
most striking difference with mammalian sequence is that fruit fly lncRNA loci are moderately well
conserved across fly species (86, 131), particularly in their exons (48) (Figure 4). This could reflect
a more central functional role of these noncoding genes in fruit fly evolution and biology. Never-
theless, such differences are best explained through the application of the theoretical framework
provided by population genetics. Indeed, more generally, the ability to distinguish between tran-
scriptional noise and signal, or between random and consequential binding events, or between
ephemeral or long-lasting functional sequences requires methods that go beyond species-level
conservation. To achieve these distinctions requires evidence of constraint among individuals of
a species.
The converse is expected for advantageous mutations: They should become fixed or segregate at a
higher frequency than neutral mutations. Accordingly, the recent 1000 Genomes Project analysis
as well as an analysis of 6,515 exomes demonstrated a negative correlation between the age of a
DAF: derived allele
frequency mutation and its deleterious effect, with most deleterious mutations within the human populations
having arisen only within the past 5,000 years (2, 42, 62).
Genetic hitchhiking:
the process by which Population-based studies initially investigated the density of polymorphic sites in different se-
positive selection at quence categories, based on the idea that polymorphisms in functional sites will have been more
one site leads to an frequently purged if they are deleterious and more rapidly fixed if they are advantageous. For
increased allele example, both 5 and 3 sequences flanking genes show an excess of low-frequency polymorphisms
frequency at a linked
relative to intergenic sites, presumably owing to a combination of background selection origi-
neutral site
nating from linked protein-coding sequence and their cis-regulatory element content (73, 119).
Furthermore, within human populations, both 5 and 3 UTRs exhibit a significantly lower den-
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
sity of polymorphic sites than do introns or putatively neutral sequences (pseudogenes, ancestral
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
repeats) (73, 88). Although polymorphic sites tend to be evenly distributed across 3 UTRs, single-
nucleotide polymorphism (SNP) density is significantly lower at the UTR boundaries (and, more
specifically, at the polyadenylation sites) and within miRNA binding sites (20, 117).
Analyses of the site frequency spectrum, in particular minor allele frequencies or derived allele
frequencies (DAFs), can infer selection not for single functional elements but for many such
elements when they are considered together as a class. Low frequencies of alleles could reflect the
purging of deleterious variants from the population but could also reflect population structure or
past demographic events. Consequently, selection is best inferred by comparing allele frequencies
in the sequence class of interest to frequencies in a set of presumed neutrally evolving sequences
(Table 1). The strength of purifying selection acting on recently arisen deleterious variants can
be estimated from the excess density of low-frequency derived alleles in a sequence class (e.g.,
lncRNA exons) compared with such alleles within putatively neutral sequences (Table 1).
As expected, the different strengths of selection at various genomic elements, estimated using
this DAF test, mirror those derived from cross-species comparisons. In humans, the skew toward
rarer derived alleles, and thus stronger selective constraints, is strongest for variants within both
nonsynonymous and splice sites. In contrast, lower selective constraints act on variants in UTRs,
introns, and intergenic regions (1, 2, 88). Only 6% of synonymous substitutions are predicted
to be deleterious in humans (116), a lower proportion than the 20% inferred from cross-species
comparisons (33). As it does not rely on interspecies sequence conservation, the DAF test can
identify constraint that has arisen recently. For example, it was used to predict that 30–50% of
nonconserved MREs are functional when their mRNA and cognate miRNA are coexpressed (20).
The method has also been used in an important demonstration that sets of human noncoding
sequences are conserved not because they are subject to low mutation rates but rather because
mutations within them have been preferentially purged from the human population (6, 26, 58,
119).
The intersection of human sequence variation (2) and functional annotations (32) permits
population-based estimation of selection acting on open chromatin, transcription factor binding,
or more generally transcribed sequence (5, 48, 118). TFBSs, such as those for POU-, HOX-,
or FOX-domain-containing transcription factors, include both a lower density of SNPs and
an excess of low-frequency SNPs relative to putatively neutral sites or similar sequences in the
genome that are without evidence of binding (5, 111, 121). Using a more rigorous approach that
better corrects for selection’s effect of distorting the pattern of polymorphism at neutral sites
(background selection, genetic hitchhiking), Arbiza et al. (5) estimated that, on average, 33% of
the nucleotides in a set of high-confidence TFBSs are under selection. They also reported that
patterns of selection identified in TFBSs varied greatly depending on the transcription factor
82 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
H. sapiens D. melanogaster
0.2
Figure 5
Comparison of derived allele frequency spectra in Homo sapiens and Drosophila melanogaster for zerofold (dark blue) and fourfold (light
blue) degenerate sites, intergenic long noncoding RNAs (lncRNAs) (red ), and neutrally evolving ancestral repeats (H. sapiens) or small
introns (D. melanogaster) (dark yellow). Adapted from Haerty & Ponting (48).
(5, 88) and that the strength of selection depended on both the size of the binding site and the
number of degenerate positions in the consensus motif.
It is striking that applying the DAF test to sequences and populations from other species yields
results that contrast with those obtained from the human population. Drosophila conserved non-
coding sequences, for example, appear to be under greater purifying selection than such sequences
in humans (15). This is most evident for intergenic lncRNAs, which in fruit flies, but not in hu-
mans, exhibit a clear excess of low-frequency alleles (48) (Figure 5). These differences extend to
other mammals because selective constraints on conserved noncoding sequences, including pro-
moters, in rodents appear to be almost twice as strong as those in the human population (0.53 ±
0.01 and 0.29 ± 0.02, respectively) (60, 115). These differences are explained below as a direct
effect of these species’ contrasting effective population sizes.
Comparison of DAF test results between species is complicated by their population size histories
and structures. The human population underwent multiple bottlenecks of various constrictions
that reduced its effective population size to approximately 1,200 individuals and were then followed
by demographic expansion (76). These changes strongly bias allele frequencies at neutral sites
and, if not properly accounted for, lead to false inferences of selection. This is because there is an
enrichment of rare variants at neutral sites in cases of population expansion (43) and an increased
frequency of intermediate variants in cases of population contraction or bottlenecks (123, 127).
Additionally, population range expansions associated with founder effects at the margin of the
population spatial distribution can lead to a strong bias in allele frequency that mimics the action
of positive selection (also called allele surfing) (65).
Methods that correct for the otherwise confounding demographic effects have been developed
that infer the distribution of fitness effects from allele frequencies (127). Three methods, developed
by Eyre-Walker et al. (36), Boyko et al. (11), and Keightley & Eyre-Walker (59), estimate demo-
graphic parameters from user-provided putatively neutral sites to correct the frequency spectrum
of sites of interest (11, 34, 36, 59, 127). Each of these methods is based on predictions from the
nearly neutral theory of evolution (64, 92, 93) together with assumptions that adaptive polymor-
phisms should be very rare, because they are fixed rapidly, and that the majority of nonneutral
mutations are deleterious. Consequently, based on the product of the effective population size
(Ne ) and the selection coefficient, mutations are classified as being effectively neutral (|Ne s| < 1),
weakly deleterious (1 < |Ne s| < 10), deleterious (10 < |Ne s| < 100), or strongly deleterious
(|Ne s| > 100). These algorithms have been applied to polymorphism data in humans, mice, and
fruit flies to assess the variable strength of selection at both coding and putatively functional non-
coding sites (Figure 6). In humans, 27–38% of mutations at nonsynonymous sites were deemed
effectively neutral, 21–30% were weakly deleterious, and the remainder were strongly deleterious
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
(11, 34). By comparison, the DAF spectrum of SNPs within intergenic lncRNAs was not different
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
from putatively neutrally evolving neighboring sequences, providing no evidence for the action
of purifying selection acting contemporarily on human lncRNAs. Nevertheless, the same anal-
ysis applied to intergenic lncRNAs from D. melanogaster predicted that approximately 30% of
mutations within these loci are weakly deleterious (48) (Figure 6).
84 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
N EU T R A L D E LE T E RIO U S
|Nes|: 0–1 1 – 10 10 – 100 10 – ∞ 100 – ∞
1.00
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
0.75
Proportion of sites
0.50
0.25
Zerofold 5’ UTR 3’ UTR lncRNA Zerofold 5’ UTR 3’ UTR CNE Zerofold 5’ UTR Intron 3’ UTR lncRNA
degenerate degenerate degenerate
Type of site
Figure 6
Comparison of the distribution of fitness effects of mutations occurring within 5 and 3 untranslated regions (UTRs), zerofold
degenerate sites, introns, conserved noncoding elements (CNEs), and long noncoding RNAs (lncRNAs) in Homo sapiens, Mus musculus
castaneus, and Drosophila melanogaster. For the CNEs in M. musculus castaneus, only three categories are present: |Ne s| < 1, 1 < |Ne s| <
10, and 10 < |Ne s| < ∞. Data from Haerty & Ponting (48), Eyre-Walker & Keightley (34), Halligan et al. (50), and Kousathanas et al.
(69).
sites, particularly in high-Ne species, will be the least applicable. Ancestral repeats, defined as
transposable element sequences that inserted prior to the last common ancestor of two species
(often humans and mice), are proposed to be the most reliable neutral proxy because virtually all
of them have been demonstrated to have evolved neutrally (79, 102). Only 0.2% and 1% of bases
in rat–mouse and human–mouse ancestral repeat sequences, respectively, depart from neutrality
(84).
Neutral sites are commonly collated across the whole genome, thereby forming a putatively
neutral reference against which test sequence is compared. However, this overlooks important
confounding issues. First, because test and neutral sites are most often not fully interdigitated,
they are likely to not share the same genealogical history (4, 85, 133), and comparisons will
not necessarily account for mutation rates, which vary greatly across the genome (31). Second,
selection acting at linked sites, via either background selection (see sidebar, Background Selection)
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
The optimal choice of neutral proxy is sequence that has escaped selection and is interdigi-
tated or that lies adjacent to and is compositionally equivalent to sites of interest (for example,
nonsynonymous sites, MREs, and lncRNAs) (4, 133). These requirements are met by matching
candidate neutral sequence that lies in close physical proximity, and is compositionally similar,
to each test sequence. Nevertheless, matches are often not possible when, for example, ancestral
repeats of the required composition are absent from the genomic vicinity of the test sequence.
Consequently, a compromise procedure, adopted by Halligan et al. (50), Haerty & Ponting (48),
and Arbiza et al. (5), is to employ proxy neutral sequence that is noncoding and not conserved
in diverse species and that lies outside of annotations such as TFBSs or DNase I–hypersensitive
sites. Such sequence constitutes the majority of the reference human genome but only a small
proportion of the fruit fly genome.
SUMMARY POINTS
1. Sequences that have long retained functionality are characterized by an increased inter-
species sequence conservation that derives from the past purging of mutations that had
a deleterious impact on fitness.
2. Purifying selection is strongest on protein-coding sequences, more moderate on flank-
ing noncoding functional sequences (5 and 3 untranslated regions), and very weak on
intergenic transcribed or bound functional elements that tend to turn over rapidly in
evolution.
3. The most powerful approach to assess the current biological relevance of these rapidly
evolving elements is to analyze variants’ site frequency spectra.
4. Because a species’ effective population size directly affects the efficacy of natural selection
on mutations, classes of elements identified as evolving under selective constraints in one
species may be effectively neutral in another species of smaller effective population size.
5. The observation of a biochemical event at a locus is necessary but not sufficient to infer
that it is functional. Instead, indication of biochemical activity, such as transcription or
transcription factor binding, combined with the inference of selective constraint between
or within species provides the most compelling evidence for the functionality of elements
that can then be prioritized for subsequent experimental validation.
86 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
FUTURE ISSUES
1. With decreasing sequencing costs of large numbers of samples, population genomic ap-
proaches that assess the functionality of elements within the genome will spread. This will
allow the in silico assessment of the biological relevance of a locus that draws upon both
evolutionary and functional (for example, transcription or transcription factor binding)
evidence.
2. Not all transcriptional or binding events will be biologically relevant. However, we do
not have a clear understanding of the expected background levels of such biological
noise when interpreting functional genomic data. The use of more quantitative methods
to investigate the strength of transcription factor binding (77) will also greatly help to
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
identify genomic regions in which binding is only transient and likely inconsequential.
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
3. Many tests for selection were initially developed for protein-coding sequence and only
later adapted for non-protein-coding sequence. There is a considerable need for new tests
that consider the deleterious effects of mutation and are tailored specifically to functional
noncoding elements, because these are the primary substrate of rapid turnover along
different evolutionary lineages.
4. We know almost nothing about the relative impacts that disruption of classes of func-
tional noncoding sequence—including microRNA response elements, transcription fac-
tor binding sites, and long noncoding RNA loci—have on fitness and phenotypes. There
is thus a pressing need for large-scale in vivo and cellular mutagenesis and phenotyping
projects.
DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.
ACKNOWLEDGMENTS
We thank Jim Kent (University of California, Santa Cruz) for helpful discussions on the conser-
vation of mouse/human genes. We are grateful to Chris Rands for his comments on the early
versions of this review. W.H. and C.P.P. are supported by the Medical Research Council, and this
work was also supported by a European Research Council Advanced Grant to C.P.P.
LITERATURE CITED
1. 1000 Genomes Proj. Consort. 2010. A map of human genome variation from population-scale sequenc-
ing. Nature 467:1061–73
2. 1000 Genomes Proj. Consort. 2012. An integrated map of genetic variation from 1,092 human genomes.
Nature 491:56–65
3. Ahituv N, Zhu Y, Visel A, Holt A, Afzal V, et al. 2007. Deletion of ultraconserved elements yields viable
mice. PLoS Biol. 5:e234
4. Andolfatto P. 2008. Controlling type-I error of the McDonald–Kreitman test in genomewide scans for
selection on noncoding DNA. Genetics 180:1767–71
5. Arbiza L, Gronau I, Aksoy BA, Hubisz MJ, Gulko B, et al. 2013. Genome-wide inference of natural
selection on human transcription factor binding sites. Nat. Genet. 45:723–29
6. Asthana S, Noble WS, Kryukov G, Grant CE, Sunyaev S, Stamatoyannopoulos JA. 2007. Widely dis-
tributed noncoding purifying selection in the human genome. Proc. Natl. Acad. Sci. USA 104:12410–15
7. Barreiro LB, Laval G, Quach H, Patin E, Quintana-Murci L. 2008. Natural selection has driven popu-
lation differentiation in modern humans. Nat. Genet. 40:340–45
8. Bartel DP. 2004. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116:281–97
9. Blow MJ, McCulley DJ, Li Z, Zhang T, Akiyama JA, et al. 2010. ChIP-Seq identification of weakly
conserved heart enhancers. Nat. Genet. 42:806–10
10. Boffelli D, Nobrega MA, Rubin EM. 2004. Comparative genomics at the vertebrate extremes. Nat. Rev.
Genet. 5:456–65
11. Boyko AR, Williamson SH, Indap AR, Degenhardt JD, Hernandez RD, et al. 2008. Assessing the
evolutionary impact of amino acid mutations in the human genome. PLoS Genet. 4:e1000083
12. Cabili MN, Trapnell C, Goff L, Koziol M, Tazon-Vega B, et al. 2011. Integrative annotation of human
large intergenic noncoding RNAs reveals global properties and specific subclasses. Genes Dev. 25:1915–27
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
13. Calin GA, Liu CG, Ferracin M, Hyslop T, Spizzo R, et al. 2007. Ultraconserved regions encoding
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
ncRNAs are altered in human leukemias and carcinomas. Cancer Cell 12:215–29
14. Carninci P, Sandelin A, Lenhard B, Katayama S, Shimokawa K, et al. 2006. Genome-wide analysis of
mammalian promoter architecture and evolution. Nat. Genet. 38:626–35
15. Casillas S, Barbadilla A, Bergman CM. 2007. Purifying selection maintains highly conserved noncoding
sequences in Drosophila. Mol. Biol. Evol. 24:2222–34
16. Cesana M, Cacchiarelli D, Legnini I, Santini T, Sthandier O, et al. 2011. A long noncoding RNA controls
muscle differentiation by functioning as a competing endogenous RNA. Cell 147:358–69
17. Chamary JV, Hurst LD. 2004. Similar rates but different modes of sequence evolution in introns and at
exonic silent sites in rodents: evidence for selectively driven codon usage. Mol. Biol. Evol. 21:1014–23
18. Chamary JV, Parmley JL, Hurst LD. 2006. Hearing silence: non-neutral evolution at synonymous sites
in mammals. Nat. Rev. Genet. 7:98–108
19. Charlesworth B, Morgan MT, Charlesworth D. 1993. The effect of deleterious mutations on neutral
molecular variation. Genetics 134:1289–303
20. Chen K, Rajewsky N. 2006. Natural selection on human microRNA binding sites inferred from SNP
data. Nat. Genet. 38:1452–56
21. Clark AG, Eisen MB, Smith DR, Bergman CM, Oliver B, et al. 2007. Evolution of genes and genomes
on the Drosophila phylogeny. Nature 450:203–18
22. Clemente F, Vogl C. 2012. Evidence for complex selection on four-fold degenerate sites in Drosophila
melanogaster. J. Evol. Biol. 25:2582–95
23. Cretekos CJ, Wang Y, Green ED, Martin JF, Rasweiler JJ IV, Behringer RR. 2008. Regulatory divergence
modifies limb length between mammals. Genes Dev. 22:141–51
24. Derrien T, Johnson R, Bussotti G, Tanzer A, Djebali S, et al. 2012. The GENCODE v7 catalog of
human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome Res.
22:1775–89
25. Dobzhansky T. 1964. Biology, molecular and organismic. Am. Zool. 4:443–52
26. Drake JA, Bird C, Nemesh J, Thomas DJ, Newton-Cheh C, et al. 2006. Conserved noncoding sequences
are selectively constrained and not mutation cold spots. Nat. Genet. 38:223–27
27. Ebisuya M, Yamamoto T, Nakajima M, Nishida E. 2008. Ripples from neighbouring transcription. Nat.
Cell Biol. 10:1106–13
28. Eddy SR. 2005. A model of the statistical power of comparative genome sequence analysis. PLoS Biol.
3:e10
29. Eißmann,Gutschner T, Hämmerle M, Günther S, Caudron-Herger M, et al. 2012. Loss of the abundant
nuclear non-coding RNA MALAT1 is compatible with life and development. RNA Biol. 9:1076–87
30. Elgar G, Vavouri T. 2008. Tuning in to the signals: noncoding sequence conservation in vertebrate
genomes. Trends Genet. 24:344–52
31. Ellegren H, Smith NG, Webster MT. 2003. Mutation rate variation in the mammalian genome. Curr.
Opin. Genet. Dev. 13:562–68
32. ENCODE Proj. Consort. 2012. An integrated encyclopedia of DNA elements in the human genome.
Nature 489:57–74
88 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
33. Eory L, Halligan DL, Keightley PD. 2010. Distributions of selectively constrained sites and deleterious
mutation rates in the hominid and murid genomes. Mol. Biol. Evol. 27:177–92
34. Eyre-Walker A, Keightley PD. 2009. Estimating the rate of adaptive molecular evolution in the presence
of slightly deleterious mutations and population size change. Mol. Biol. Evol. 26:2097–108
35. Eyre-Walker A, Keightley PD, Smith NG, Gaffney D. 2002. Quantifying the slightly deleterious mu-
tation model of molecular evolution. Mol. Biol. Evol. 19:2142–49
36. Eyre-Walker A, Woolfit M, Phelps T. 2006. The distribution of fitness effects of new deleterious amino
acid mutations in humans. Genetics 173:891–900
37. Fairbrother WG, Holste D, Burge CB, Sharp PA. 2004. Single nucleotide polymorphism-based valida-
tion of exonic splicing enhancers. PLoS Biol. 2:E268
38. Fisher S, Grice EA, Vinton RM, Bessling SL, McCallion AS. 2006. Conservation of RET regulatory
function from human to zebrafish without sequence similarity. Science 312:276–79
39. Fisher WW, Li JJ, Hammonds AS, Brown JB, Pfeiffer BD, et al. 2012. DNA regions bound at low
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
occupancy by transcription factors do not drive patterned reporter gene expression in Drosophila. Proc.
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
60. Keightley PD, Kryukov GV, Sunyaev S, Halligan DL, Gaffney DJ. 2005. Evolutionary constraints in
conserved nongenic sequences of mammals. Genome Res. 15:1373–78
61. Keightley PD, Lercher MJ, Eyre-Walker A. 2005. Evidence for widespread degradation of gene control
regions in hominid genomes. PLoS Biol. 3:e42
62. Kiezun A, Pulit SL, Francioli LC, van Dijk F, Swertz M, et al. 2013. Deleterious alleles in the human
genome are on average younger than neutral alleles of the same frequency. PLoS Genet. 9:e1003301
63. Kim SY, Pritchard JK. 2007. Adaptive evolution of conserved noncoding elements in mammals. PLoS
Genet. 3:1572–86
64. Kimura M. 1983. The Neutral Theory of Molecular Evolution. Cambridge, UK: Cambridge Univ. Press
65. Klopfstein S, Currat M, Excoffier L. 2006. The fate of mutations surfing on the wave of a range expansion.
Mol. Biol. Evol. 23:482–90
66. Knowles DG, McLysaght A. 2009. Recent de novo origin of human protein-coding genes. Genome Res.
19:1752–59
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
67. Kondrashov AS, Sunyaev S, Kondrashov FA. 2002. Dobzhansky-Muller incompatibilities in protein
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
90 Haerty · Ponting
GG15CH04-Ponting ARI 15 July 2014 11:5
87. Mouse Genome Seq. Consort. 2002. Initial sequencing and comparative analysis of the mouse genome.
Nature 420:520–62
88. Mu XJ, Lu ZJ, Kong Y, Lam HYK, Gerstein MB. 2011. Analysis of genomic variation in non-coding
elements using population-scale sequencing data from the 1000 Genomes Project. Nucleic Acids Res.
39:7058–76
89. Nakagawa S, Ip JY, Shioi G, Tripathi V, Zong X, et al. 2012. Malat1 is not an essential component of
nuclear speckles in mice. RNA 18:1487–99
90. Nakagawa S, Naganuma T, Shioi G, Hirose T. 2011. Paraspeckles are subpopulation-specific nuclear
bodies that are not essential in mice. J. Cell Biol. 193:31–39
91. Odom DT, Dowell RD, Jacobsen ES, Gordon W, Danford TW, et al. 2007. Tissue-specific transcrip-
tional regulation has diverged significantly between human and mouse. Nat. Genet. 39:730–32
92. Ohta T. 1973. Slightly deleterious mutant substitutions in evolution. Nature 246:96–98
93. Ohta T, Gillespie JH. 1996. Development of neutral and nearly neutral theories. Theor. Popul. Biol.
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
49:128–42
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
94. Osada N, Hirata M, Tanuma R, Kusuda J, Hida M, et al. 2005. Substitution rate and structural divergence
of 5 UTR evolution: comparative analysis between human and cynomolgus monkey cDNAs. Mol. Biol.
Evol. 22:1976–82
95. Parmley JL, Urrutia AO, Potrzebowski L, Kaessmann H, Hurst LD. 2007. Splicing and the evolution
of proteins in mammals. PLoS Biol. 5:e14
96. Parsch J, Novozhilov S, Saminadin-Peter SS, Wong KM, Andolfatto P. 2010. On the utility of short
intron sequences as a reference for the detection of positive and negative selection in Drosophila. Mol.
Biol. Evol. 27:1226–34
97. Pennacchio LA, Ahituv N, Moses AM, Prabhakar S, Nobrega MA, et al. 2006. In vivo enhancer analysis
of human conserved non-coding sequences. Nature 444:499–502
98. Piskol R, Stephan W. 2008. Analyzing the evolution of RNA secondary structures in vertebrate introns
using Kimura’s model of compensatory fitness interactions. Mol. Biol. Evol. 25:2483–92
99. Pollard KS, Salama SR, Lambert N, Lambot MA, Coppens S, et al. 2006. An RNA gene expressed during
cortical development evolved rapidly in humans. Nature 443:167–72
100. Ponjavic J, Ponting CP, Lunter G. 2007. Functionality or transcriptional noise? Evidence for selection
within long noncoding RNAs. Genome Res. 17:556–65
101. Ponting CP, Hardison RC. 2011. What fraction of the human genome is functional? Genome Res.
21:1769–76
102. Ponting CP, Lunter G. 2006. Signatures of adaptive evolution within human non-coding sequence.
Hum. Mol. Genet. 15(Suppl. 2):R170–75
103. Ponting CP, Nellåker C, Meader S. 2011. Rapid turnover of functional sequence in human and other
genomes. Annu. Rev. Genomics Hum. Genet. 12:275–99
104. Ponting CP, Oliver PL, Reik W. 2009. Evolution and functions of long noncoding RNAs. Cell 136:629–
41
105. Ramani AK, Chuluunbaatar T, Verster AJ, Na H, Vu V, et al. 2012. The majority of animal genes are
required for wild-type fitness. Cell 148:792–802
106. Rands CM, Meader S, Ponting CP, Lunter G. 2014. 8.2% of the human genome is constrained: variation
in rates of turnover across functional element classes in the human lineage. PLoS Genet. In press
107. Roy SW, Penny D, Neafsey DE. 2007. Evolutionary conservation of UTR intron boundaries in Cryp-
tococcus. Mol. Biol. Evol. 24:1140–48
108. Schorderet P, Duboule D. 2011. Structural and functional differences in the long non-coding RNA
Hotair in mouse and human. PLoS Genet. 7:e1002071
109. Siepel A, Bejerano G, Pedersen JS, Hinrichs AS, Hou M, et al. 2005. Evolutionarily conserved elements
in vertebrate, insect, worm, and yeast genomes. Genome Res. 15:1034–50
110. Singh ND, Larracuente AM, Clark AG. 2008. Contrasting the efficacy of selection on the X and auto-
somes in Drosophila. Mol. Biol. Evol. 25:454–67
111. Spivakov M, Akhtar J, Kheradpour P, Beal K, Girardot C, et al. 2012. Analysis of variation at transcription
factor binding sites in Drosophila and humans. Genome Biol. 13:R49
112. Stark A, Brennecke J, Bushati N, Russell RB, Cohen SM. 2005. Animal microRNAs confer robustness
to gene expression and have a significant impact on 3 UTR evolution. Cell 123:1133–46
113. Stone EA, Cooper GM, Sidow A. 2005. Trade-offs in detecting evolutionarily constrained sequence by
comparative genomics. Annu. Rev. Genomics Hum. Genet. 6:143–64
114. Struhl K. 2007. Transcriptional noise and the fidelity of initiation by RNA polymerase II. Nat. Struct.
Mol. Biol. 14:103–5
115. Taylor MS, Kai C, Kawai J, Carninci P, Hayashizaki Y, Semple CAM. 2006. Heterotachy in mammalian
promoter evolution. PLoS Genet. 2:e30
116. Tennessen JA, Bigham AW, O’Connor TD, Fu W, Kenny EE, et al. 2012. Evolution and functional
impact of rare coding variation from deep sequencing of human exomes. Science 337:64–69
117. Thomas LF, Sætrom P. 2012. Single nucleotide polymorphisms can create alternative polyadenylation
signals and affect gene expression through loss of microRNA-regulation. PLoS Comput. Biol. 8:e1002621
118. Thurman RE, Rynes E, Humbert R, Vierstra J, Maurano MT, et al. 2012. The accessible chromatin
Annu. Rev. Genom. Hum. Genet. 2014.15:71-92. Downloaded from www.annualreviews.org
119. Torgerson DG, Boyko AR, Hernandez RD, Indap A, Hu X, et al. 2009. Evolutionary processes acting
on candidate cis-regulatory regions in humans inferred from patterns of polymorphism and divergence.
PLoS Genet. 5:e1000592
120. Ulitsky I, Bartel DP. 2013. lincRNAs: genomics, evolution, and mechanisms. Cell 154:26–46
121. Vernot B, Stergachis AB, Maurano MT, Vierstra J, Neph S, et al. 2012. Personal and population genomics
of human regulatory variation. Genome Res. 22:1689–97
122. Visel A, Prabhakar S, Akiyama JA, Shoukry M, Lewis KD, et al. 2008. Ultraconservation identifies a
small subset of extremely constrained developmental enhancers. Nat. Genet. 40:158–60
123. Wakeley J, Aliacar N. 2001. Gene genealogies in a metapopulation. Genetics 159:893–905
124. Ward LD, Kellis M. 2012. Evidence of abundant purifying selection in humans for recently acquired
regulatory functions. Science 337:1675–78
125. Warnecke T, Parmley JL, Hurst LD. 2008. Finding exonic islands in a sea of non-coding sequence:
Splicing related constraints on protein composition and evolution are common in intron-rich genomes.
Genome Biol. 9:R29
126. Warnefors M, Pereira V, Eyre-Walker A. 2010. Transposable elements: insertion pattern and impact
on gene expression evolution in hominids. Mol. Biol. Evol. 27:1955–62
127. Williamson SH, Hernandez R, Fledel-Alon A, Zhu L, Nielsen R, Bustamante CD. 2005. Simultaneous
inference of selection and population growth from patterns of variation in the human genome. Proc. Natl.
Acad. Sci. USA 102:7882–87
128. Woolfe A, Goodson M, Goode DK, Snell P, McEwen GK, et al. 2005. Highly conserved non-coding
sequences are associated with vertebrate development. PLoS Biol. 3:e7
129. Xu J, Zhang R, Shen Y, Liu G, Lu X, Wu CI. 2013. The evolution of evolvability of microRNA target
sites in vertebrates. Genome Res. 23:1810–16
130. Yokoyama KD, Thorne JL, Wray GA. 2011. Coordinated genome-wide modifications within proximal
promoter cis-regulatory elements during vertebrate evolution. Genome Biol. Evol. 3:66–74
131. Young RS, Marques AC, Tibbit C, Haerty W, Bassett AR, et al. 2012. Identification and properties of
1,119 lincRNA loci in the Drosophila melanogaster genome. Genome Biol. Evol. 4:427–42
132. Zhang B, Arun G, Mao YS, Lazar Z, Hung G, et al. 2012. The lncRNA Malat1 is dispensable for mouse
development but its transcription plays a cis-regulatory role in the adult. Cell Rep. 2:111–23
133. Zhen Y, Andolfatto P. 2012. Methods to detect selection on noncoding DNA. Methods Mol. Biol. 856:141–
59
92 Haerty · Ponting
GG15-FrontMatter ARI 14 July 2014 13:54
Annual Review of
Genomics and
Human Genetics
Leon E. Rosenberg p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Access provided by WIB6106 - University Munchen on 04/17/16. For personal use only.
vi
GG15-FrontMatter ARI 14 July 2014 13:54
Errata
An online log of corrections to Annual Review of Genomics and Human Genetics articles
may be found at http://www.annualreviews.org/errata/genom
Contents vii