ISSN 0006-2979, Biochemistry (Moscow), 2024, Vol. 89, No. 6, pp. 987-1001 © The Author(s) 2024. This article is an open access publication.
987
REVIEW
Irreducible Complexity of Hox Gene:
Path to the Canonical Function of the Hox Cluster
Milana A. Kulakova
1,a
*, Georgy P. Maslakov
1
, and Liudmila O. Poliushkevich
1
1
Department of Embryology, Faculty of Biology, St. Petersburg State University, 199034 St. Petersburg, Russia
a
e-mail: m.kulakowa@spbu.ru
Received November 19, 2023
Revised March 22, 2024
Accepted March 27, 2024
AbstractThe evolution of major taxa is often associated with the emergence of new gene families. In all mul-
ticellular animals except sponges and comb jellies, the genomes contain Hox genes, which are crucial regula-
tors of development. The canonical function of Hox genes involves colinear patterning of body parts in bilateral
animals. This general function is implemented through complex, precisely coordinated mechanisms, not all of
which are evolutionarily conserved and fully understood. We suggest that the emergence of this regulatory com-
plexity was preceded by a stage of cooperation between more ancient morphogenetic programs or their individual
elements. Footprints of these programs may be present in modern animals to execute non-canonical Hox func-
tions. Non-canonical functions of Hox genes are involved in maintaining terminal nerve cell specificity, autophagy,
oogenesis, pre-gastrulation embryogenesis, vertical signaling, and a number of general biological processes. These
functions are realized by the basic properties of homeodomain protein and could have triggered the evolution of
ParaHoxozoa and Nephrozoa subsequently. Some of these non-canonical Hox functions are discussed in our review.
DOI: 10.1134/S0006297924060014
Keywords: homeodomain, ANTP, Hox genes, non-canonical functions of Hox genes, Metazoa, ParaHoxozoa,
Nephrozoa, neurogenesis, developmental autophagy, oogenesis, vertical signaling
Abbreviations: ANTP,Antennapedia; HOX,homeotic homeo-
box; GRN, gene regulatory network; PG, paralogue group;
Ubx,ultrabithorax.
* To whom correspondence should be addressed.
INTRODUCTION
The history of multicellular animal emergence is
tightly linked to the appearance of a new class of tran-
scription factors– Antennapedia (ANTP), members of
the superclass of homeodomain proteins [1]. Therapid
structural and functional evolution of the ANTP genes
led to the emergence of the most numerous and diverse
clade in the animal kingdom, now referred to as Para-
Hoxozoa [2]. This clade includes approximately 7mil-
lion species of bilaterians (Bilateria), about 10thousand
species of cnidarians (Cnidaria), and several species of
placozoans (Placozoa). The name of the clade indicates
an evolutionary boundary within Metazoa, which sep-
arates taxa with Hox/ParaHox genes from comb jellies
and sponges, which lack these genes [2-5].
Hox (homeotic homeobox) genes were the first
genes shown to be involved in development and evo-
lution [6]. Their discovery led to the emergence of a
new science – evolutionary developmental biology
(evo-devo) and made Hox genes the most stud-
ied group among all homeobox genes in animals.
The homeobox is a conserved region of the primary
sequence that encodes a DNA-binding motif of the
homeodomain, which is required for the Hox protein
to interact with enhancers of downstream target genes
[7, 8]. Hox genes are organized in a cluster, i.e., they
are physically linked. Traditionally, they are divid-
ed into 9 paralogous groups (PG1-8 and PG9/14). This
classification is based on the differences in Hox pro-
tein sequences and their relative positions in clusters.
The level of evolutionary conservation within a pa-
ralogous group (e.g., between lab (PG1) of the fruit fly
and Hox1 (PG1) of the amphioxus) is always higher
than outside of it (between lab (PG1) and pb (PG2) of
the fruit fly). Overall, the structural expansion of the
Hox cluster and the formation of the most paralogous
KULAKOVA et al.988
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
groups occurred before the appearance of the three
major Bilaterian clades [9, 10].
The main feature of the Hox cluster is the abili-
ty to exhibit a colinear expression. Colinearity refers
to the correspondence between the location of genes
on a chromosome and the order of their expression
along the anterior-posterior axis of the body [11, 12].
Thecloser is a Hox gene to the 3′-end of the cluster, the
closer to the anterior end of the embryo it will func-
tion. This is called spatial colinearity. Colinearity can
also be temporal, when genes are expressed sequential-
ly in time, starting from the 3′-end of the cluster [13].
The Hox gene cluster is the result of tandem cis-du-
plications of the ancestral sequence, starting from the
single proto-Hox gene that belonged to the NK family
[14, 15]. This event occurred before the sister branches
Bilateria and Cnidaria were formed because the Hox
genes belonging to the paralogous groups PG1, PG2,
and PG4/14 are already present in both these branch-
es [16]. These paralogous groups emerged as a result
of the diversification of cis-duplicates under two sce-
narios: neofunctionalization and subfunctionalization.
In the former case, a copy acquires a new function,
while in the latter, the ancestral function is shared
between two copies [17]. The cluster can be intact
(compact or relaxed) or contain rearrangements and
discontinuities (breakages) up to complete atomiza-
tion [18-21]. Integrity of the Hox cluster is essential for
maintaining temporal colinearity but is not crucial for
spatial colinearity [22].
Paradoxically, Hox genes are both highly con-
served and functionally flexible. Their functions are
universal during the establishment of the bilateral
body plan and species-specific at the level of local pat-
terns, such as the formation of bristles on the legs of
different Drosophila species [23]. This systemic prop-
erty is known as scalability. In the case of Hox genes,
it is manifested at both ontogenetic and phylogenetic
levels.
The staggering diversity of the bilaterian animals
results from the rapid evolution of developmental pro-
grams, which are simultaneously stable and flexible.
Bilaterian animals maintain a common body plan due
to the conserved members and parts of gene regula-
tory networks (GRNs), which start functioning shortly
after cleavage and are necessary for regionalization
and patterning. This is particularly evident in verte-
brates and other segmented animals, which undergo a
phylotypic period or “zootype” stage [24]. During this
period, representatives of a particular phylum (or sub-
phylum) are morphologically very similar to each other
(for example, all vertebrates at the pharyngula stage).
Athe molecular regulation level, the similarity is even
broader, as segmented animals from different phyla
(vertebrates, arthropods, annelids) exhibit orderly ex-
pression of Hox genes shortly before or during gastru-
lation [25-27]. The zootype concept is graphically rep-
resented as the hourglass model, where the constricted
waist corresponds to the onset of colinear transcription
of the Hox clusters along the anterior-posterior axis of
the body. Such expression is conceptually similar in all
segmented Bilateria, despite significant differences in
implementation mechanisms. Hox proteins determine
the fate of cells in broad spatial domains of the embryo
along the anterior-posterior axis of the body. During
this early period, their targets are genes of signaling
pathways and transcription factors, the sets of which
will qualitatively and quantitatively differ depending
on the Hox code. This difference ultimately will lead
to morphological and functional differences between
embryo regions.
This function is traditionally considered basic,
i.e., canonical. It is this function that is implied when
discussing the role of Hox genes in development, and
there are several reasons for this. Firstly, animals in
which we observe early colinear transcription of Hox
genes in broad spatial domains (i.e., regionalizing func-
tion) belong to the three superphyla– Deuterostomia,
Ecdysozoa, and Lophotrochozoa, which are grouped
into the clade Nephrozoa. The likelihood that such a
function of the Hox genes arose independently (conver-
gently) in these groups appears lower than the likeli-
hood of its direct inheritance from a common ancestor
of all Nephrozoa. Secondly, the sequential in-time ear-
ly activation of Hox cluster genes is always associated
with regionalization, and such a mode of activation
depends on the intactness or minimally damaged state
of the cluster [22, 28]. If the structure of the cluster
determines its regionalizing function and if a whole
cluster is apriori considered primary, it is logical to
assume that this function itself is inseparably linked to
the emergence of the Hox cluster.
The constricted waist in the hourglass model in-
dicates a lack of variability available for selection,
so it can be confidently stated that the action of Hox
genes determines the organization of the body plan, at
least in the segmented Bilateria. However, with this ap-
proach, questions about the early stages of system evo-
lution remain unsolved. The coordinated temporal and
spatial early colinear expression of Hox clusters looks
very complex. It is difficult to imagine the primary and
intermediate steps that formed this hyper network.
Moreover, if the canonical function of Hox genes is pri-
mary, then the last common ancestor of Nephrozoa is
a complex animal, not inferior in terms of organization
to the amphioxus, fruit fly, or Platynereis. This leads us
to the old paradox of irreducible complexity. Perhaps
among the multitude of functions of Hox genes, which
are not canonical ones, there are those that provide a
clue to the primary state of Hox regulation in the Para-
Hoxozoa lineage and its subsequent evolution. Some of
them are discussed in our review.
IRREDUCIBLE COMPLEXITY OF HOX GENE 989
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
HOX GENES AND NEUROGENESIS
The older the trait is, the more likely it is to be
found in the phylogenetically distant descendants of
the species that acquired this trait. The idea that the
ancestral function of Hox genes is patterning of the
nervous system was first suggested by Jordi Garcia
Fernandez [29], and not without reason. If we put the
canonical function of Hox genes out of consideration,
we are left with the most widespread and most dam-
age-resistant function– control of neurogenesis [30-32].
Control persists even in the animals that have aban-
doned early regionalizing function (leeches, appendic-
ularia, rotifers), or use it in isolation from the anteri-
or-posterior axis specification (mollusks) [19-21, 33-36].
It is noteworthy that in most of the studied mollusks,
the ganglia of the nervous system show signs of Hox
colinearity [21, 35, 36].
Modern experimental methods make it possible to
locally turn on or off selected genes at different stages
of development of model animals (nematodes, Drosoph-
ila, mouse). These experiments revealed an important
pattern. It turned out that the Hox genes do not simply
determine the cellular territories in which neuroblasts
are laid down. But they control pathways of their dif-
ferentiation and, more interestingly, establish terminal
specificity of the mature postmitotic neurons [32]. Ter-
minal specificity (neuronal terminal identity) brings
the neuron to a functional state. It begins to form neu-
rites, synthesize neuropeptides, proteins necessary
for the production of neurotransmitters, receptors to
them, and components of ion channels. All these and
many other changes in the postmitotic neurons occur
due to unrelated regulatory proteins called terminal
selectors. One such selector in the nematode Caenor-
habditis elegans, Unc-3 (an ortholog of EBF/Olf/Collier),
determines the terminal differentiation of the cholin-
ergic motor neurons. The Unc-3 protein binds directly
to the cis-regulatory sites of acetylcholine biosynthesis
genes, ion channels, and numerous other genes. It does
not work alone but in cooperation with various Hox
proteins that act as its cofactors. Different Hox proteins
(depending on the site of the body) determine differ-
ences in the number and length of neurites, synaptic
connections, and electrical activity of the motor neu-
rons. This general scheme is also valid for other types
of neurons (sensory, motor, and intermediate) with
other terminal selectors [37]. Itis important that the
Hox proteins are needed by the worm not only for the
correct tuning of the neuron at the time of its termi-
nal differentiation but also for its further work. It has
been shown that the Hox protein of C. elegans Lin-39
(PG 4/5) is necessary in adulthood to maintain the ter-
minal specificity of the motor neurons [38, 39].
C. elegans is a simply organized animal. An adult
worm (hermaphrodite) has only 302 mature neurons[40].
An adult Drosophila brain contains about 200,000 neu-
rons [41], but their differentiation, targeting, and num-
ber of synapses are determined by similar processes.
The fly neuroblasts acquire unique fates under the
action of Hox proteins [42], and, more surprisingly,
neuromuscular synapse formation is under the Hox
control [31, 43]. It is hypothesized that the assembly of
synaptic contact between a neuron and a muscle cell is
possible if they express the same Hox protein (or set of
Hox proteins). This is true for at least one model sys-
tem, where the Hox protein Dfd (PG4) directly turns
on the expression of ankyrin (Ankyrin2-XL; synaptic
protein) and turns off the expression of Con (an adhe-
sive protein that selectively works in neuromuscular
synapses of another type) in the motor neurons and
muscles they innervate [43].
Finally, a type of neurons (leucokinergic neurons)
has been described in Drosophila for which Hox pro-
teins from the BX-C complex [Ubx (ultrabithorax),
abd-A (abdominalA), Abd-B (abdominalB)] are direct
terminal selectors because they turn on (Ubx, abd-A)
and turn off (Abd-B) synthesis of the neuropeptide leu-
cokinin [44].
Without going into details, it should be noted that
mammals (mouse, human) have fundamentally similar
rules for establishing proneural territories, differenti-
ation of neurons, and their postmitotic settings under
the control of Hox genes [30]. Mammals have also been
shown to have Hox proteins that function as terminal
selectors of motor neurons [45] and are required in
adulthood. Hindbrain development is controlled by 24
Hox genes, and they continue to work in the mature
brains of adult mice, while forebrain development in
vertebrates is a Hox independent process [46].
This is all the more surprising that the Hox genes
from several anterior paralogical groups (PG1,3-5)
begin to be expressed in the postnatal neocortex and
thalamus of the mouse [46].
Thus, modern experimental data obtained with
different model animals lead us to the idea that the
ancestral function of Hox genes is the terminal dif-
ferentiation of neurons, most likely motor neurons.
Thishypothesis has a strong theoretical and evidence
base. Firstly, it has recently become known that the
homeobox-containing factors in general tend to trigger
and maintain neurogenic differentiation. The C. ele-
gans genome encodes 102 homeodomain-containing
proteins from different families that selectively and
combinatorially function as terminal selectors or their
partners in the mature neurons [47, 48]. Therefore, the
specification of neurons using Hox proteins is a special
case of a general principle.
Secondly, in the GRNs, direct linkage between
high-level regulatory genes and terminal differentia-
tion genes may indicate the ancestral state of the sys-
tem. Within the GRNs master genes, target genes and
KULAKOVA et al.990
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
intermediate regulatory genes are distinguished. These
intermediate genes form a complex architecture of the
developmental GRNs, and according to the hypothesis
of intercalary evolution, they are the result of evolu-
tionary intercalations (insertions) between the master
gene and its target. Such as in the case of the homeo-
box gene Pax6 and the light-sensitive transmembrane
protein rhodopsin [49-51]. The Hox genes are univer-
sally implicated in the establishment and maintenance
of the terminal neuronal specification in Protostomia
and Deuterostomia, allowing for the existence of a sim-
ple organized ancestor of all Nephrozoa, which used
Hox genes for the same purpose. Initially, the simple
GRNs of such ancestor gradually and independently be-
came more complex in different evolutionary lineages
due to the involvement of the new clade-specific genes
under the control of the Hox cluster. In the course of
evolution, heterochronies shifted the activity onset of
all participants to earlier developmental stages and
brought their expression to the canonical state. It is
possible that in the first stages of this evolutionary pro-
cess, Hox genes coordinated the formation of synapses
between the motor neurons and muscles. Therefore,
the general principle of colinear transcription of Hox
genes relates Drosophila and mammals at the level of
two germ layers, ectodermal and mesodermal. Impor-
tantly, the principle of intercalary evolution allows the
GRN growth by duplication of the master gene and sub-
functionalization of the descendant genes with partial
preservation of the ancestral function [51].
This attractive hypothesis has internal contradic-
tions. Firstly, the physical linkage of terminal selectors
is not necessary to specify neurons. Most homeobox
genes that create the neural code in nematodes are not
clustered [47, 48]. Secondly, the level of complexity of
the last common ancestor of all Nephrozoa remains
in question because its Hox cluster already consisted
of at least 7 or 8 genes– five anterior (PG1-5), one or
two middle (PG6/8), and one posterior (PG9/14) [9]. It is
known that functions of the genes from different paral-
ogous groups overlap significantly [52], which means
that the ancestral cluster formed very quickly before
its members began to differ greatly in the spectra of
their targets. If the quantitative information realized
by the Hox proteins was at some point more important
than the qualitative (paralog-specific) information, this
could have pushed the Hox cluster to a rapid structur-
al expansion with minimal divergence of participants.
However, the paralog-specific functions began to ap-
pear later. This explanation looks logical but raises the
following questions: why are the Hox proteins from
different paralogous groups qualitatively important
for the neuron specification, and why do these proteins
differ structurally, while some of their paralog-specif-
ic functions are conserved (common to Nephrozoa)?
It seems that selection drove the evolution of the Hox
cluster in several directions at once, and this could be
explained by the scenario when the neurogenic func-
tion was not the only one. Whether this is true or not,
we could figure it out by referring to basal taxa.
Outside the Nephrozoa group, expression of the
Hox genes has been studied less, but it is known that
small and disjointed clusters of Acoelomorpha (a sister
branch of Nephrozoa, previously classified as a clade
within flatworms) operate in the nervous, muscular,
and reproductive systems [53-56]. In a single study
on embryos [53], three Hox genes of Convolutriloba
longifissura (PG1, PG5, and PG9/14) colinearly turn on
in the proneural territories shortly after gastrulation.
Two of these three genes operate in parenchymatous
internal domains slightly later or simultaneously with
thisevent.
The cnidarian Hox genes have been extensively
studied [57-59], and expression of some of them can
be associated with the nervous system, e.g., Hox1 (PG1)
of Clytia hemisphaerica functions in statocysts and
Anthox1 (PG9-like) in the apical tuft of the Nematostella
vectensis planula larva. However, in comparison with
the diverse neural differentiation involving other ho-
meobox genes in cnidarians, this is a very modest
outcome [60, 61]. Surprisingly, many cnidarian neu-
rotransmitters (including acetylcholine) and enzymes
of their biogenesis are synthesized not in neurons but
in gastrodermal cells [62]. Analyzing expression data
is also difficult because direct correspondence of the
genes from Hox/ParaHox classes in cnidarians and bi-
laterians is not obvious due to the high divergence or
loss of orthologs [63]. However, there are no direct reg-
ulators of neurogenesis among the genes, which exact-
ly match the category of “PG1-like” or “PG2-like”, but
there are genes with broad expression domains at the
level of ectoderm and endomesoderm.
Thus, before the origin of the last common ances-
tor of Nephrozoa, at the level of Acoelomorpha, Hox
genes were already engaged in several different de-
velopmental programs. Their functions in cnidarians
are diverse and not associated with terminal neuronal
specification. The path from the common ancestor of
cnidarians and bilaterians to modern Nephrozoa was
accompanied by structural expansion of the Hox clus-
ter and complex rearrangements of the regulatory rela-
tionships between the ancient developmental programs
hidden to us. Some of these programs may have relied
on the shared paralog-nonspecific functions of Hox
genes, which are realized in isolation from the spa-
tial colinear transcription. We hypothesize that these
functions have remained in modern animals, and to
investigate them we need to look at the general biolog-
ical processes in which Hox genes are involved. There
are several examples where the paralog-nonspecific
function of Hox genes is realized particularly clearly.
Wewill discuss them in the following sections.
IRREDUCIBLE COMPLEXITY OF HOX GENE 991
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
HOX PROTEINS CONTROL DEVELOPMENTAL
TIMING THROUGH AUTOPHAGY
Autophagy is a cellular degradation process neces-
sary for maintaining cell homeostasis and renewing its
cytoplasmic components. It is highly conserved, and its
influence on various biological functions has been de-
scribed in a wide range of organisms, from plants and
yeast to humans [64]. For bilaterians, autophagy is an
important tool in early development, as it participates
in cellular differentiation and tissue remodeling [65,
66]. For instance, in the Drosophila larvae, autophagy
activity is very high in the fat body during the wander-
ing L3 (L3W) stage, when the larva is rapidly growing
and undergoing metamorphosis, but not in the younger
feeding L3 (L3F) stage. It has been shown that the tran-
sition from the L3F to L3W stage is controlled by ecdys-
one, and the main regulators of autophagy in this case
are Hox proteins, which suppress premature autopha-
gy at the L3F stage [67]. During normal development,
colocalization of Hox proteins from multiple paralo-
gous groups is observed in the fat body of L3F larvae,
but it is not colinear. Here, the Hox proteins suppress
the expression of the atg genes (18 genes), which are
responsible for autophagy. It has been shown that in-
activation of the individual Hox genes (Dfd, Scr, Ubx,
abd-A, AbdB) does not lead to premature induction of
autophagy. Only the simultaneous shutdown of all Hox
paralogs in the L3F larval experiment initiates this
process [67]. Conversely, prolonged expression of the
investigated Hox genes inhibits autophagy in the lar-
val fat body cells. Such animals enter the wandering
stage 6-7 days later than the controls, indicating that
the forced maintenance of the Hox gene expression de-
lays development in Drosophila.
Thus, in the larval fat body, the universal activi-
ty of Hox proteins carries temporal rather than spa-
tial information, regulating the onset of autophagy at
the required stage of development. It is worth noting
that in the culture of mammalian fibroblasts, HoxB8
and HoxA9 also inhibit autophagy. The same proteins
exhibit a similar effect on the Drosophila larvae after
transgenesis [67]. These preliminary studies do not rule
out the paralog-nonspecific involvement of Hox genes
in autophagy control in the last common ancestor of
insects and vertebrates but require additional analysis
across a wide range of models.
WHY DO HOX GENES WORK
BEFORE DIFFERENTIATION BEGINS?
In multicellular animals, Hox genes are not ex-
pressed in totipotent and pluripotent cells because this
expression induces differentiation. In mammalian em-
bryonic stem cells, Hox loci have an ambivalent epigen-
etic status. Their histone code contains both repressive
and permissive tags [68]. These cells do not express
Hox genes, but expression can start rapidly in case of
additional permissive signals, which will lead the cells
to the beginning of the differentiation path.
Despite the blockage of Hox genes activity in toti-
potent cells, their maternal transcripts have been found
in oocytes of mammals (mouse, cow, and human [69],
amphibians (Xenopus laevis [70]), annelids (Platynereis
dumerilii[71]), myriapods (Strigamia maritima[72];
Trigo niulus corallinus [73], hymenopterans (ants of the
tribe Camponotini [74]), decapods (Macrobrachium olfer-
sii [75]), and even in oocytes of hydroid polyp Clitya
hemisphaerica [58]. In all animals, except Xenopus lae-
vis, the genes from several paralogous groups are ex-
pressed in oocytes.
The structure of oocyte transcripts of Hox genes
may provide clues about their functions. The exam-
ple of centipede Strigamia maritime (Chilopoda) [72]
showed that the maternal RNAs of Hox genes are
polyadenylated, but some of them do not contain an
open reading frame. Possibly, part of the maternal
RNAs of the centipede belongs to the class of regulatory
(protein-noncoding) RNAs. On the other hand, the Hox
gene transcripts in mammalian oocytes are deadenylat-
ed [69]. This is reasonable if such maternal transcripts
are required for later developmental stages and stored
in a stable (non-translated) form [76]. In addition, the
HOXB9 protein was detected in the nuclei of oocytes
and the cells of early mammalian (mouse, cow) embry-
os [77], and the Ubx and AbdA proteins were found in
ant oocytes [74].
The function of Hox genes in oocytes differs from
the canonical one since transcripts of different paral-
ogous groups are localized in a single cell. The wide-
spread nature of this phenomenon suggests that it is
not random. So far, there are no successful experiments
unambiguously indicating the functions of the Hox
gene transcripts in oocytes, but several hypotheses can
be put forward.
Most of the maternal transcripts of the Hox genes
may not be translated. They may represent an element
of epigenetic tuning of the zygotic genome. It is known
that the non-coding RNAs often function as scaffolds
for the assembly of chromatin remodeling proteins,
targeting them to subordinate loci. This function has
been described for the regulatory RNAs that are read
from the vertebrate Hox clusters [78] and it is consis-
tent with the presence of transcripts without an open
reading frame in the Strigamia oocytes. The importance
of the non-coding RNAs in the first cleavage divisions
was shown in the mouse embryos [79].
On the other hand, the mRNAs of Hox genes
can be translated. For example, the HoxB9 protein
is present in the nuclei of oocytes (both mature and
immature) and the cell nuclei of early embryos [77].
KULAKOVA et al.992
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
This doesnotexclude an early function of Hox genes
directed to oogenesis. It is known that in the mouse
several transcripts of Hox genes and their cofactors are
already present at the stage of the growing oocyte [80].
Interestingly, the homeodomain protein Nobox from the
family of the same name, which is close to both the Antp
and PRD classes [81], is present in the mouse oocytes
and regulates the functioning of the genes important
for oogenesis [82]. It cannot be ruled out that the oocyte
RNAs and proteins of the Hox genes are required for the
transcriptional control of the early zygotic genes [83].
The Hox proteins can be not only transcriptional
regulators, but also regulators of the cell cycle, RNA
splicing, DNA replication and repair [84-86]. For exam-
ple, some Hox proteins have recognition sites for the
serine/threonine kinase ATM [87], whose role in the
DNA double-strand break repair is widely known [88].
At least one vertebrate Hox protein, HoxB7, has been
experimentally shown to be involved in this process
[85]. In the mammary epithelial cells, HoxB7 increas-
es the probability of nonhomologous DNA end joining
by binding to the complex of Ku70/80 heterodimers
(proteins that recognize double-stranded breaks) and
DNA protein kinase [85]. HoxB7 expression stimulates
the DNA-protein kinase activity, which correlates with
the repair efficiency, whereas this effect is lost when
HoxB7 is knocked out. In addition, Hox proteins pro-
mote the assembly of pre-replicative complexes. For
example, HoxD13 and HoxC13 in different cell cultures
interact through homeodomains with the proteins
ofORC and Cdc6 pre-replicative complexes [86, 89, 90].
Although the mentioned functions of Hox proteins were
not described in embryogenesis, we assume that they
can still participate in the first stages of the develop-
ment of multicellular animals. In the early develop-
ment, synchronous divisions of blastomeres occur with
a minimum interval (there are no G1 and G2 phases of
the cycle), therefore, a very precise adjustment of the
molecular machinery of the oocyte is necessary for the
successful completion of the cleavage stage. In this case,
Hox proteins in oocytes may accelerate the assembly of
pre-replicative complexes and enhance DNA repair to
maintain the integrity of the embryo’s genetic material.
Undoubtedly, the role of Hox genes in oogenesis
will not be clarified without functional tests on a wide
range of models. It cannot be ruled out that the oocyte
RNAs and proteins of the Hox genes could be a “tran-
scriptional noise” or by-products of the previous stages
of oogenesis [91].
DOSE-DEPENDENT FUNCTIONS
OF HOX PROTEINS
There are dose-dependent functions of Hox pro-
teins when their concentration determines the mor-
phology of the anlage [92]. In mammals, Hox proteins
specify vertebral morphology in a dose-dependent
manner [93] and set the number and length of dig-
its [94]. In both cases, a gradual decrease in the dose
of Hox proteins increases intensity of morphological
changes. For instance, when the dose of any protein
from the posterior paralogs of HoxA and HoxD clus-
ters (Hoxd11, Hoxd12, Hoxa13, Hoxd13) is gradually
decreased, the digits shorten linearly and paralog-
independently depending on the proportion of mutant
alleles of the Hox genes [94].
In invertebrate animals, the most obvious exam-
ple of a dose-dependent function of Hox proteins is the
regulation of wing morphology. This function has been
described in insects from different clades and, appar-
ently, is universal for the diversification of the wing
shape and size in the second (T2) and third (T3) thorac-
ic segments [92]. It involves the Antp and Ubx proteins
in its realization. In the wild-type Drosophila, the Antp
protein is present only in T2 and the Ubx protein in
T3, with a lower concentration of Antp in T2 than of
Ubx in T3 (Fig.1a) [95]. In the Ubx
–/–
mutants, a pair
of wings is formed at T3 instead of halteres (Fig.1b).
If the dose of Antp in T3 is increased to the level of
Ubx in such mutants, the normal phenotype is restored
(halteres are formed on T3 (Fig.1c) [95]. On the con-
trary, when the Ubx dose is decreased, wings grow in-
stead of halteres at T3 (Fig.1d) [95]. Likewise, when
the Antp dose is increased in T2, halteres grow instead
ofwings (Fig.1e)[95].
HOX FACTORS CAN BE SECRETED BY CELLS
The difficulty faced by a researcher who decides to
elucidate the ancestral function of the ANTP-class ho-
meobox genes is related to the fact that such function
was not originally a single function, at least at the level
of Metazoa. This follows from the arrangement of mul-
titasking ANTP-class homeodomain proteins (Fig. 2).
Atthe end of the last century, it was discovered that a
synthetic homeodomain protein of 60 amino acid res-
idues, repeating the sequence of the Drosophila Antp
homeodomain, can penetrate the membranes of rat
nerve cells without the mediation of any receptors.
After penetration, it is transported into the nucleus
and increases the level of differentiation of the re-
cipient cells [96]. Later it was found out that the nat-
ural homeodomain proteins Emx1, Emx2, Engrailed-2
(En2), Hoxa5, Hoxb4, Hoxc8, Knotted1, Otx2, Pax6, and
Vax1 are present in the cells that do not express their
mRNA[97].
Apparently, most of the homeodomain proteins, in-
cluding those beyond the boundaries of the ANTP class
[97-100], have the ability for intercellular transport like
a signaling or morphogen molecule. The mechanism
IRREDUCIBLE COMPLEXITY OF HOX GENE 993
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
Fig. 1. Effect of Antp and Ubx protein dose on the Drosophila phenotype. a)“Wild-type” (WT); b)Ubx mutant –/–; c)Antp dosage
increase in T3 to the Ubx level restores normal phenotype; d)decrease of Ubx dosage in T3 to the Antp level in T2 leads to wing
formation; e) Antp dosage increase in T2 to the Ubx level in T3 leads to halteres formation; f) hypothetical phenotype that
could be modeled on Drosophila. It is characteristic of insects from the order Strepsiptera. The illustration is based on the data
ofPauletal. (2021) and Merabet and Carnesecchi (2024) [92, 95].
bywhich this is realized is not yet fully understood.
Itis known that secretion and internalization depend
on two overlapping motifs localized in the most con-
served regions of the homeodomain [100]. In addition,
secretion depends on individual hydrophobic amino
acids outside of the homeodomain (Fig.2) [97]. Itseems
that homeodomain proteins can enter any cell type
through macropinocytosis, but the efficiency of the
process depends on the structure of the glycocalyx of
the receiving cell [97, 101].
In the impressive study conducted in 2019 [97], 162
human homeodomain proteins from different classes
were tested for their secretion and transfer abilities,
and the test was performed simultaneously on three
different cell cultures (secretion – HEK 293T, GT1-7,
and MDCK; internalization – HeLa). It was shown
that secretion efficiency strongly depends on the cell
type and characteristics of the primary sequence of
the homeodomain proteins themselves. For example,
the proteins EN2, HOXC8, PAX6, and VAX1 were secreted
KULAKOVA et al.994
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
Fig. 2. Structural motifs of proteins from the ANTP class and their participation in the realization of the Hox protein functions.
The scheme describes the functional significance of the main domains of Hox factors. The homeodomain, the most conserved
region of ANTP-class proteins, is required for Hox factors to realize both canonical (transcription) and non-canonical functions.
In addition to the homeodomain, a high degree of conservatism is also characteristic of the short hexapeptide motif by which
Hox proteins interact with cofactors [102]. Surprisingly, the transport capacity of the Hox factors is also mediated by these do-
mains: the homeodomain contains secretion and internalizing (penetratin) motifs, and the conserved tryptophan residue in
the hexapeptide is required for Hox proteins to be exported from the nucleus [103]. For some Hox factors, it has been shown
that their secretion could depend on individual hydrophobic amino acids localized at less conserved sites on the protein [97].
Colored letters in the sequences indicate conserved positions, colored boxes surround motifs, non-canonical functions are in
bold; Mm,Mus musculus; Dm,Drosophila melanogaster; aa,amino acid residues.
in all three cell cultures, while HOXA5 and OTX2 were
secreted in only two. Ten proteins, including only
one Hox (HOXA10), were not secreted at all, which,
however, does not exclude this possibility in other
cell types. All tested proteins (including non-secreted
ones) turned out to be capable of internalization. This
means that secretion of the homeodomain proteins is
the crucial stage of transduction at which cells control
this process.
Homeodomain proteins can diffuse over two or
three-cell diameters like paracrine signaling factors,
but there are examples of their extensive diffusion.
In the mouse, the Otx2 protein is synthesized in the
vascular plexus, secreted into the cerebrospinal fluid,
and accumulates throughout the cerebral cortex [104].
Importantly, the endogenous Hox proteins prefer-
entially function as transcription factors. Immediately
after translation, they are taken up by karyopherins,
which recognize the nuclear localization signal, and
transport Hox proteins to the nucleus. The exogenous
Hox proteins exhibit a broader range of functions.
Ithas been shown that the secreted mouse Otx2 moves
into mitochondria, where it binds to mitochondrial ATP
synthases and enhances ATP synthesis [105]. Earlier
it was reported that the exogenous En2 in Xenopus
accumulates in the growth cones of neurons, controls
their axonal targeting, and indirectly enhances trans-
lation [106].
This does not exclude penetration of the exoge-
nous proteins into the nuclei, where they trigger tran-
scription of their mRNAs and other specific targets.
A classic example of this type is vertical signaling
during gastrulation in Xenopus. It was shown that the
Hox proteins from presomitic mesoderm sequentially
switch on the expression of their own Hox genes in the
gastrula neuroectoderm, i.e., copying of positional in-
formation from one germ layer to another occurs in
this case [104]. Direct exchange of transcription factors
between the cell layers coordinates the operation of de-
velopmental programs without the mediation of mor-
phogens and signaling cascades.
Remarkably, the signaling and regulatory func-
tions in general cases are mediated by the same evolu-
tionarily ancient and conserved motif– homeodomain.
The ability to solve two tasks with a single tool could
have been used to coordinate growth and development
by the first Metazoa, even before the emergence of
modern relationships between the long-range signaling
ligands, their messengers, and targets [104].
HOW DID NON-CANONICAL FUNCTIONS
BRING HOX CLUSTER ACTIVITY
TO A CANONICAL STATE?
The coherent operation of Hox genes required for
canonical function remains a great evolutionary mys-
tery because it is a multi-event process. It consists of:
– epigenetic tuning of Hox loci, including through
the regulatory RNAs encoded in the Hox clusters;
– establishment of the topology-associated do-
mains (TADs), which are stabilized according
tothe position of cells along the anteroposterior
axis;
– responses to multidirectional signals from mor-
phogens (retinoic acid, Wnt, Fgf, Bmp) in three-
dimensional coordinates of the embryo;
– coordinated expression in cells of different germ
layers due to the mechanism unique for home-
odomain proteins – vertical signaling;
IRREDUCIBLE COMPLEXITY OF HOX GENE 995
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
Fig. 3. Control of Hox cluster transcription. a)Some control mechanisms of the Hox cluster in bilateral animals. 1)GCR, Global
controlling element. Regulatory elements of this type have been described in vertebrates; 2)local cis-regulatory modules (in-
dividual and shared), site-specific transcription factors, and microRNAs (miR) encoded in Hox cluster sequences. These modes
ofregulation are present in protostomes and deuterostomes; 3)vertical signaling has been described in vertebrates. b)Non-ca-
nonical functions of Hox genes that could be realized by separate elements of the controls; c)Simplified scheme of Hox clus-
ter regulation during realization of the canonical function of axial patterning(d). Not all mechanisms are represented in the
scheme, and not all of those represented are universal. GCR and vertical signaling are not found in the model arthropods. Spiral
animals are generally poorly studied at the level of Hox cluster activation mechanisms, but among them, there are organisms
with temporal colinearity and early mesodermal transcription. RA (Retinoic Acid), FGF, and WNT are gradients of morphogens.
– response to individual signals from upstream
regulatory proteins, which can turn on/off indi-
vidual genes;
– reciprocal interactions of Hox genes (posterior
suppression and beyond).
Some of the regulatory mechanisms are shown
in Fig.3. How could this very complex picture emerge
from non-canonical functions? Perhaps the different
modes of Hox cluster regulation needed to perform
separate tasks were co-opted into a new program.
It seems intuitively correct to assume that this new
program is gastrulation in its “bilateral version.” This
seems to be suggested by the temporal conjugation of
the Hox cluster activation and gastrulation processes in
deuterostomes. Moreover, there is evidence for pre-ad-
aptation– two Hox genes of the cnidarian Nematostella
vectensis (anterior NvAx6 and mid-posterior NvAx1)
are important for gastrulation and specification of the
oral-aboral axis. Their expression sites label the oral
and aboral poles, and the morpholino knockdown sup-
presses gastrulation [107]. However, these genes do not
form a cluster and operate in different germ layers.
InDrosophila, the axial pattern of Hox genes is estab-
lished before the onset of gastrulation, and there are no
studies to reliably confirm or refute the involvement of
the Hox genes in arthropods in gastrulation. Insome
spiral animals (annelids, brachiopods, mollusks), early
activation of the Hox genes coincides with the onset
or continuation of gastrulation, but functional tests are
still lacking [26, 108, 109].
With a high degree of confidence, it can be stat-
ed that the last common ancestor of bilateral animals
KULAKOVA et al.996
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
hada single Hox cluster. This cluster could have been
used for different tasks, such as the specification of mo-
tor neurons, the establishment of neuromuscular junc-
tions, and a function requiring quantitative changes in
transcripts in response to a stimulus (probably a mor-
phogen concentration). Such an ancient function may
have been related to autophagy, control of prolifera-
tion, or gametogenesis. Importantly, this function was
enabled by the gradient distribution of Hox proteins
along the axis and kept the Hox genes in the cluster.
The Nephrozoa ancestor could use the whole cluster
or some of its genes to control gastrulation, but this is
not the case in the lineage of modern Acoelomorpha
(sister branch of Nephrozoa) because all Hox genes of
Convolutriloba are turned on after gastrulation [53].
Different ways of controlling the same cluster
could lead to errors in its regulation, and some of the
resulting aberrant variants were preserved by natural
selection. Two important events could have occurred in
the evolution of bilateral animals from the Nephrozoa
lineage. Firstly, several mechanisms controlling the
Hox cluster could have united to control a single mor-
phogenetic program. Secondly, there might have been
aheterochronic shift in the activation of this program
towards earlier development. The least catastrophic
variant suggests a series of heterochronic shifts of co-
linear expression of the Hox genes in internal, meso-
dermal in origin, structures up to the gastrula stage.
Then, through the vertical signaling, the Hox genes
began to turn on colinearly in the adjacent ectoderm
(future neuroectoderm), providing animals with a
new powerful tool for controlling early development.
Thistool could be easily scaled by the gradients of mor-
phogens, it coordinated the development of ectodermal
and mesodermal tissues, and it was evolutionarily plas-
tic due to many controls coming from the older pro-
grams. Perhaps it was this new molecular mechanism
that “detonated” and triggered the “Cambrian Explo-
sion” because of its ability to rapidly alter early devel-
opment.
CONCLUSION
The diversity of non-canonical functions of Hox
genes is determined by the structure of the home-
odomain protein itself, which can work not only as a
transcription factor, but also as a regulator of general
biological processes, such as DNA repair, replication,
translation, and RNA splicing.
The “hourglass” model, while illustrative, leaves
non-canonical functions of the Hox genes invisible.
According to the inverse hourglass model, which is
valid for Metazoa as a whole [110], there is a fundamen-
tal similarity in gene functioning at the earliest stages
of development (pluripotent state of cells, cleavage)
and at later stages (differentiation, organogenesis).
However, animals from different phyla will vary sig-
nificantly in the ensembles of regulatory genes and
nature of their involvement in morphogenesis in the
middle of development, just between cleavage and
committed differentiation [110]. These are the differ-
ences that define fundamental distinction between
the phyla within Metazoa. In other words, it is possi-
ble to identify distinct sets of signaling pathways and
transcription factors that interact during the estab-
lishment of Metazoa organization plans. Their specific
combination defines the appearance of each phylum.
Itturned out that homeobox genes in general and Hox
genes in particular do not fall into the category of such
“phylum- specific” regulators because their functions
are broader and more conserved during the divergent
period of development.
We assume that the proto-Hox gene initially pos-
sessed a wide repertoire of functions, some of which
relied on the signaling nature of its protein. Animals
from the ParaHoxozoa branch turned out to be heirs
of this regulatory complexity. They further enhanced it
through cooperation between developmental programs
that used different functional capabilities of Hox pro-
teins. These programs emerged at various stages of
evolution, and their traces are preserved in modern
animals in the form of distinct paralog-nonspecific and
dose-dependent functions.
Acknowledgments. In the study, Geneious®
2023.2.1 software was used for sequence analysis, with
access provided by the research resource center “Chro-
mas” of Saint Petersburg State University.
Contributions. M.A.K. developed the study con-
cept, supervised the study, and prepared and edited
the manuscript; G.P.M. prepared and edited the manu-
script; L.O.P. prepared and edited the manuscript.
Funding. This work was financially supported by
the Russian Science Foundation (project no.23-24-00426).
Ethics declarations. This work does not contain
any studies involving human and animal subjects.
Theauthors of this work declare that they have nocon-
flicts of interest.
Open access. This article is licensed under a Cre-
ative Commons Attribution4.0 International License,
which permits use, sharing, adaptation, distribution,
and reproduction in any medium or format, as long
as you give appropriate credit to the original author(s)
and the source, provide a link to the Creative Commons
license, and indicate if changes were made. Theimages
or other third-party material in this article are includ-
ed in the article’s Creative Commons license, unless
indicated otherwise in a credit line to the material.
Ifmaterial is not included in the article’s Creative Com-
mons license and your intended use is not permitted
by statutory regulation or exceeds the permitted use,
IRREDUCIBLE COMPLEXITY OF HOX GENE 997
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit
https://creativecommons.org/licenses/by/4.0/.
REFERENCES
1. Degnan, B. M., Vervoort, M., Larroux, C., and Rich-
ards, G. S. (2009) Early evolution of metazoan tran-
scription factors, Curr. Opin. Genet. Dev., 19, 591-599,
doi:10.1016/j.gde.2009.09.008.
2. Ryan, J.F., Pang,K., Mullikin, J.C., Martindale, M.Q.,
and Baxevanis, A.D. (2010) The homeodomain com-
plement of the ctenophore Mnemiopsis leidyi suggests
that Ctenophora and Porifera diverged prior to the
ParaHoxozoa, Evodevo, 1, 9, doi:10.1186/2041-9139-1-9.
3. Pang,K., and Martindale, M.Q. (2008) Developmen-
tal expression of homeobox genes in the ctenophore
Mnemiopsis leidyi, Dev. Genes Evol., 218, 307-319,
doi:10.1007/s00427-008-0222-3.
4. Srivastava, M., Begovic, E., Chapman, J., Putnam,
N.H., Hellsten,U., Kawashima,T., Kuo,A., Mitros,T.,
Salamov,A., and Carpenter, M.L. (2008) The Trichop-
lax genome and the nature of placozoans, Nature,
454, 955-960, doi:10.1038/nature07191.
5. Pastrana, C.C., DeBiasse, M.B., and Ryan, J.F. (2019)
Sponges lack ParaHox genes, Genome Biol. Evol., 11,
1250-1257, doi:10.1093/gbe/evz052.
6. Lewis, E. B. (1978) A gene complex controlling seg-
mentation in Drosophila, Nature, 276, 565-570,
doi:10.1038/276565a0.
7. Korchagina, N. M., Bakalenko, N. I., and Kulakova,
M.A. (2010) Hox-cluster and evolution of morphogen-
eses, Russ.J. Dev. Biol., 41, 353-363.
8. Hubert, K. A., and Wellik, D. M. (2023) Hox genes
in development and beyond, Development, 150,
dev192476, doi:10.1242/dev.192476.
9. De Rosa, R., Grenier, J. K., Andreeva, T., Cook, C. E.,
Adoutte,A., Akam,M., Carroll, S.B., and Balavoine,G.
(1999) Hox genes in brachiopods and priapulids
and protostome evolution, Nature, 399, 772-776,
doi:10.1038/21631.
10. Balavoine, G., de Rosa, R., and Adoutte, A. (2002)
Hox clusters and bilaterian phylogeny, Mol. Phy-
logenet. Evol., 24, 366-373, doi: 10.1016/s1055-7903
(02)00237-3.
11. Dressler, G. R., and Gruss,P. (1989) Anterior bound-
aries of Hox gene expression in mesoderm-derived
structures correlate with the linear gene order
along the chromosome, Differentiation, 41, 193-201,
doi:10.1111/j.1432-0436.1989.tb00747.x.
12. Duboule, D., and Dolle, P. (1989) The structural and
functional organization of the murine HOX gene
family resembles that of Drosophila homeotic genes,
EMBOJ., 8, 1497-1505, doi:10.1002/j.1460-2075.1989.
tb03534.x.
13. Duboule, D. (1994) Temporal colinearity and the
phylotypic progression: a basis for the stability of a
vertebrate Bauplan and the evolution of morpholo-
gies through heterochrony, Dev. Suppl., doi:10.1242/
dev.1994.Supplement.135.
14. Larroux, C., Fahey, B., Degnan, S. M., Adamski, M.,
Rokhsar, D.S., and Degnan, B. M. (2007) The NK ho-
meobox gene cluster predates the origin of Hox genes,
Curr. Biol., 17, 706-710, doi:10.1016/j.cub.2007.03.008.
15. Copley, R. R. (2023) The ancestry of Antennapedia-
like homeobox genes, bioRxiv, doi: 10.1101/2023.
03.14.532566.
16. DuBuc, T. Q., Ryan, J. F., Shinzato, C., Satoh, N., and
Martindale, M. Q. (2012) Coral comparative genom-
ics reveal expanded Hox cluster in the cnidarian-
bilaterian ancestor, Integr. Comp. Biol., 52, 835-841,
doi:10.1093/icb/ics098.
17. Lynch, V. J., Roth, J. J., and Wagner, G. P. (2006)
Adaptive evolution of Hox-gene homeodomains
after cluster duplications, BMC Evol. Biol., 6, 86,
doi:10.1186/1471-2148-6-86.
18. Duboule, D. (2007) The rise and fall of Hox gene
clusters, Development, 134, 2549-2560, doi: 10.1242/
dev.001065.
19. Seo, H. C., Edvardsen, R. B., Maeland, A. D., Bjord-
al, M., Jensen, M. F., Hansen, A., Flaat, M., Weissen-
bach, J., Lehrach,H., Wincker, P., Reinhardt, R., and
Chourrout,D. (2004) Hox cluster disintegration with
persistent anteroposterior order of expression in
Oikopleura dioica, Nature, 431, 67-71, doi: 10.1038/
nature02709.
20. Fröbius, A. C., and Funch, P. (2017) Rotiferan Hox
genes give new insights into the evolution of meta-
zoan bodyplans, Nat. Commun., 8, 9, doi: 10.1038/
s41467-017-00020-w.
21. Lee, P. N., Callaerts, P., De Couet, H. G., and Mar-
tindale, M. Q. (2003) Cephalopod Hox genes and
the origin of morphological novelties, Nature, 424,
1061-1065, doi:10.1038/nature01872.
22. Monteiro, A. S., and Ferrier, D. E. (2006) Hox genes
are not always Colinear, Int. J. Biol. Sci., 2, 95-103,
doi:10.7150/ijbs.2.95.
23. Stern, D. L. (1998) A role of Ultrabithorax in mor-
phological differences between Drosophila species,
Nature, 396, 463-466, doi:10.1038/24863.
24. Slack, J. M., Holland, P.W., and Graham, C.F. (1993)
The zootype and the phylotypic stage, Nature, 361,
490-492, doi:10.1038/361490a0.
25. Lemaire, L., and Kessel, M. (1997) Gastrulation and
homeobox genes in chick embryos, Mech. Dev.,
67, 3-16, doi:10.1016/s0925-4773(97)00102-0.
26. Kulakova,M., Bakalenko,N., Novikova,E., Cook, C.E.,
Eliseeva,E., Steinmetz, P.R., Kostyuchenko, R.P., Don-
dua,A., Arendt,D., Akam,M., and Andreeva,T. (2007)
Hox gene expression in larval development of the
polychaetes Nereis virens and Platynereis dumerilii
KULAKOVA et al.998
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
(Annelida, Lophotrochozoa), Dev. Genes Evol., 217,
39-54, doi:10.1007/s00427-006-0119-y.
27. Michaut,L., Jansen, H.J., Bardine,N., Durston, A.J.,
and Gehring, W. J. (2011) Analyzing the function
of a hox gene: an evolutionary approach, Dev.
Growth Differ., 53, 982-993, doi:10.1111/j.1440-169X.
2011.01307.x.
28. Fröbius, A. C., Matus, D. Q., and Seaver, E. C. (2008)
Genomic organization and expression demonstrate
spatial and temporal Hox gene colinearity in the
lophotrochozoan Capitella sp. I, PLoS One, 3, e4004,
doi:10.1371/journal.pone.0004004.
29. Garcia-Fernàndez, J. (2005) The genesis and evolu-
tion of homeobox gene clusters, Nat. Rev. Genet., 6,
881-892, doi:10.1038/nrg1723.
30. Philippidou, P., and Dasen, J. S. (2013) Hox genes:
choreographers in neural development, architects of
circuit organization, Neuron, 80, 12-34, doi:10.1016/
j.neuron.2013.09.020.
31. Joshi, R., Sipani, R., and Bakshi, A. (2021) Roles of
Drosophila Hox genes in the assembly of neuromus-
cular networks and behavior, Front. Cell Dev. Biol.,
9, 786993, doi:10.3389/fcell.2021.786993.
32. Feng,W., Li,Y., and Kratsios,P. (2021) Emerging roles
for hox proteins in the last steps of neuronal devel-
opment in worms, flies, and mice, Front. Neurosci.,
15, 801791, doi:10.3389/fnins.2021.801791.
33. Kourakis, M.J., Master, V.A., Lokhorst, D.K., Nardel-
li-Haefliger, D., Wedeen, C.J., Martindale, M.Q., and
Shankland,M. (1997) Conserved anterior boundaries
of Hox gene expression in the central nervous sys-
tem of the leech Helobdella, Dev. Biol., 190, 284-300,
doi:10.1006/dbio.1997.8689.
34. Samadi,L., and Steiner,G. (2010) Expression of Hox
genes during the larval development of the snail,
Gibbula varia (L.)-further evidence of non-colin-
earity in molluscs, Dev. Genes Evol., 220, 161-172,
doi:10.1007/s00427-010-0338-0.
35. Hinman, V. F., O’Brien, E. K., Richards, G. S., and
Degnan, B. M. (2003) Expression of anterior Hox
genes during larval development of the gastropod
Haliotis asinina, Evol. Dev., 5, 508-521, doi: 10.1046/
j.1525-142x.2003.03056.x.
36. Barrera Grijalba, C. C., Rodríguez Monje, S. V.,
Gestal,C., and Wollesen,T. (2023) Octopod Hox genes
and cephalopod plesiomorphies, Sci. Rep., 13, 15492,
doi:10.1038/s41598-023-42435-0.
37. Kratsios,P., Kerk, S.Y., Catela,C., Liang,J., Vidal,B.,
Bayer, E. A., Feng, W., De La Cruz, E. D., Croci, L.,
Consalez, G.G., Mizumoto, K., and Hobert,O. (2017)
An intersectional gene regulatory strategy defines
subclass diversity of C.elegans motor neurons, Elife,
6, e25751, doi:10.7554/eLife.25751.
38. Feng, W., Li, Y., Dao, P., Aburas, J., Islam, P., El-
baz, B., Kolarzyk, A., Brown, A. E., and Kratsios, P.
(2020) A terminal selector prevents a Hox tran-
scriptional switch to safeguard motor neuron iden-
tity throughout life, Elife, 9, e50065, doi: 10.7554/
eLife.50065.
39. Li,Y., Osuma,A., Correa,E., Okebalama, M.A., Dao,P.,
Gaylord, O., Aburas, J., Islam, P., Brown, A. E., and
Kratsios, P. (2020) Establishment and maintenance
of motor neuron identity via temporal modular-
ity in terminal selector function, Elife, 9, e59464,
doi:10.7554/eLife.59464.
40. White, J.G., Southgate,E., Thomson, J.N., and Bren-
ner,S. (1986) The structure of the nervous system of
the nematode Caenorhabditis elegans, Philos. Trans.
R Soc. Lond. B Biol. Sci., 314, 1-340, doi: 10.1098/
rstb.1986.0056.
41. Raji, J. I., and Potter, C. J. (2021) The number of
neurons in Drosophila and mosquito brains, PLoS
One, 16, e0250381, doi: 10.1371/journal.pone.
0250381.
42. Estacio-Gómez, A., and Díaz-Benjumea, F. J. (2014)
Roles of Hox genes in the patterning of the central
nervous system of Drosophila, Fly (Austin), 8, 26-32,
doi:10.4161/fly.27424.
43. Friedrich, J., Sorge, S., Bujupi, F., Eichenlaub, M. P.,
Schulz, N. G., Wittbrodt, J., and Lohmann, I. (2016)
Hox function is required for the development
and maintenance of the Drosophila feeding mo-
tor unit, Cell Rep., 14, 850-860, doi: 10.1016/j.celrep.
2015.12.077.
44. Estacio-Gómez, A., Moris-Sanz, M., Schäfer, A. K.,
Perea,D., Herrero,P., and Díaz-Benjumea, F.J. (2013)
Bithorax-complex genes sculpt the pattern of leucoki-
nergic neurons in the Drosophila central nervous
system, Development, 140, 2139-2148, doi: 10.1242/
dev.090423.
45. Catela,C., Chen,Y., Weng,Y., Wen,K., and Kratsios,P.
(2022) Control of spinal motor neuron terminal dif-
ferentiation through sustained Hoxc8 gene activity,
Elife, 11, e70766, doi:10.7554/eLife.70766.
46. Hutlet, B., Theys, N., Coste, C., Ahn, M. T., Doshish-
ti-Agolli, K., Lizen, B., and Gofflot, F. (2016) System-
atic expression analysis of Hox genes at adulthood
reveals novel patterns in the central nervous system,
Brain Struct. Funct., 221, 1223-1243, doi: 10.1007/
s00429-014-0965-8.
47. Reilly, M. B., Cros,C., Varol, E., Yemini, E., and Hob-
ert, O. (2020) Unique homeobox codes delineate all
the neuron classes of C.elegans, Nature, 584, 595-601,
doi:10.1038/s41586-020-2618-9.
48. Hobert, O. (2021) Homeobox genes and the specifi-
cation of neuronal identity, Nat. Rev. Neurosci., 22,
627-636, doi:10.1038/s41583-021-00497-x.
49. Sheng,G., Thouvenot,E., Schmucker,D., Wilson, D.S.,
and Desplan, C. (1997) Direct regulation of rhodop-
sin 1 by Pax-6/eyeless in Drosophila: evidence for a
conserved function in photoreceptors, Genes Dev.,
11, 1122-1131, doi:10.1101/gad.11.9.1122.
IRREDUCIBLE COMPLEXITY OF HOX GENE 999
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
50. Gehring, W. J., and Ikeo, K. (1999) Pax 6: mastering
eye morphogenesis and eye evolution, Trends Genet.,
15, 371-377, doi:10.1016/s0168-9525(99)01776-x.
51. Gehring, W.J. (2005) New perspectives on eye devel-
opment and the evolution of eyes and photorecep-
tors, J.Hered., 96, 171-184, doi:10.1093/jhered/esi027.
52. Merabet,S., and Mann, R.S. (2016) To be specific or
not: the critical relationship between Hox and TALE
proteins, Trends Genet., 32, 334-347, doi:10.1016/j.tig.
2016.03.004.
53. Hejnol,A., and Martindale, M.Q. (2009) Coordinated
spatial and temporal expression of Hox genes during
embryogenesis in the acoel Convolutriloba longifis-
sura, BMC Biol., 7, 65, doi:10.1186/1741-7007-7-65.
54. Moreno, E., Nadal, M., Baguñà, J., and Martínez, P.
(2009) Tracking the origins of the bilaterian Hox
patterning system: insights from the acoel flatworm
Symsagittifera roscoffensis, Evol. Dev., 11, 574-581,
doi:10.1111/j.1525-142X.2009.00363.x.
55. Moreno,E., De Mulder,K., Salvenmoser,W., Ladurn-
er,P., and Martínez,P. (2010) Inferring the ancestral
function of the posterior Hox gene within the bila-
teria: controlling the maintenance of reproductive
structures, the musculature and the nervous system
in the acoel flatworm Isodiametra pulchra, Evol. Dev.,
12, 258-266, doi:10.1111/j.1525-142X.2010.00411.x.
56. Moreno, E., Permanyer, J., and Martinez, P. (2011)
The origin of patterning systems in bilateria-insights
from the Hox and ParaHox genes in Acoelomorpha,
Genom. Proteom. Bioinform., 9, 65-76, doi: 10.1016/
s1672-0229(11)60010-7.
57. Finnerty, J. R., and Martindale, M.Q. (1999) Ancient
origins of axial patterning genes: Hox genes and
ParaHox genes in the Cnidaria, Evol. Dev., 1, 16-23,
doi:10.1046/j.1525-142x.1999.99010.x.
58. Chiori, R., Jager, M., Denker, E., Wincker, P., Da Sil-
va,C., Le Guyader,H., Manuel,M., and Quéinnec, E.
(2009) Are Hox genes ancestrally involved in axial
patterning? Evidence from the hydrozoan Clytia hemi-
sphaerica (Cnidaria), PLoS One, 4, e4231, doi:10.1371/
journal.pone.0004231.
59. Nong,W., Cao,J., Li,Y., Qu,Z., Sun,J., Swale,T., Yip,
H.Y., Qian, P.Y., Qiu, J.W., Kwan, H.S., Bendena,W.,
Tobe,S., Chan, T.F., Yip, K.Y., Chu, K.H., Ngai, S.M.,
Tsim, K.Y., Holland, P.W.H., and Hui, J.H.L. (2020)
Jellyfish genomes reveal distinct homeobox gene clus-
ters and conservation of small RNA processing, Nat.
Commun., 11, 3051, doi:10.1038/s41467-020-16801-9.
60. Galliot, B., Quiquand, M., Ghila, L., de Rosa, R., Mil-
jkovic-Licina, M., and Chera, S. (2009) Origins of
neurogenesis, a cnidarian view, Dev. Biol., 332, 2-24,
doi:10.1016/j.ydbio.2009.05.563.
61. Faltine-Gonzalez, D., Havrilak, J., and Layden, M. J.
(2023) The brain regulatory program predates cen-
tral nervous system evolution, Sci. Rep., 13, 8626,
doi:10.1038/s41598-023-35721-4.
62. Oren, M., Brickner, I., Appelbaum, L., and Levy, O.
(2014) Fast neurotransmission related genes are ex-
pressed in non nervous endoderm in the sea anem-
one Nematostella vectensis, PLoS One, 9, e93832,
doi:10.1371/journal.pone.0093832.
63. Steinworth, B. M., Martindale, M. Q., and Ryan, J. F.
(2023) Gene Loss may have shaped the Cnidarian and
Bilaterian Hox and ParaHox complement, Genome
Biol. Evol., 15, evac172, doi:10.1093/gbe/evac172.
64. Reggiori, F., and Klionsky, D. J. (2002) Autopha-
gy in the eukaryotic cell, Eukaryot. Cell, 1, 11-21,
doi:10.1128/ec.01.1.11-21.2002.
65. Wada, Y., Sun-Wada, G. H., Kawamura, N., and Aoy-
ama, M. (2014) Role of autophagy in embryogene-
sis, Curr. Opin. Genet. Dev., 27, 60-66, doi: 10.1016/
j.gde.2014.03.010.
66. Tsukamoto,S., Kuma,A., and Mizushima,N. (2008) The
role of autophagy during the oocyte-to-embryo transi-
tion, Autophagy, 4, 1076-1078, doi:10.4161/auto.7065.
67. Banreti,A., Hudry,B., Sass,M., Saurin, A.J., and Gra-
ba, Y. (2014) Hox proteins mediate developmental
and environmental control of autophagy, Dev. Cell,
28, 56-69, doi:10.1016/j.devcel.2013.11.024.
68. Sachs, M., Onodera, C., Blaschke, K., Ebata, K. T.,
Song, J. S., and Ramalho-Santos, M. (2013) Bivalent
chromatin marks developmental regulatory genes in
the mouse embryonic germline in vivo, Cell Rep., 3,
1777-1784, doi:10.1016/j.celrep.2013.04.032.
69. Paul, D., Bridoux, L., Rezsöhazy, R., and Donnay, I.
(2011) HOX genes are expressed in bovine and mouse
oocytes and early embryos, Mol. Reprod. Dev., 78,
436-449, doi:10.1002/mrd.21321.
70. Kondo,M., Yamamoto,T., Takahashi,S., and Taira,M.
(2017) Comprehensive analyses of hox gene expres-
sion in Xenopus laevis embryos and adult tissues, Dev.
Growth Differ., 59, 526-539, doi:10.1111/dgd.12382.
71. Maslakov, G.P., Kulishkin, N. S., Surkova, A.A., and
Kulakova, M. A. (2021) Maternal transcripts of Hox
genes are found in oocytes of Platynereis dumerilii
(Annelida, Nereididae), J.Dev. Biol., 9, 37, doi:10.3390/
jdb9030037.
72. Chipman, A.D., Ferrier, D.E., Brena,C., Qu,J., Hughes,
D.S., Schröder,R., Torres-Oliva,M., Znassi,N., Jiang,H.,
Almeida, F.C., Alonso, C.R., Apostolou,Z., Aqrawi,P.,
Arthur,W., Barna, J.C., Blankenburg, K.P., Brites,D.,
Capella-Gutiérrez,S., Coyle,M., Dearden, P. K., etal.
(2014) The first myriapod genome sequence reveals
conservative arthropod gene content and genome or-
ganisation in the centipede Strigamia maritima, PLoS
Biol., 12, e1002005, doi:10.1371/journal.pbio.1002005.
73. Qu, Z., Nong, W., So, W. L., Barton-Owen, T., Li, Y.,
Leung, T.C.N., Li,C., Baril,T., Wong, A.Y.P., Swale,T.,
Chan, T.F., Hayward,A., Ngai, S.M., and Hui, J.H.L.
(2020) Millipede genomes reveal unique adaptations
during myriapod evolution, PLoS Biol., 18, e3000636,
doi:10.1371/journal.pbio.3000636.
KULAKOVA et al.1000
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
74. Rafiqi, A.M., Rajakumar,A., and Abouheif,E. (2020)
Origin and elaboration of a major evolutionary
transition in individuality, Nature, 585, 239-244,
doi:10.1038/s41586-020-2653-6.
75. Jaramillo, M. L., Ammar, D., Quispe, R. L., Bonat-
to Paese, C. L., Gruendling, A. P., Müller, Y. M., and
Nazari, E.M. (2022) Identification of Hox genes and
their expression profiles during embryonic devel-
opment of the emerging model organism, Macro-
brachium olfersii, J.Exp. Zool. B Mol. Dev. Evol., 338,
292-300, doi:10.1002/jez.b.23118.
76. Ferreira, E.M., Vireque, A.A., Adona, P.R., Meirelles,
F. V., Ferriani, R. A., and Navarro, P. A. (2009) Cyto-
plasmic maturation of bovine oocytes: structural and
biochemical modifications and acquisition of devel-
opmental competence, Theriogenology, 71, 836-848,
doi:10.1016/j.theriogenology.2008.10.023.
77. Sauvegarde, C., Paul, D., Bridoux, L., Jouneau, A.,
Degrelle, S., Hue, I., Rezsohazy, R., and Donnay, I.
(2016) Dynamic pattern of HOXB9 protein localiza-
tion during oocyte maturation and early embryonic
development in mammals, PLoS One, 11, e0165898,
doi:10.1371/journal.pone.0165898.
78. Rinn, J. L., Kertesz, M., Wang, J. K., Squazzo, S. L.,
Xu, X., Brugmann, S. A., Goodnough, L. H., Helms,
J.A., Farnham, P.J., Segal,E., and Chang, H.Y. (2007)
Functional demarcation of active and silent chroma-
tin domains in human HOX loci by noncoding RNAs,
Cell, 129, 1311-1323, doi:10.1016/j.cell.2007.05.022.
79. Iyyappan,R., Aleshkina,D., Zhu,L., Jiang,Z., Kintero-
va, V., and Susor, A. (2021) Oocyte specific lncRNA
variant Rose influences oocyte and embryo develop-
ment, Noncoding RNA Res., 6, 107-113, doi: 10.1016/
j.ncrna.2021.06.001.
80. Kageyama,S., Gunji,W., Nakasato,M., Murakami,Y.,
Nagata,M., and Aoki,F. (2007) Analysis of transcrip-
tion factor expression during oogenesis and preim-
plantation development in mice, Zygote, 15, 117-128,
doi:10.1017/s096719940700411x.
81. Holland, P. W., Booth, H. A., and Bruford, E. A.
(2007) Classification and nomenclature of all hu-
man homeobox genes, BMC Biol., 5, 47, doi:10.1186/
1741-7007-5-47.
82. Rajkovic, A., Pangas, S. A., Ballow, D., Suzumori, N.,
and Matzuk, M. M. (2004) NOBOX deficiency dis-
rupts early folliculogenesis and oocyte-specific gene
expression, Science, 305, 1157-1159, doi: 10.1126/
science.1099755.
83. Jukam, D., Shariati, S. A. M., and Skotheim, J. M.
(2017) Zygotic genome activation in vertebrates,
Dev. Cell, 42, 316-332, doi:10.1016/j.devcel.2017.07.026.
84. Carnesecchi, J., Boumpas, P., van Nierop, Y. S. P.,
Domsch, K., Pinto, H. D., Borges Pinto, P., and
Lohmann, I. (2022) The Hox transcription factor
Ultrabithorax binds RNA and regulates co-transcrip-
tional splicing through an interplay with RNA poly-
meraseII, Nucleic Acids Res., 50, 763-783, doi:10.1093/
nar/gkab1250.
85. Rubin, E., Wu, X., Zhu, T., Cheung, J. C., Chen, H.,
Lorincz, A., Pandita, R. K., Sharma, G. G., Ha, H. C.,
Gasson,J., Hanakahi, L. A., Pandita, T. K., and Suku-
mar, S. (2007) A role for the HOXB7 homeodomain
protein in DNA repair, Cancer Res., 67, 1527-1535,
doi:10.1158/0008-5472.Can-06-4283.
86. Salsi, V., Ferrari, S., Ferraresi, R., Cossarizza, A.,
Grande,A., and Zappavigna,V. (2009) HOXD13 binds
DNA replication origins to promote origin licens-
ing and is inhibited by geminin, Mol. Cell Biol., 29,
5775-5788, doi:10.1128/MCB.00509-09.
87. Primon, M., Hunter, K. D., Pandha, H. S., and Mor-
gan, R. (2019) Kinase regulation of HOX transcrip-
tion factors, Cancers (Basel), 11, 508, doi: 10.3390/
cancers11040508.
88. Lavin, M. F., and Kozlov, S. (2007) ATM activation
and DNA damage response, Cell Cycle, 6, 931-942,
doi:10.4161/cc.6.8.4180.
89. Comelli, L., Marchetti, L., Arosio, D., Riva, S., Abdu-
rashidova,G., Beltram,F., and Falaschi,A. (2009) The
homeotic protein HOXC13 is a member of human
DNA replication complexes, Cell Cycle, 8, 454-459,
doi:10.4161/cc.8.3.7649.
90. Marchetti, L., Comelli, L., D’Innocenzo, B., Puzzi, L.,
Luin, S., Arosio, D., Calvello, M., Mendoza-Maldona-
do,R., Peverali,F., Trovato,F., Riva,S., Biamonti, G.,
Abdurashidova, G., Beltram, F., and Falaschi, A.
(2010) Homeotic proteins participate in the function
of human-DNA replication origins, Nucleic Acids Res.,
38, 8105-8119, doi:10.1093/nar/gkq688.
91. Vieux, K.F., and Clarke, H.J. (2018) CNOT6 regulates a
novel pattern of mRNA deadenylation during oocyte
meiotic maturation, Sci. Rep., 8, 6812, doi: 10.1038/
s41598-018-25187-0.
92. Merabet, S., and Carnesecchi, J. (2024) Hox dosage
and morphological diversification during develop-
ment and evolution, Semin. Cell Dev. Biol., 152-153,
70-75, doi:10.1016/j.semcdb.2022.11.009.
93. Horan, G. S., Ramírez-Solis, R., Featherstone, M. S.,
Wolgemuth, D. J., Bradley, A., and Behringer, R. R.
(1995) Compound mutants for the paralogous
hoxa-4, hoxb-4, and hoxd-4 genes show more com-
plete homeotic transformations and a dose-depen-
dent increase in the number of vertebrae trans-
formed, Genes Dev., 9, 1667-1677, doi: 10.1101/
gad.9.13.1667.
94. Zákány, J., Fromental-Ramain, C., Warot, X., and
Duboule, D. (1997) Regulation of number and size
of digits by posterior Hox genes: a dose-dependent
mechanism with potential evolutionary implica-
tions, Proc. Natl. Acad. Sci. USA, 94, 13695-13700,
doi:10.1073/pnas.94.25.13695.
95. Paul,R., Giraud,G., Domsch,K., Duffraisse,M., Mar-
migère, F., Khan, S., Vanderperre, S., Lohmann, I.,
IRREDUCIBLE COMPLEXITY OF HOX GENE 1001
BIOCHEMISTRY (Moscow) Vol. 89 No. 6 2024
Stoks, R., Shashidhara, L. S., and Merabet, S. (2021)
Hox dosage contributes to flight appendage mor-
phology in Drosophila, Nat. Commun., 12, 2892,
doi:10.1038/s41467-021-23293-8.
96. Joliot, A., Pernelle,C., Deagostini-Bazin, H., and Pro-
chiantz, A. (1991) Antennapedia homeobox peptide
regulates neural morphogenesis, Proc. Natl. Acad. Sci.
USA, 88, 1864-1868, doi:10.1073/pnas.88.5.1864.
97. Lee, E.J., Kim,N., Park, J.W., Kang, K.H., Kim, W.I.,
Sim, N.S., Jeong, C.S., Blackshaw,S., Vidal,M., Huh,
S. O., Kim, D., Lee, J.H., and Kim, J. W. (2019) Glob-
al analysis of intercellular homeodomain protein
transfer, Cell Rep., 28, 712-722.e713, doi: 10.1016/
j.celrep.2019.06.056.
98. Lucas, W. J., Bouché-Pillon, S., Jackson, D. P., Nguy-
en, L., Baker, L., Ding, B., and Hake, S. (1995) Selec-
tive trafficking of KNOTTED1 homeodomain protein
and its mRNA through plasmodesmata, Science, 270,
1980-1983, doi:10.1126/science.270.5244.1980.
99. Prochiantz, A., and Di Nardo, A. A. (2015) Homeo-
protein signaling in the developing and adult ner-
vous system, Neuron, 85, 911-925, doi: 10.1016/
j.neuron.2015.01.019.
100. Spatazza,J., Di Lullo,E., Joliot,A., Dupont,E., Moya,
K.L., and Prochiantz,A. (2013) Homeoprotein signal-
ing in development, health, and disease: a shaking of
dogmas offers challenges and promises from bench
to bed, Pharmacol. Rev., 65, 90-104, doi: 10.1124/
pr.112.006577.
101. Joliot, A.H., Triller,A., Volovitch,M., Pernelle,C., and
Prochiantz, A. (1991) Alpha-2,8-polysialic acid is the
neuronal surface receptor of antennapedia homeo-
box peptide, New Biol., 3, 1121-1134.
102. Merabet,S., Kambris, Z., Capovilla, M., Bérenger, H.,
Pradel, J., and Graba, Y. (2003) The hexapep-
tide and linker regions of the AbdA Hox pro-
tein regulate its activating and repressive func-
tions, Dev. Cell, 4, 761-768, doi: 10.1016/s1534-5807
(03)00126-6.
103. Duffraisse, M., Paul, R., Carnesecchi, J., Hudry, B.,
Banreti,A., Reboulet, J., Ajuria,L., Lohmann, I., and
Merabet, S. (2020) Role of a versatile peptide motif
controlling Hox nuclear export and autophagy in
the Drosophila fat body, J. Cell Sci., 133, jcs241943,
doi:10.1242/jcs.241943.
104. Bardine, N., Lamers, G., Wacker, S., Donow, C.,
Knoechel,W., and Durston,A. (2014) Vertical signal-
ling involves transmission of Hox information from
gastrula mesoderm to neurectoderm, PLoS One, 9,
e115208, doi:10.1371/journal.pone.0115208.
105. Kim, H.T., Kim, S.J., Sohn, Y.I., Paik, S.S., Caplette,R.,
Simonutti,M., Moon, K.H., Lee, E.J., Min, K.W., Kim,
M.J., Lee, D.G., Simeone,A., Lamonerie,T., Furuka-
wa,T., Choi, J.S., Kweon, H.S., Picaud,S., Kim, I.B.,
Shong,M., and Kim, J.W. (2015) Mitochondrial protec-
tion by exogenous Otx2 in mouse retinal neurons, Cell
Rep., 13, 990-1002, doi:10.1016/j.celrep.2015.09.075.
106. Brunet, I., Weinl, C., Piper, M., Trembleau, A., Volo-
vitch, M., Harris, W., Prochiantz, A., and Holt, C.
(2005) The transcription factor Engrailed-2 guides
retinal axons, Nature, 438, 94-98, doi: 10.1038/
nature04110.
107. DuBuc, T.Q., Stephenson, T.B., Rock, A.Q., and Mar-
tindale, M. Q. (2018) Hox and Wnt pattern the pri-
mary body axis of an anthozoan cnidarian before
gastrulation, Nat. Commun., 22, 2007, doi: 10.1038/
s41467-018-04184-x.
108. Schiemann, S.M., Martín-Durán, J.M., Børve,A., Vel-
lutini, B.C., Passamaneck, Y.J., and Hejnol,A. (2017)
Clustered brachiopod Hox genes are not expressed
collinearly and are associated with lophotrochozoan
novelties, Proc. Natl. Acad. Sci. USA, 114, 1913-1922,
doi:10.1073/pnas.1614501114.
109. Salamanca-Díaz, D. A., Calcino, A. D., de Oliveira,
A. L., and Wanninger, A. (2021) Non-collinear Hox
gene expression in bivalves and the evolution of mor-
phological novelties in mollusks, Sci. Rep., 11, 3575,
doi:10.1038/s41598-021-82122-6.
110. Levin,M., Anavy,L., Cole, A.G., Winter,E., Mostov,N.,
Khair, S., Senderovich, N., Kovalev, E., Silver, D. H.,
Feder, M., Fernandez-Valverde, S. L., Nakanishi, N.,
Simmons,D., Simakov, O., Larsson,T., Liu, S. Y., Jer-
afi-Vider, A., Yaniv,K., Ryan, J.F., Martindale, M. Q.,
et al. (2016) The mid-developmental transition and
the evolution of animal body plans, Nature, 531,
637-641, doi:10.1038/nature16994.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in pub-
lished maps and institutional affiliations.