Download A flexible genetic toolkit for arthropod neurogenesis

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Convolutional neural network wikipedia , lookup

Synaptogenesis wikipedia , lookup

Optogenetics wikipedia , lookup

Nervous system network models wikipedia , lookup

Neurogenomics wikipedia , lookup

Adult neurogenesis wikipedia , lookup

Biology and consumer behaviour wikipedia , lookup

Subventricular zone wikipedia , lookup

Artificial neural network wikipedia , lookup

Metastability in the brain wikipedia , lookup

Types of artificial neural networks wikipedia , lookup

Neuroanatomy wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Gene expression programming wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Recurrent neural network wikipedia , lookup

Neural engineering wikipedia , lookup

Development of the nervous system wikipedia , lookup

Transcript
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
A flexible genetic toolkit for arthropod
neurogenesis
rstb.royalsocietypublishing.org
Angelika Stollewerk
School of Biological and Chemical Sciences, Queen Mary University of London, Mile End Road, London E1 4NS, UK
Review
Cite this article: Stollewerk A. 2016 A flexible
genetic toolkit for arthropod neurogenesis.
Phil. Trans. R. Soc. B 371: 20150044.
http://dx.doi.org/10.1098/rstb.2015.0044
Accepted: 21 August 2015
One contribution of 16 to a discussion meeting
issue ‘Homology and convergence in nervous
system evolution’.
Subject Areas:
evolution, neuroscience
Keywords:
neurogenic potential, neural precursors,
neural progenitors, asymmetric division,
patterning genes, Notch signalling
Author for correspondence:
Angelika Stollewerk
e-mail: [email protected]
Arthropods show considerable variations in early neurogenesis. This
includes the pattern of specification, division and movement of neural precursors and progenitors. In all metazoans with nervous systems, including
arthropods, conserved genes regulate neurogenesis, which raises the question of how the various morphological mechanisms have emerged and
how the same genetic toolkit might generate different morphological
outcomes. Here I address this question by comparing neurogenesis
across arthropods and show how variations in the regulation and function
of the neural genes might explain this phenomenon and how they might
have facilitated the evolution of the diverse morphological mechanisms
of neurogenesis.
1. Introduction
During neurogenesis neurons and glia are formed and their number, types,
morphologies, final positions and connections shape the architecture of the
brain. Consequently, variations in the morphology and function of metazoan
nervous systems must originate from evolutionary changes of the developmental
processes that generate the neural cells and regulate their differentiation.
In order to understand the causes and sequence of evolutionary variations,
Volker Hartenstein and I, in a recent review [1], compared early neurogenesis
across the animal kingdom by subdividing the developmental processes of
early neurogenesis into distinct morphological modules. Neurogenesis starts
with patterning cells with neurogenic potential (module A) and further early
steps include the patterning (module B), proliferation (module C) and movement (module D) of neural progenitors. Variants of these modules can be
detected across the animal kingdom and allow conclusions to be drawn on
the mechanisms of neurogenesis in the urbilaterian [1]. For example, the
module ‘patterning cells with neurogenic potential’ (module A) is present as
three different variants. In animals like cnidarians that have a nerve net,
the whole ectoderm has neurogenic potential (‘generalized neurogenic
ectoderm’, module A1), whereas arthropods and vertebrates have restricted
ectodermal domains that generate the central nervous system (CNS) (‘neuroectoderm’, module A2). In some clades, the neurogenic potential is restricted
to subectodermal cells (module A3) such as the neoblasts in flatworms
(Platyhelminthes) [1].
Interestingly, the genetic toolkit of early neurogenesis is highly conserved in
all animal phyla despite significant variations in the morphological modules. In
this review, I will address the question of how the various morphological modules might have emerged and how the same genetic toolkit might generate the
different modules of neurogenesis, using arthropods as model organisms.
Arthropods are particularly well suited to analyse this question because of
the large size of the phylum, great variations in shapes and behaviours, and
the availability of comparative data on neurogenesis for all groups [1,2].
Furthermore, large-scale molecular phylogenies have greatly improved the resolution of arthropod relationships and thus facilitate evolutionary
investigations of arthropod neurogenesis (figure 1) [3]. It is now widely
accepted that the paraphyletic crustaceans group with the insects forming the
pancrustacean clade (also named Tetraconata clade). The pancrustaceans are
a sister group of the myriapods and are united with the latter in the
& 2015 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
Arthropoda
2
Onychophora
Myriapoda
Euchelicerata
Crustacea
Hexapoda
C3/D2
C3/D1
+C3/D1
C1/D3/D4
C1/D3/D4
B1/C2/D2
B2 (invariant neural
progenitors/precursors)
A2 (neuroectoderm)
Figure 1. Arthropod phylogeny and the evolution of the neurogenesis module. The arthropod relationships shown in the phylogenetic tree are the framework for
the discussions in this review. A2– D4 represent parts of the module variants recently described by [1]. Module A: patterning cells with neurogenic potential. In
contrast with animals with a generalized ectoderm (module A1, not shown), all arthropods have a restricted area with neurogenic potential (neuroectoderm) that
gives rise to the central nervous system (module A2). Module B: patterning of neural progenitors. Neural progenitors/precursor groups show a similar, invariant
pattern in all euarthropods (module B2), while the neural progenitors of onychophorans appear in random positions (module B1). The ancestral pattern of euarthropod neurogenesis is the selection of neural precursor groups. Modules C and D: proliferation and movement of neural progenitors, respectively. The neural
precursors of chelicerates and myriapods (mainly) do not divide (module C1) and directly differentiate into neurons and glia. The neural precursor groups are
internalized by ingression (module D3) or invagination (module D4). Both mechanisms are seen in chelicerates (including pycnogonids) and myriapods. In addition,
pycnogonids have asymmetrically dividing progenitors (module C3) that remain in the epithelium (module D1). By contrast, the neural progenitors of onychophorans
divide symmetrically (module C2) after their delamination (module D2). Insect and crustacean neuroblasts divide asymmetrically (module C3), however insect
neuroblasts delaminate (module D2), while crustacean neuroblasts remain in the epithelium (module D1). The scheme and description refer to the processes
in the ventral neuroectoderm.
Mandibulata group [3–6]. The chelicerates represent a basal
branch and are a sister group of the Mandibulata. Insects,
crustaceans, myriapods and chelicerates together compose
the euarthropods. The onychophorans (velvet worms) are
the sister group of the euarthropods and together with the
latter form the phylum Arthropoda (figure 1). The following
discussion of the neurogenesis modules is based on this
phylogenetic framework (but see [7– 10] for discussion).
2. Patterning regions with neurogenic potential
In arthropods, the procephalic and the ventral neuroectoderm
generate the tripartite brain and the ventral ganglia, respectively (module A2; figure 1) [11 –13]. The size of the
neuroectoderm differs from a few cell rows in crustaceans
and insects to wide areas containing many small cells in myriapods and chelicerates (figure 2) [14–19]. Data from insects
and a representative of a basal branch of the euarthropods,
the spider Parasteatoda tepidariorum (chelicerate), show that
dorsoventral (DV) cell fates, including the regions with neurogenic potential, are determined by morphogen gradients
involving Dpp/Sog [20–24]. The determination and patterning of the ventral neurogenic region by Dpp/Sog might,
therefore, be an ancestral feature of arthropod neurogenesis.
This is further supported by data from other Ecdysozoa,
Lophotrochozoa and Deuterostomia, which indicate that the
BMP(Dpp)/chordin(Sog) morphogen system is conserved
across bilaterians. However, there seem to be differences in
the degree of the contribution of the dpp/sog homologues in
patterning the neurogenic region in arthropods. In Drosophila
melanogaster sog null mutants, the DV extension is slightly
reduced [22], while in the flour beetle Tribolium castaneum
and in the spider (chelicerate) P. tepidariorum, sog RNAi results
in an (almost) complete loss of the ventral neuroectoderm
[20,23]. Data are not available for myriapods and data from
crustaceans are inconclusive because the only sog gene that
has been identified shows a restricted expression in the mandibular ventral midline cells [25]. The fact that the regulation
of DPP signalling has considerably diverged even within
insects [23,24] suggests that evolutionary variations in the
regulation of the DV patterning genes and their targets
might correlate with the different extensions of the neurogenic
regions in arthropods (figure 2). This assumption is further
supported by comparative studies of the DV extension of the
ventral midline in Diptera and Hymenoptera [26]. While in
D. melanogaster and in the mosquito Anopheles gambiae the ventral midline is 1–2 cells wide along the DV axis, it extends over
5–6 cells in the honeybee Apis mellifera. This difference is
because of variations in the regulation of single-minded—a transcription factor which is essential for the development of the
midline cells—by DV patterning genes [26].
In all arthropods the neurogenic potential differs along
the anterior –posterior (AP) axis. The (anterior) procephalic
Phil. Trans. R. Soc. B 371: 20150044
Pycnogonida
rstb.royalsocietypublishing.org
Euarthropoda
Mandibulata
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
(a)
pc
(b)
11
(f)
13
vne
Glomeris marginata msh
(c)
t1
4-1 4-2 4-3 4-4 3-5
5-5 5-6
5-3
5-1
5-4
5-2
6-4
6-1
6-2
7-1
7-3 7-4
MNB
7-2
t3
(g)
Glomeris marginata SoxB1
(d)
Tribolium castaneum msh
(e)
1-2
1-3
2-4
2-5
2-1 2-2 2-3
3-4 3-5
3-1 3-2 3-3
4-4
4-3
4-1 4-2
5-6
5-4
5-1 5-2 5-3
1-1
6-2 6-3 6-4
6-1
7-1 7-2 7-3 7-4
MNB
Glomeris marginata msh
Cupiennius salei msh
Tribolium castaneum
vnd, ind msh
Figure 2. Patterning of the neurogenic region in euarthropods. The dashed lines indicate the ventral midline of the neuroectoderm. (a) Flat preparation of a
Glomeris marginata embryo (millipede) showing the RNA expression pattern of the SoxB1 gene in all neuroectodermal cells. (b,c) Comparison of the msh RNA
expression pattern in flat preparations of three ventral neuromeres of G. marginata and Tribolium castaneum (beetle). Note the difference in the width of the
neuroectoderm in these species. The boxes frame the neuromere of the second leg segment in G. marginata and the second thoracic neuromere of T. castaneum,
respectively. (d,e) Schematic of the arrangement of the neural precursor groups in hemi-neuromeres of G. marginata and the spider Cupiennius salei. The precursor
groups expressing msh are shown in brown. ( f,g) Schematic drawings comparing the expression of the dorsoventral patterning genes vnd (blue), ind (red) and msh
(brown) in the neuroblasts of Drosophila melanogaster and T. castaneum. The neuroblasts 6-2 and 7-2 express both vnd and ind in D. melanogaster, while neuroblasts
4-2 and 5-3 express ind and msh in T. castaneum. l1–l3, leg segment 1 to 3; pc, procephalic neuroectoderm; t1–t3, thoracic segments 1 to 3; vne, ventral
neuroectoderm.
neuroectoderm not only generates a higher and more
diverse number of neurons than the ventral neuroectoderm
but also forms the complex brain centres such as the mushroom bodies that integrate sensory information and regulate
behaviour [27]. In vertebrates, the Wnt/b-catenin pathway
patterns the CNS along the AP axis and in some species
determines the overall AP cell fates (reviewed in [28]). An
involvement of this signalling pathway in patterning the
AP axis has also been demonstrated in lophotrochozoans,
in particular in regenerating planarians, where the
reduction of Wnt signalling results in radial hypercephalized
individuals [29].
In arthropods, Wnt signalling seems to regulate posterior
elongation and segmentation but is neither required for the
formation of anterior (head) segments nor the overall AP patterning of the CNS [28,30]; rather, the function of Wnt in
establishing AP fates in the arthropod CNS seems to be confined to neuromeres, the metameric units of the CNS [31– 34].
The mechanisms establishing different AP fates in the
anterior part of the embryo show considerable variations
within and between arthropod groups. The AP morphogen
Bicoid, for example, is not present outside of dipterans and
Hunchback, another head morphogen, does not regulate
the expression of AP patterning genes (e.g. Hox genes)
in the spider P. tepidariorum, in contrast with the case in
D. melanogaster [35,36]. If the overall AP patterning mechanisms are connected with the regionalization of the CNS,
these differences might contribute to existing variations of
the arthropod CNS. In D. melanogaster, AP axis formation is
in fact directly linked to the AP regionalization of the CNS.
AP morphogens (e.g. Bicoid, Hunchback) regulate the genes
that pattern the CNS along the AP axis (e.g. orthodenticle,
ems, Hox genes) (e.g. [37,38]). However, there are no data
available on how overall AP patterning is linked to CNS
regionalization in arthropods other than insects.
Thus, while the overall function of BMP in DV axis formation and patterning of the neurogenic region seems to be
conserved across the animal kingdom, the function of
Wnt in AP patterning of the CNS has been restricted to the
metameric units in arthropods. Furthermore, the upstream
mechanisms that set up the initial differences in CNS AP
fates have diverged in arthropods.
However, many of the genes downstream of the primary
AP/DV subdivision are conserved across the animal kingdom, such as the transcriptional regulators of the SoxB
family, which maintain the proliferative state of the neuroepithelium and at the same time inhibit neural
differentiation (figure 2) [39–42].
Furthermore, many genes that translate the positional
information into neural/lineage subtype identity are
Phil. Trans. R. Soc. B 371: 20150044
Drosophila melanogaster
vnd, ind msh
t2
rstb.royalsocietypublishing.org
12
3
1-1 1-2
1-3
2-1 2-2 2-3 2-4 2-5
3-1 3-2 3-3 3-4
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
Neural precursors/progenitors can appear in either random
(module B1) or fixed (module B2) positions within regions
of neurogenic potential (figure 3) [1]. Furthermore, single
(modules D1 and D2) or groups of contiguous precursors
can be selected (modules D3 and D4). In the arthropod
neuroectoderm, different types of founder cells of the
nervous system are generated over a period of time ranging
from hours (e.g. in the insect D. melanogaster) to months
(e.g. in the onychophoran Euperpatoides kanangrensis) [55]. In
chelicerates and myriapods groups of contiguous mainly
postmitotic neural precursors are selected in the neuroectoderm, while in onychophorans, crustaceans and insects
single neural progenitors are selected that divide either symmetrically or in a stem-cell like mode (see below) (figure 3)
[7,14,15,17,18,55,56]. In onychophorans, the sister group of
euarthropods (chelicerates, myriapods, crustaceans, insects),
neural progenitors appear in random positions in the
neuroectoderm and their number per hemi-neuromere is
variable (60– 100 progenitors; figure 3). By contrast, in all
euarthropods neural progenitors/precursors are generated
in fixed positions and show species-specific invariable
numbers within the same neuromeres (figures 2 and 3).
The morphological differences can be explained by
variations in the regulation of two classes of genes, the
achaete-scute homologues (ASH) and the Notch signalling pathway. These genes are used for the generation and regulation of
4
Phil. Trans. R. Soc. B 371: 20150044
3. Patterning of neural precursors/progenitors
neural cells throughout the animal kingdom; however, they
show variations in their spatial and temporal expression
(reviewed by [1]). In insects, chelicerates and myriapods the
invariant spatial–temporal expression of ASH genes sets up
the areas where neural progenitors/precursor groups form
[17,18,57,58]. The positions of the neural progenitors/precursors are refined by a transcriptional feedback loop between
the Notch signalling pathway and the ASH genes, which
results in the selection of regular arranged, spaced neural
precursor groups/progenitors [59,60]. In branchiopod crustaceans, the ASH genes are not expressed before neural
progenitors (neuroblasts) are formed and therefore are not
involved in positioning them (figure 3) [15]. However, an
invariant spatial pattern of ASH expression appears with a
time delay. Neuroblasts are also arranged in an invariant
pattern in another crustacean group, the malacostracans, but
molecular expression data are not available [14].
By contrast, in the onychophoran E. kanangrensis, ASH is
expressed at very low, homogeneous levels in the neuroectoderm [55]. High levels of expression can be detected and are
maintained in the neural progenitors after their delamination,
suggesting that Ek ASH is required for neural progenitor
maturation (figure 3).
Two mechanisms might contribute to the variations in patterning neural precursors/progenitors in the neuroectoderm
in arthropods. Firstly, the variations could be due to changes
in the prepatterning mechanisms regulating ASH expression.
Given the homogeneous low ASH expression in onychophorans, molecular changes in the achaete-scute homologues must
have occurred in the lineage leading to the euarthropods that
facilitated the spatio-temporal regulation of these genes allowing for the upregulation of transcripts in invariant positions.
However, data on ASH regulation in the embryonic CNS are
fragmentary in insects and not available at all in other arthropods. Skeath et al. [61] report that the D. melanogaster ASH
genes achaete, scute and lethal of scute are regulated by the
products of the pair-rule and DV axis patterning genes (e.g.
dpp) in the ventral neuroectoderm and their expression
patterns are maintained by the segment polarity proteins. In
the D. melanogaster procephalic neuroectoderm, lethal of scute,
which is the main proneural gene in the brain, is regulated
by the head gap genes tailless (tll), orthodenticle (otd), empty
spiracles (ems) and buttonhead (bhd) [62]. It can be speculated
that evolutionary changes in the spatial– temporal expression
of ASH genes are associated with variations in the expression
and interactions of the segmentation genes, which have
considerably diverged in arthropods, in particular regarding
the gap and pair-rule genes (e.g. [63 –69]).
However, segmentation genes might not be used at all for
positioning neural precursors/progenitors in some groups.
For example, in the neuroectoderm of the branchiopod
crustacean Daphnia magna, the default neural cell fate is suppressed in the central area of the hemi-neuromeres by the
activation of Notch signalling so that neural progenitors are
arranged in a ring-like structure (figure 3). This pattern
seems to be representative for branchiopods; however, in
malacostracan crustaceans the neuroblast arrangement is
not disrupted by a non-neural central area, suggesting that
the mechanisms of neuroblast positioning have diverged in
crustaceans [14].
Secondly, variations in the regulation of Notch activity
could underlie the variation in the pattern of neural
precursors/progenitors in arthropods.
rstb.royalsocietypublishing.org
conserved in metazoans. Downstream of the primary AP and
DV patterning mechanisms, additional signalling pathways
as well as homeodomain proteins establish neuronal subtype
identity [43]. For example, the function of Hox genes in determining segmental identity is conserved across the animal
kingdom [43]. However, although these genes seem to be
an integral downstream part of the molecular processes that
pattern the neurogenic region along the AP axis of the CNS
in all arthropods (e.g. [44–47]), there are considerable functional variations even within individual arthropod groups.
For example, the expression of the Hox gene labial in the
head segment that gives rise to the tritocerebrum is conserved in all arthropods [48]; however, studies in insects
show divergent functions of Labial in brain development.
While in D. melanogaster labial mutants the tritocerebrum
and additional brain structures are affected, labial RNAi
does not show a phenotype in the milkweed bug Oncopeltus
fasciatus and the functional defects are restricted to the
tritocerebrum in the flour beetle T. castaneum [49–51].
The genes patterning the neurogenic region along the DV
axis are also highly conserved across the animal kingdom
[52]. Yet again, detailed comparative analyses of the
expression of the DV patterning genes muscle segment homeobox (msh), intermediate neuroblasts defective (ind) and ventral
nervous system (vnd) defective in arthropods revealed considerable variations in the expression of these genes relative
to the neural progenitors/precursors in the respective
domains (figure 2) [53,54]. Furthermore, evolutionary
changes have occurred in the neuronal subtype specific
genes that are regulated by the DV patterning genes. For
example, msh regulates the expression of the motor- and
interneuronal marker islet in the spider Cupiennius salei,
whereas this is not the case in D. melanogaster [53].
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
onychophorans
chelicerates
insects
crustaceans
intermediate
neural precursor
neuron/glia
neural
precursor
neural
progenitor
neuro/glia
ganglion
mother
cell
neuroblast
symmetrically
dividing
neuroblast
Figure 3. Pattern of neural precursor/progenitor specification and division. Small sections of the neuroectoderm (white squares: neuroectodermal cells) are shown
for each group. (a) Onychophorans: ASH shows a low homogeneous expression in the neuroectoderm (light grey) and is upregulated in the neural progenitors (dark
grey) after their delamination. Chelicerates: ASH is upregulated in large domains and becomes restricted to groups of cells. A similar pattern is seen in myriapods
(not shown). Insects: ASH is expressed in small proneural clusters. Expression becomes restricted to single neuroblasts. Crustaceans (branchiopods): ASH is upregulated after neuroblast specification. Over time more neuroblasts express ASH and the gene is expressed at different levels indicated by lighter and darker grey.
(b) Division pattern and movement of neural precursors/progenitors. Onychophorans: single neural progenitors delaminate and divide symmetrically to produce
intermediate neural precursors, which divide again. The expression of snail and prospero has not been studied in onychophorans. Chelicerates: most neural precursors
are postmitotic. Neural precursor groups express snail (orange). Neural precursors co-express prospero (red) before detaching from the group. Insects: neuroblasts
express snail and prospero and divide asymmetrically to produce GMCs, which divide again to produce neurons and glia. Crustaceans show the same division pattern
but snail is expressed long before prospero. Snail-positive neuroblasts can divide symmetrically.
The core mechanism of the Notch signalling pathway is
the activation of the Notch receptor by the ligand Delta,
which results in the intramembrane proteolysis of the receptor and the release of an intracellular fragment that acts as a
transcriptional regulator [70]. The interactions between the
transmembrane proteins Notch and Delta are complex,
and different regulatory mechanisms are used in different
developmental processes such as ligand– receptor trans- and
cis-activation, receptor cis-inhibition by the ligand or
mutual receptor– ligand cis-inhibition [70,71].
Biological and mathematical models show that the selection pattern of precursors can be modulated by regulating
Notch signalling in different ways [71,72]. The selection of
regular arranged, spaced neural precursor groups/progenitors can be explained by the ‘lateral inhibition model’
where Notch signalling and the ASH genes are linked in a
transcriptional feedback mechanism [70]. In insects, the
ASH genes are expressed in small clusters in the ventral
neuroectoderm. One of the cells adopts the neural fate and
expresses higher levels of ASH, which in turn leads to an
increase of Delta expression [59]. Delta trans-activates the
Notch receptor in the neighbouring cells. The products of
the Notch target genes, the hairy/Enhancer of Split (HES)
genes, downregulate ASH and inhibit the neighbouring
cells from adopting the neural fate. In chelicerates and
myriapods, lateral inhibition operates between the neural
precursors and the neuroectoermal cells surrounding the
selected precursor group [18,60]. However, the neural precursor group itself seems to be insensitive to Notch signalling.
This is reflected in the high levels of Delta expressed by the
precursor groups suggesting that the Notch receptor is
cis-inactivated [60,71].
In onychophorans, the homogeneous, low expression of
ASH and the strong expression of Notch and Delta in the ventral neuroectoderm suggest that Notch signalling keeps ASH
expression at low levels in all cells (figure 3) [55]. Random
changes in the sensitivity to Notch signalling, possibly by
cis-inactivation of the Notch receptor by the ligand Delta,
could allow for individual cells to adopt the neural fate and
delaminate. This mechanism could explain the seemingly
random positions and great range in progenitor numbers
per hemi-neuromere.
The pattern of neural precursors/progenitors is much
more complex in the developing brain. In chelicerates and
myriapods, the same types of founder cells (neural precursor
groups) are visible in the procephalic and ventral neuroectoderm [73]. In both neurogenic areas, the ASH genes are
expressed in large domains from which several neural precursor groups arise (although the underlying morphological
processes are more complex in the procephalic neuroectoderm, see below). By contrast, in insects, ASH genes are
expressed in large domains (rather than small clusters) in
Phil. Trans. R. Soc. B 371: 20150044
(b)
5
rstb.royalsocietypublishing.org
time
(a)
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
Arthropods show also variations in the developmental modules following neural determination. Selected cells can either
differentiate directly into neural cells (neurons/glia) (module
C1), as is the case for many of the neural precursors in chelicerates and myriapods, or they can undergo symmetric (e.g.
onychophorans) or asymmetric divisions (insects, crustaceans)
(modules C2 and C3; figure 1). They can also show combinations of these variants (e.g. sea spiders; figure 1). The
associated movements either of the neural precursors/progenitors themselves and/or of their progeny lead to the
formation of internal cell layers, basal to the neuroectoderm
(figure 3). The movements show variations that are partially
linked to the differences in the specification patterns of neural
precursors/progenitors. Insect neuroblasts start to divide
asymmetrically to produce ganglion mother cells (GMCs)
after segregating individually from the neuroectoderm (delamination; modules C3 and D2), while crustacean neuroblasts
remain in the neuroectoderm and generate GMCs at the basal
side by directed asymmetric mitosis (module C3 and D1;
figures 1 and 3) [14–16,77]. Onychophoran neural progenitors delaminate before they divide symmetrically to
produce intermediate neural precursors (modules C2 and
D2; figures 1 and 3) [7,55]. By contrast, there are no cell divisions within the neural precursor groups of chelicerates and
myriapods (module C1) and only few neural precursors
divide after detaching from the precursor groups (ingression;
module D3) or the internalization of the whole group (invagination; module D4; figures 1 and 3).
However, in myriapods single and groups of mitotic cells
seem to prefigure the areas where neural precursor groups
are selected, suggesting a link between the formation of
neural precursors and cell proliferation [18]. However, the
6
Phil. Trans. R. Soc. B 371: 20150044
4. Proliferation and movement of neural
precursors
division pattern has not been analysed in detail and it is
not known if the neural precursors within a group are related
by lineage. Interestingly, sea spiders, which are most
probably the sister group of all other chelicerates (euchelicerates), show a combination of asymmetrically dividing
neural progenitors and neural precursor groups [78]. However, in this case, neural precursor groups are generated
before the neural progenitors appear.
Despite these differences, committed neural progenitors/
precursors express a set of conserved factors that were shown
to be involved in asymmetric division and neural cell fate
determination in D. melanogaster [79]. This raises the question
of how these genes can support the various morphological
outcomes in all euarthropod groups (data are not available
for onychophorans).
In the following, I will discuss a few of the factors for
which comparative data exist in euarthropods. In D. melanogaster, Asense activates many genes in neuroblasts, among
others those that are required for asymmetric division [80].
The neuroblasts immediately delaminate after their formation and start to divide (figure 3). The neural
differentiation factor Prospero is asymmetrically distributed
into the GMCs where it enters the nucleus, acts antagonistically to Asense and promotes cell-cycle exit [81–83]. The
Snail family members are required for the mitotic activity
of the neuroblast and the spindle rotation, which occurs
prior to the production of GMCs [79]. The asymmetric distribution of Prospero to GMCs is affected in mutants
carrying a deletion of three Snail family genes [79]. Furthermore,
the Snail family members might be directly involved in triggering the delamination of neuroblasts by downregulating
E-cadherin [84,85].
Although crustacean neuroblasts do not delaminate, they
show the same division pattern as insect neuroblasts, and it is
therefore conceivable to assume that the function of Snail and
Asense in asymmetric cell division is conserved in both
groups. Furthermore, prospero is asymmetrically distributed
into the GMCs in crustaceans and its expression coincides
with the onset of asymmetric division of the neuroblasts,
suggesting that it acts as a neural cell fate determinant similar
to the case in insects [15,86]. However, Snail must have an
additional earlier function in crustacean neurogenesis
because it is the earliest gene to be expressed in neuroblasts
and there is a considerable time delay between neuroblast
formation and the start of asymmetric divisions [15,86]. The
fact that snail is expressed before ASH suggests that the
former rather than the latter gene is required for neural cell
fate determination. This potential function of Snail is supported by data from D. melanogaster showing that Snail can
activate neuroblast fate genes such as grainyhead, numb,
prospero and seven-up [80]. Furthermore, Snail could be
involved in the symmetric divisions of crustacean neuroblasts, which occur before the start of GMC production
(figure 3) [15]. The function of Snail in regulating cell-cycle
genes and the proliferative divisions of stem cells has
been demonstrated both in D. melanogaster and in
vertebrates [80,87 –89].
The delayed progression of crustacean neuroblasts to
asymmetric divisions compared with insects can again be
explained by changes in the use of the participating genes
[86]. Notch signalling is used not only for binary cell fate
decisions in the hemi-neuromeres between the central presumptive epidermal cells and the surrounding neuroblasts,
rstb.royalsocietypublishing.org
some areas of the procephalic neuroectoderm, from which
several neuroblasts delaminate [74]. Furthermore, in the
dorsal midline of the procephalic neuroectoderm of
D. melanogaster, groups of neural precursors form within a
large lethal of scute domain [74]. Yet again, different patterns
and regulations of ASH and Notch signalling can be seen in
different ontogenetic stages (e.g. larvae) and in the sensory
nervous system (e.g. [75,76]). The mechanisms seen in the
developing insect brain show similarities to the ancestral
patterns of neurogenesis described in onychophorans and
chelicerates/myriapods.
These data have led to the hypothesis that the neural precursor groups of chelicerates and myriapods have evolved
from areas of massive, random neural progenitor delamination due to changes in the regulation of the ASH genes and
in the regulation of Notch signalling [55]. There is strong evidence that neural precursor groups were present in the last
common ancestor of euarthropods and thus represent the
ancestral pattern of euarthropod neurogenesis [55].
Additional changes in the regulation of the participating
genes have led to the evolution of neuroblasts from the
neural precursor groups in the last common ancestor of
insects and crustaceans. Thus, changes in the prepatterning
mechanisms and the versatility of the Notch signalling pathway might explain the various patterns of neural precursors/
progenitors in arthropods.
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
5. Conservation of structure despite changes
of gene expression
Phil. Trans. R. Soc. B 371: 20150044
So far, I have discussed how the expression of conserved genes
in neural precursors/progenitors can lead to different morphological outcomes. However, in arthropod neurogenesis
the reverse mechanism also exists: the expression of conserved
genes in different patterns resulting in the same morphological
outcome. This phenomenon might contribute to internal buffering mechanisms ensuring that evolutionary modifications
do not result in destabilization and functional breakdown of
neuronal structures. In the following, I will give two examples
relating to the axonal scaffold of euarthropods.
In all euarthropods, axonal tracts are established in the
developing ventral nerve chord in a highly conserved pattern
[53]. Two longitudinal tracts connect the brain and the ventral
ganglia. In each ventral neuromere, two transverse tracts, the
anterior and posterior commissures, connect the longitudinal
fascicles. During later development this arrangement can be
obscured by the fusion of ganglia. The conserved axonal scaffold seems to be related to the equally highly conserved
arrangement of the neural precursor groups/progenitors in
euarthropods. In all four groups, they are arranged in seven
rows and three to six columns in each hemi-neuromere
(figure 2) [53]. Spatial and temporal cues within the neuroectoderm lead to the expression of a unique set of genes in
the neural precursors/progenitors, which in turn determines
the subtype identity of the progeny [43,53,54,98 –100]. The
neural precursors/progenitors produce early sets of pioneer
neurons that establish the axonal scaffold [101]. One set of
pioneer neurons is characterized by the expression of the
motor and interneuronal marker even-skipped [54,102,103].
The position of eve-positive neurons is highly conserved in
insects and crustaceans. However, a recent comparison of
the neuroblast expression patterns in D. melanogaster and in
the flour beetle T. castaneum revealed that despite their conserved positions the expression profiles of the neuroblasts
have diverged considerably, including the profiles of those
generating the eve-positive pioneer neurons [54,99]. Interestingly, these evolutionary changes include the expression of
patterning genes which are highly conserved throughout
the animal kingdom, such as the DV patterning genes msh,
ind and vnd (figure 2). These data show that the early neuronal programme, which generates the pioneer neurons,
tolerates considerable molecular variations, thus ensuring
the formation of the conserved axonal scaffold.
In addition to the pioneer neurons, a sophisticated system
of axonal guidance molecules at the ventral CNS midline
plays an important role in generating the axonal scaffold in
euarthropods [104–106]. For example, Netrin, a conserved
chemoattractant, attracts commissural axons to the ventral
midline [107]. In insects and crustaceans, specialized midline
cells consisting of neurons and glia are the source of Netrin
expression [106,107]. In spiders, however, neuronal midline
cells are absent and the midline consists of epithelial cells
[108]. Furthermore, during embryogenesis, spiders undergo
a process called inversion whereby the germband splits
into two halves at the ventral side along the AP axis. The
two halves move apart and during this process a ventral
midline epithelium appears which extends between the
germband halves. After dorsal closure the germband halves
come together again ventrally and the midline epithelium
7
rstb.royalsocietypublishing.org
but also for keeping neuroblasts in the transition phase which
is characterized by the expression of neural genes but the
absence of asymmetric divisions. Neuroblasts individually
leave the transition phase showing strong expression of
ASH, asense and prospero and start producing GMCs [86].
Both ASH and prospero expression are derepressed when
Notch activity is lowered, leading to premature production
of GMCs. Thus in this context, ASH seems to be required
for the maturation of the neuroblasts, rather than the
specification of the neural fate as is the case in insects.
Furthermore, in embryos with reduced Notch activity
the epithelial morphology is disturbed, indicating that
Notch signalling might have an additional function in keeping neuroblasts within the neuroectoderm [86]. This
function of Notch has been described both in vertebrates
and in arthropods. For example, inactivation of the Notch
effector genes Hes 1, 3 and 5 in mice results in reduction
of tight and adherens junctions at the apical end feed of
neural stem cells and loss of these cells from the spinal
cord neuroepithelium [90]. In the spider C. salei (chelicerate)
neural precursor groups disintegrate prematurely in the
CNS and peripheral nervous system (PNS), showing that
this function of Notch can also support the formation and
maintenance of neural precursor groups [17,91].
In spiders (chelicerates), most of the neural precursors
expressing snail and prospero are postmitotic and the few precursors that proliferate divide symmetrically, thus excluding
a function of these panneural factors in asymmetric division
[92]. However, Prospero seems to act as neural differentiation factor and might promote cell-cycle exit similar to the
case in insects and crustaceans because it is expressed in
the nuclei of neurons and their precursors in the CNS and
PNS [92,93]. Snail, on the other hand, must have a different
function. Interestingly, it has been shown in vertebrates and
recently also in D. melanogaster that members of the Snail
family are sufficient to induce cell-shape changes leading to
the delamination of neural precursors from the neuroepithelium [94–96]. It is, therefore, conceivable that Snail is
required for the morphological changes that occur during
invagination/ingression of the neural precursors both in chelicerates and myriapods, where snail is expressed in the same
pattern [35].
Asense is neither present in chelicerates nor in myriapods;
however, the spider C. salei has two ASH genes, which arose
by independent duplication in the spider lineage [97]. While
CsASH1 is initially expressed in large proneural domains,
CsASH2 is exclusively expressed in neural precursors and is
able to rescue asense-specific defects in D. melanogaster
mutants [97]. This suggests that there is yet another mechanism that could account for the morphological differences of
neurogenesis in arthropods—the convergent evolution of
genes promoting the neural programme.
To summarize, these explanations show that the emergence of the various neurogenesis modules might have
been facilitated by the diverse functions of the participating
genes as well as their flexibility of acting in different morphological contexts. Furthermore, the modularity of early
neurogenesis has allowed for combining various modules
so that neurogenesis could be adapted to different requirements not only in the individual arthropod groups/species
but also in different areas of neurogenesis in the same
organism.
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
Molecular genetic changes that positively influenced the flexible use of the gene product, for example by introducing
Competing interests. I declare we have no competing interests.
Funding. I received no funding for this study.
Acknowledgements. I am grateful to Nick Strausfeld, Frank Hirth and the
Royal Society for inviting me to participate in the March 2015 discussion meeting ‘Origin and evolution of the nervous system’. I also
gratefully acknowledge the participants of the discussion meeting
for thought-provoking and helpful comments.
References
1.
Hartenstein V, Stollewerk A. 2015 The evolution of
early neurogensis. Dev. Cell 32, 390–407. (doi:10.
1016/j.devcel.2015.02.004)
2. Storch V, Welsch U. 2004 Systematische zoologie.
Berlin, Germany: Spektrum akademischer Verlag.
3. Regier JC, Shultz JW, Zwick A, Hussey A, Ball B,
Wetzer R, Martin JW, Cunningham CW. 2010
Arthropod relationships revealed by phylogenomic
analysis of nuclear protein-coding sequence. Nature
463, 1079 –1084. (doi:10.1038/nature08742)
4. Rehm P, Meusemann K, Borner J, Misof B,
Burmester T. 2014 Phylogenetic position of
Myriapoda revealed by 454 transcriptome
sequencing. Mol. Phylogenet. Evol. 77, 25 –33.
(doi:10.1016/j.ympev.2014.04.007)
5. Rota-Stabelli O, Campbell L, Brinkmann H,
Edgecombe GD, Longhorn SJ, Peterson KJ, Pisani D,
Philippe H, Telford MJ. 2011 A congruent solution
to arthropod phylogeny: phylogenomics, microRNAs
and morphology support monophyletic
Mandibulata. Proc. R. Soc. B 278, 298–306.
(doi:10.1098/rspb.2010.0590)
6. Meusemann K, von Reumont BM, Simon S,
Roeding F, Strauss S, Kuck P, Ebersberger I, Walzl M,
Pass GBS. 2010 A phylogenomic approach to resolve
the arthropod tree of life. Mol. Biol. Evol. 27,
2451–2464. (doi:10.1093/molbev/msq130)
7. Mayer G, Whitington PM. 2009 Velvet worm
development links myriapods with chelicerates.
Proc. R. Soc. B 276, 3571 –3579. (doi:10.1098/rspb.
2009.0950)
8. Friedrich M, Tautz D. 1995 Ribosomal DNA
phylogeny of the major extant arthropod classes
and the evolution of myriapods. Nature 376,
165–167. (doi:10.1038/376165a0)
9. Boore JL, Collins TM, Stanton D, Daehler LL, Brown
WM. 1995 Deducing the pattern of arthropod
phylogeny from mitochondrial DNA rearrangements.
Nature 376, 163 –165. (doi:10.1038/376163a0)
10. Kusche K, Burmester T. 2001 Diplopod hemocyanin
sequence and the phylogenetic position of the
Myriapoda. Mo. Biol. Evol. 18, 1566– 1573. (doi:10.
1093/oxfordjournals.molbev.a003943)
11. Bullock TH, Horridge GA. 1965 Structure and
function of the nervous system of invertebrates.
San Francisco, CA: Freeman.
12. Mittmann B, Scholtz G. 2003 Development of the
nervous system in the ’head’ of Limulus
polyphemus (Chelicerata, Xiphosura): morphological
evidence for a correspondence between the
segments of the chelicerae and of the (first)
antennae of Mandibulata. Dev. Genes Evol.
213, 9 –17.
13. Strausfeld NJ. 1998 Crustacean –insect relationships:
the use of brain characters to derive phylogeny
amongst segmented inverebrates. Brain Behav.
Evol. 52, 186–206. (doi:10.1159/000006563)
14. Ungerer P, Scholtz G. 2008 Filling the gap between
identified neuroblasts and neurons in crustaceans
adds new support for Tetraconata. Proc. R. Soc. B
275, 369 –376. (doi:10.1098/rspb.2007.1391)
15. Ungerer P, Eriksson BJ, Stollewerk A. 2011
Neurogenesis in the water flea Daphnia magna
(Crustacea, Branchiopoda) suggests different
mechanisms of neuroblast formation in insects
and crustaceans. Dev. Biol. 357, 42– 52. (doi:10.
1016/j.ydbio.2011.05.662)
16. Scholtz G. 1990 The formation, differentiation and
segmentation of the post-naupliar germ band of
the amphipod Gammarus pulex L. (Crustacea,
Malacostraca, Peracarida). Proc. R. Soc. Lond. B 239,
163 –211. (doi:10.1098/rspb.1990.0013)
17. Stollewerk A, Weller M, Tautz D. 2001 Neurogenesis
in the spider Cupiennius salei. Development 128,
2673 –2688.
18. Dove H, Stollewerk A. 2003 Comparative analysis of
neurogenesis in the myriapod Glomeris marginata
(Diplopoda) suggests more similarities to
chelicerates than to insects. Development 130,
2161 –2171. (doi:10.1242/dev.00442)
19. Mittmann B. 2002 Early neurogenesis in the
horseshoe crab Limulus polyphemus and its
implication for arthropod relationships. Bio.
Bull. 203, 221–222. (doi:10.2307/1543407)
20. Akiyama-Oda Y, Oda H. 2006 Axis specification in
the spider embryo: dpp is required for radial-to-
21.
22.
23.
24.
25.
26.
27.
28.
29.
axial symmetry transformation and sog for ventral
patterning. Development 133, 2347– 2357. (doi:10.
1242/dev.02400)
Ferguson EL, Anderson KV. 1992 decapentaplegic
acts as a morphogen to organize dorsal –ventral
pattern in the Drosophila embryo. Cell 71,
451–461. (doi:10.1016/0092-8674(92)90514-D)
Francois V, Solloway M, O’Neill JS, Emery J, Bier E.
1994 Dorsal –ventral patterning of the Drosophila
embryo depends on a putative negative growth
factor encoded by the short gastrulation gene.
Genes Dev. 8, 2602– 2616. (doi:10.1101/gad.8.
21.2602)
van der Zee M, Stockhammer O, Levetzow C, Nunes
da Fonseca R, Roth S. 2006 Sog/Chordin is required
for ventral-to-dorsal Dpp/BMP transport and head
formation in a short germ insect. Proc. Natl Acad.
Sci. USA 103, 16 307–16 312. (doi:10.1073/pnas.
0605154103)
Özüak O, Buchta T, Roth S, Lynch JA. 2014
Dorsoventral polarity of the Nasonia embryo
primarily relies on a BMP gradient formed without
input from Toll. Curr. Biol. 24, 2393–2398. (doi:10.
1016/j.cub.2014.08.035)
Vargas-Vila MA, Hannibal RL, Parchem RJ, Liu PZ,
Patel NH. 2010 A prominent requirement for singleminded and the ventral midline in patterning the
dorsoventral axis of the crustacean Parhyale
hawaiensis. Development 137, 3469–3476. (doi:10.
1242/dev.055160)
Zinzen RP, Cande J, Ronshaugen M, Papatsenko D,
Levine M. 2006 Evolution of the ventral midline in
insect embryos. Dev. Cell 11, 895 –902. (doi:10.
1016/j.devcel.2006.10.012)
Heisenberg M. 2003 Mushroom body memoir: from
maps to models. Nat. Rev. Neurosci. 4, 266–275.
(doi:10.1038/nrn1074)
Niehrs C. 2010 On growth and form: a Cartesian
coordinate system of Wnt and BMP signaling
specifies bilaterian body axes. Development 137,
845–857. (doi:10.1242/dev.039651)
Adell T, Cebriá F, Saló E. 2010 Gradients in planarian
regeneration and homeostasis. Cold Spring Harb.
8
Phil. Trans. R. Soc. B 371: 20150044
6. Conclusion
additional regulatory mechanisms and expanding the molecular interactions of the protein, might have been the
starting point of the evolution of the diverse morphological
mechanisms of neurogenesis. The conservation of the genetic
tools of neurogenesis across the animal kingdom suggests
that this might be a general principle of morphological
evolution.
rstb.royalsocietypublishing.org
disappears [108]. Interestingly, commissural axons cross the
space between the germband halves during inversion and
Netrin, which is expressed in transverse stripes in the midline
epithelium, is required for attracting the axons and guiding
them across the midline [105]. This example shows an
evolutionary change of gene expression to accommodate
the morphological peculiarity of spider embryogenesis so
that an important axonal guidance molecule can perform
its conserved function in establishing the commissural tracts.
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
31.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57. Cabrera CV, Martinez-Arias A, Bate M. 1987 The
expression of three members of the achaete-scute
gene complex correlates with neuroblast
segregation in Drosophila. Cell 50, 425 –433.
(doi:10.1016/0092-8674(87)90496-X)
58. Kadner D, Stollewerk A. 2004 Neurogenesis in
the chilopod Lithobius forficatus suggests more
similarities to chelicerates than to insects. Dev.
Genes Evol. 214, 367 –379. (doi:10.1007/s00427004-0419-z)
59. Cabrera CV. 1990 Lateral inhibtion and cell fate
during neurogenesis in Drosophila: interactions
between scute, Notch, Delta. Development 109,
733–742.
60. Stollewerk A. 2002 Recruitment of cell groups
through Delta/Notch signalling during spider
neurogenesis. Development 129, 5339–5348.
(doi:10.1242/dev.00109)
61. Skeath JB, Panganiban G, Selegue J, Carroll SB.
1992 Gene regulation in two dimensions: the
proneural achaete and scute genes are controlled by
combinations of axis patterning genes through a
common intergenic control region. Genes Dev. 6,
2606– 2619. (doi:10.1101/gad.6.12b.2606)
62. Younossi-Hartenstein A, Green P, Liaw GJ, Rudolph
K, Lengyel J, Hartenstein V. 1997 Control of
early neurogenesis of the Drosophila brain by
the head gap genes tll, otd, ems, and btd. Dev.
Biol. 182, 270–283. (doi:10.1006/dbio.
1996.8475)
63. Choe CP, Brown SJ. 2007 Evolutionary fexibility of
pair-rule patterning revealed by functional analysis
of secondary pair-rule genes, paired and slopypaired in the short germ insect, Tribolium
castaneum. Dev. Biol. 302, 281–294. (doi:10.1016/
j.ydbio.2006.09.037)
64. Peel AD, Chipman AD, Akam M. 2005 Arthropod
segmentation: beyond the Drosophila paradigm.
Nat. Rev. Genet. 6, 905 –916. (doi:10.1038/
nrg1724)
65. Janssen R, Budd GE. 2013 Deciphering the
onychophoran ’segmentation gene cascade’: Gene
expression reveals limited involvement of pair rule
gene orthologs in segmentation, but a highly
conserved segment polarity gene network. Dev.
Biol. 382, 224–234. (doi:10.1016/j.ydbio.2013.
07.010)
66. Schinko JB, Kreuzer N, Offen N, Posnien N, Wimmer
EA, Bucher G. 2008 Divergent functions of
orthodenticle, empty spiracles and buttonhead
in early head patterning of the beetle Tribolium
castaneum (Coleoptera). Dev. Biol. 317, 600– 613.
(doi:10.1016/j.ydbio.2008.03.005)
67. Simonnet F, Célérier ML, Quéinnec E. 2006
Orthodenticle and empty spiracles genes are
expressed in a segmental pattern in chelicerates.
Dev. Genes Evol. 216, 467 –480. (doi:10.1007/
s00427-006-0093-4)
68. Kanayama M, Akiyama-Oda Y, Nishimura O, Tarui H,
Agata K, Oda H. 2011 Travelling and splitting of a
wave of hedgehog expression involved in spiderhead segmentation. Nat. Comm. 2, 500. (doi:10.
1038/ncomms1510)
9
Phil. Trans. R. Soc. B 371: 20150044
32.
44.
Cell 32, 408–422. (doi:10.1016/j.devcel.2015.
01.008)
Janssen R, Damen WG. 2006 The ten Hox genes of
the millipede Glomeris marginata. Dev. Genes Evol.
216, 451 –465. (doi:10.1007/s00427-006-0092-5)
Brena C, Chipman AD, Minelli A, Akam M. 2006
Expression of trunk Hox genes in the centipede
Strigamia maritima: sense and anti-sense transcripts.
Evol. Dev. 8, 252–265. (doi:10.1111/j.1525-142X.
2006.00096.x)
Schaeper ND, Pechmann M, Damen WG, Prpic N-M,
Wimmer EA. 2010 Evolutionary plasticity of collier
function in head development of diverse arthropods.
Dev. Biol. 344, 363 –376. (doi:10.1016/j.ydbio.
2010.05.001)
Janssen R, Eriksson BJ, Tait NN, Budd GE. 2014
Onychophoran Hox genes and the evolution of
arthropod Hox gene expression. Front. Zool. 11, 22.
(doi:10.1186/1742-9994-11-22)
Hughes CL, Kaufman TC. 2002 Hox genes and the
evolution of the arthropod body plan. Evol. Dev. 4,
459 –499. (doi:10.1046/j.1525-142X.2002.02034.x)
Merrill VKL, Diederich RJ, Turner FR, Kaufman TC.
1989 A genetic and developmental analysis of
mutations in labial, a gene necessary for proper
head formation in Drosophila melanogaster. Dev.
Biol. 135, 376 –391. (doi:10.1016/0012-1606(89)
90187-5)
Angelini DR, Liu PZ, Hughes CL, Kaufman TC. 2005
Hox gene function and interaction in the milkweed
bug Oncopeltus fasciatus (Hemiptera). Dev. Biol.
287, 440 –455. (doi:10.1016/j.ydbio.2005.08.010)
Posnien N, Bucher G. 2010 The Hox1 ortholog Tclabial is required for embryonic development of the
intercalary segment giving rise to lateral parts of
the head in Tribolium. Dev. Biol. 338, 107–116.
(doi:10.1016/j.ydbio.2009.11.010)
Denes AS, Jékely G, Steinmetz PRH, Raible F,
Snyman H, Prud’homme B, Ferrier DEK, Balavoine G,
Arendt D. 2007 Molecular architecture of annelid
nerve cord supports common origin of nervous
system centralization in Bilateria. Cell 129,
277 –288. (doi:10.1016/j.cell.2007.02.040)
Döffinger C, Stollewerk A. 2010 How can conserved
gene expression allow for variation? Lessons from
the dorso-ventral patterning gene muscle segment
homeobox. Dev. Biol. 345, 105 –116. (doi:10.1016/j.
ydbio.2010.06.012)
Biffar L, Stollewerk A. 2015 Evolutionary variations
in the expression of dorso-ventral patterning genes
and the conservation of pioneer neurons in
Tribolium castaneum. Dev. Biol. 400, 159–167.
(doi:10.1016/j.ydbio.2015.01.025)
Eriksson BJ, Stollewerk A. 2010 Expression patterns
of neural genes in Euperipatoides kanangrensis
suggest divergent evolution of onychophoran and
euarthropod neurogenesis. Proc. Natl Acad. Sci. USA
107, 22 576–22 581. (doi:10.1073/pnas.
1008822108)
Bate M. 1976 Embryogenesis of an insect nervous
system. I. A map of thoracic and abdominal
neuroblasts in Locusta migratoria. J. Embryol.
Exp. Morph. 35, 107–123.
rstb.royalsocietypublishing.org
30.
Perspect. Biol. 2, a000505. (doi:10.1101/cshperspect.
a000505)
Janssen R et al. 2010 Conservation, loss, and
redeployment of Wnt ligands in protostomes:
implications for understanding the evolution
of segment formation. BMC Evol. Biol. 10, 374.
(doi:10.1186/1471-2148-10-374)
Urbach R, Technau GM. 2003 Molecular markers for
identified neuroblasts in the developing brain of
Drosophila. Development 130, 3621 –3637. (doi:10.
1242/dev.00533)
Urbach R, Technau GM. 2003 Segment polarity and
D/V patterning gene expression reveals segmental
organization of the Drosophila brain. Development
130, 3607 –3620. (doi:10.1242/dev.00532)
Richter S, Hartmann B, Reichert H. 1998 The
wingless gene is required for embryonic brain
development in Drosophila. Dev. Genes Evol. 208,
37 –45. (doi:10.1007/s004270050151)
McDonald JA, Doe CQ. 1997 Establishing neuroblastspecific gene expression in the Drosophila CNS:
huckebein is activated by Wingless and Hedgehog
and repressed by Engrailed and Gooseberry.
Development 124, 1079 –1087.
Davis GK, Patel NH. 2002 Short, long, and beyond:
molecular and embryological approaches to insect
segmentation. Annu. Rev. Entomol. 47, 669–699.
(doi:10.1146/annurev.ento.47.091201.145251)
Schwager EE, Pechmann M, Feitosa NM, McGregor
AP, Damen WG. 2009 hunchback functions as a
segmentation gene in the spider Achaearanea
tepidariorum. Curr. Biol. 19, 1333–1340. (doi:10.
1016/j.cub.2009.06.061)
Dalton D, Chadwick R, McGinnis W. 1989 Expression
and embryonic function of empty spiracles: a
Drosophila homeo box gene with two patterning
functions on the anterior–posterior axis of the
embryo. Genes Dev. 3, 1940 –1956. (doi:10.1101/
gad.3.12a.1940)
Gao Q, Finkelstein R. 1998 Targeting gene
expression to the head: the Drosophila orthodenticle
gene is a direct target of the Bicoid morphogen.
Development 125, 4185 –4193.
Elkouris M, Balaskas N, Poulou M, Politis PK,
Panayiotou E, Malas S, Thomaidou D, Remboutsika
E. 2011 Sox1 maintains the undifferentiated state of
cortical neural progenitor cells via the suppression
of Prox1-mediated cell cycle exit and neurogenesis.
Stem Cells 29, 89 –98. (doi:10.1002/stem.554)
Bylund M, Andersson E, Novitch BG, Muhr J. 2003
Vertebrate neurogenesis is counteracted by Sox1 –3
activity. Nat. Neurosci. 6, 1162 –1168. (doi:10.1038/
nn1131)
Sasai Y. 2001 Roles of Sox factors in neural
determination: conserved signaling in evolution?
Int. J. Dev. Biol. 45, 321–326.
Pioro HL, Stollewerk A. 2006 The expression pattern
of genes involved in early neurogenesis suggests
distinct and conserved functions in the diplopod
Glomeris marginata. Dev. Genes Evol. 216,
417–430. (doi:10.1007/s00427-006-0078-3)
Jung H, Dasen JS. 2015 Evolution of patterning
systems and circuit elements for locomotion. Dev.
Downloaded from http://rstb.royalsocietypublishing.org/ on June 12, 2017
83.
84.
86.
87.
88.
89.
90.
91.
92.
93.
94.
95. LaBonne C, Bronner-Fraser M. 2000 Snail-related
transcriptional repressors are required in Xenopus
for both the induction of the neural crest and its
subsequent migration. Dev. Biol. 221, 195 –205.
(doi:10.1006/dbio.2000.9609)
96. Apitz H, Salecker I. 2015 A region-specific
neurogenesis mode requires migratory progenitors
in the Drosophila visual system. Nat. Neurosci. 18,
46– 55. (doi:10.1038/nn.3896)
97. Ayyar S, Negre B, Simpson P, Stollewerk A. 2010 An
arthropod cis-regulatory element functioning in
sensory organ precursor development dates back to
the Cambrian. BMC Biol. 8, 127. (doi:10.1186/17417007-8-127)
98. Chipman AD, Stollewerk A. 2006 Specification of
neural precursor identity in the geophilomorph
centipede Strigamia maritima. Dev. Biol. 290,
337–350. (doi:10.1016/j.ydbio.2005.11.029)
99. Biffar L, Stollewerk A. 2014 Conservation and
evolutionary modifications of neuroblast expression
patterns in insects. Dev. Biol. 388, 103–116.
(doi:10.1016/j.ydbio.2014.01.028)
100. Skeath JB, Thor S. 2003 Genetic control of
Drosophila nerve cord development. Curr. Opin.
Neurobiol. 13, 8–15. (doi:10.1016/S0959-4388
(03)00007-2)
101. Goodman CS. 1996 Mechanisms and molecules that
control growth cone guidance. Ann. Rev. Neurosci.
19, 341 –377. (doi:10.1146/annurev.ne.19.030196.
002013)
102. Broadus J, Skeath JB, Spana EP, Bossing T, Technau
G, Doe CQ. 1995 New neuroblast markers and the
origin of the aCC/pCC neurons in the Drosophila
central nervous system. Mech. Dev. 53, 393–402.
(doi:10.1016/0925-4773(95)00454-8)
103. Duman-Scheel M, Patel NH. 1999 Analysis of
molecular marker expression reveals neuronal
homology in distantly related arthropods.
Development 126, 2327–2334.
104. Brankatschk M, Dickson BJ. 2006 Netrins guide
Drosophila commissural axons at short range.
Nat. Neurosci. 9, 188 –194. (doi:10.1038/nn1625)
105. Linne V, Stollewerk A. 2011 Conserved and novel
functions for Netrin in the formation of the
axonal scaffold and glial sheath cells in spiders. Dev.
Biol. 353, 134–146. (doi:10.1016/j.ydbio.2011.02.
006)
106. Duman-Scheel M, Clark SM, Grunow ET, Hasley AO,
Hill BL, Simanton WL. 2007 Delayed onset of
midline netrin expression in Artemia franciscana
coincides with commissural axon growth and
provides evidence for homology of midline cells in
distantly related arthropods. Evol. Dev. 9, 131 –140.
(doi:10.1111/j.1525-142X.2007.00144.x)
107. Dickson B. 2002 Molecular mechanisms of axon
guidance. Science 298, 1959–1963. (doi:10.1126/
science.1072165)
108. Linne V, Eriksson BJ, Stollewerk A. 2012 Singleminded and the evolution of the ventral midline in
arthropods. Dev. Biol. 364, 66 –76. (doi:10.1016/j.
ydbio.2012.01.019)
10
Phil. Trans. R. Soc. B 371: 20150044
85.
renewal and differentation in Drosophila neural
stem cells. Dev. Cell 11, 775–789. (doi:10.1016/j.
devcel.2006.09.015)
Li L, Vaessin H. 2000 Pan-neural Prospero
terminates cell proliferation during Drosophila
neurogenesis. Genes Dev. 14, 147 –151.
Tepass U, Gruszynski-DeFeo E, Haag TA, Omatyar L,
Török T, Hartenstein V. 1996 shotgun encodes
Drosophila E-cadherin and is preferentially required
during cell rearrangement in the neurectoderm and
other morphogenetically active epithelia. Genes Dev.
10, 672– 685. (doi:10.1101/gad.10.6.672)
Schäfer G, Narasimha M, Vogelsang E, Leptin M.
2014 Cadherin switching during the formation and
differentiation of the Drosophila mesoderm—
implications for epithelial-to-mesenchymal
transitions. J. Cell. Sci. 127, 1511– 1522. (doi:10.
1242/jcs.139485)
Ungerer P, Eriksson BJ, Stollewerk A. 2012
Unravelling the evolution of neural stem cells in
arthropods: Notch signalling in neural stem cell
development in the crustacean Daphnia magna.
Dev. Biol. 371, 302 –311. (doi:10.1016/j.ydbio.
2012.08.025)
Ashraf SI, Ip YT. 2001 The Snail protein family
regulates neuroblast epxression of inscutable and
string, genes involved in asymmetry and cell
division in Drosophila. Development 128,
4747 –4767.
Zander MA, Burns SE, Yuang G, Kaplan D, Miller FD.
2014 Snail coordinately regulates downstream
pathways to control multiple aspects of mammalian
neural precursor development. J. Neurosci. 34,
5164 –5175. (doi:10.1523/JNEUROSCI.0370-14.2014)
Korzelius J et al. 2014 Escargot maintains stemness
and suppresses differentiation in Drosophila
intestinal stem cells. EMBO J. 33, 2967–2982.
(doi:10.15252/embj.201489072)
Kageyama R, Ohtsuka T, Shimojo H, Imayoshi I.
2009 Dynamic regulation of Notch signaling in
neural progenitor cells. Curr. Opin. Cell Biol. 21,
733 –740. (doi:10.1016/j.ceb.2009.08.009)
Gold K, Cotton JA, Stollewerk A. 2009 The role
of Notch signalling and numb function in
mechansensory organ formation in the spider
Cupiennius salei. Dev. Biol. 327, 121 –131. (doi:10.
1016/j.ydbio.2008.12.004)
Weller M, Tautz D. 2003 Prospero and Snail
expression during spider neurogenesis. Dev. Genes
Evol. 213, 554–566. (doi:10.1007/s00427-0030362-4)
Stollewerk A, Seyfarth E-A. 2008 Evolutionary
changes in sensory precursor formation in
arthropods: embryonic development of leg sensilla
in the spider Cupiennius salei. Dev. Biol. 313,
659 –673. (doi:10.1016/j.ydbio.2007.11.003)
Cheung M, Chaboissier MC, Mynett A, Hirst E,
Schedl A, Briscoe J. 2005 The transcriptional control
of trunk neural crest induction, survival, and
delamination. Dev. Cell 2, 179 –192. (doi:10.1016/j.
devcel.2004.12.010)
rstb.royalsocietypublishing.org
69. Eriksson BJ, Ungerer P, Stollewerk A. 2013 The function
of Notch signalling in segment formation in the
crustacean Daphnia magna (Branchiopoda). Dev. Biol.
383, 321–330. (doi:10.1016/j.ydbio.2013.09.021)
70. Fortini ME. 2009 Notch signaling: the core pathway
and its posttranslational regulation. Dev. Cell 16,
633–647. (doi:10.1016/j.devcel.2009.03.010)
71. del Alamo D, Rouault H, Schweisguth F. 2011
Mechanisms and significance of cis-inhibition in
Notch signalling. Curr. Biol. 21, R40–R47. (doi:10.
1016/j.cub.2010.10.034)
72. Formosan-Jordan P, Ibañes M. 2014 Competition in
Notch signaling with cis enriches cell fate decisions.
PLoS ONE 9, e95744. (doi:10.1371/journal.pone.
0095744)
73. Döffinger C, Hartenstein V, Stollewerk A. 2010
Compartmentalisation of the precheliceral
neuroectoderm in the spider Cupiennius salei:
Development of the arcuate body, the optic ganglia
and the mushroom body. J. Comp. Neurol. 518,
2612 –2632.
74. Hartenstein AY, Nassif C, Green P, Hartenstein V.
1996 Early neurogenesis of the Drosophila brain.
J. Comp. Neurol. 370, 313–329. (doi:10.1002/
(SICI)1096-9861(19960701)370:3,313::AIDCNE3.3.0.CO;2-7)
75. Egger B, Gold KS, Brand AH. 2010 Notch regulates
the switch from symmetric to asymmetric neural
stem cell division in the Drosophila optic lobe.
Development 137, 2981 –2987. (doi:10.1242/
dev.051250)
76. zur Lage PI, Jarman AP. 1999 Antagonism of EGFR
and Notch signalling in the reiterative recruitment
of Drosophila adult chordotonal sense organ
precursors. Development 126, 3149 –3157.
77. Goodman CS, Doe CQ. 1993 Embryonic development
of the Drosophila central nervous system. In The
development of Drosophila melanogaster (eds
M Bate, A Martinez-Arias), pp. 1131 –1206.
New York, NY: Cold Spring Harbor Laboratory Press.
78. Brenneis G, Stollewerk A, Scholtz G. 2013 Embryonic
neurogenesis in Pseudopallene sp. (Arthropoda,
Pycnogonida) includes two subsequent phases with
similarities to different arthropod groups. Evodevo 4,
32. (doi:10.1186/2041-9139-4-32)
79. Cai Y, Chia W, Yang X. 2001 A family of snail-related
zinc finger proteins regulates two distinct and
parallel mechanisms that mediate Drosophila
neuroblast asymmetric divisions. EMBO J. 20,
1704–1714. (doi:10.1093/emboj/20.7.1704)
80. Southall TD, Brand AH. 2009 Neural stem cell
transcriptional networks highlight genes essential
for nervous system development. EMBO J. 28,
3799–3807. (doi:10.1038/emboj.2009.309)
81. Broadus J, Fuerstenberg S, Doe CQ. 1998 Staufendependent localization of prospero mRNA
contributes to neuroblast daughter-cell fate. Nature
391, 792–795. (doi:10.1038/35861)
82. Choksi SP, Southall TD, Bossing T, Edoff K, de Wit E,
Fischer BE, van Steensel B, Micklem G, Brand AH.
2006 Prospero acts as a binary switch between self-