1 The evolutionary origin of bilaterian smooth and striated

Transcription

1 The evolutionary origin of bilaterian smooth and striated
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
The evolutionary origin of bilaterian smooth and striated myocytes
2
3
4
Thibaut Brunet1, Antje H. L. Fischer1,2, Patrick R. H. Steinmetz1,3, Antonella Lauri1,4,
5
Paola Bertucci1, and Detlev Arendt1,*
6
7
8
1.
9
10
Meyerhofstraße, 1, 69117 Heidelberg, Germany
2.
11
12
15
Present address: Ludwig-Maximilians University Munich, Biochemistry, FeodorLynen Straße 17, 81377 Munich
3.
13
14
Developmental Biology Unit, European Molecular Biology Laboratory,
Present address: Sars International Centre for Marine Molecular Biology,
University of Bergen, Thormøhlensgate 55, 5008 Bergen, Norway.
4.
Present address: Institute for Biological and Medical Imaging (IBMI), Helmholtz
Zentrum München, Neuherberg, Germany
16
17
* For correspondence: [email protected]
18
19
1
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Abstract
2
The dichotomy between smooth and striated myocytes is fundamental for bilaterian
3
musculature, but its evolutionary origin remains unsolved. Given their absence in
4
Drosophila and Caenorhabditis, smooth muscles have so far been considered a
5
vertebrate innovation. Here, we characterize expression profile, ultrastructure,
6
contractility and innervation of the musculature in the marine annelid Platynereis
7
dumerilii and identify smooth muscles around the midgut, hindgut and heart that
8
resemble their vertebrate counterparts in molecular fingerprint, contraction speed,
9
and nervous control. Our data suggest that both visceral smooth and somatic
10
striated myocytes were present in the protostome-deuterostome ancestor, and that
11
smooth myocytes later co-opted the striated contractile module repeatedly – for
12
example in vertebrate heart evolution. During these smooth-to-striated myocyte
13
conversions the core regulatory complex of transcription factors conveying myocyte
14
identity remained unchanged, reflecting a general principle in cell type evolution.
15
16
2
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Introduction
2
Musculature is composed of myocytes that are specialized in active contraction (Schmidt-
3
Rhaesa 2007). This function relies on actomyosin, a contractile module that dates back to
4
stem eukaryotes (Brunet & Arendt 2016) and incorporated accessory proteins of pre-
5
metazoan origin (Steinmetz et al. 2012). Two fundamentally distinct types of myocytes
6
are distinguished based on ultrastructural appearance. In striated myocytes, actomyosin
7
myofibrils are organized in aligned repeated units (sarcomeres) separated by transverse
8
‘Z discs’, while in smooth myocytes adjacent myofibrils show no clear alignment and are
9
separated by scattered “dense bodies” (Figure 1A). In vertebrates, striated myocytes are
10
found in voluntary skeletal muscles, but also at the anterior and posterior extremities of
11
the digestive tract (anterior esophagus muscles and external anal sphincter), and in the
12
muscular layer of the heart; smooth myocytes are found in involuntary visceral
13
musculature that ensures slow, long-range deformation of internal organs. This includes
14
the posterior esophagus and the rest of the gut, but also blood vessels, and most of the
15
urogenital system. In stark contrast, in the fruit fly Drosophila virtually all muscles are
16
striated, including gut visceral muscles (Anderson & Ellis 1967; Goldstein & Burdette
17
1971; Paniagua et al. 1996); the only exception are poorly-characterized multinucleated
18
smooth muscles around the testes (Susic-Jung et al. 2012). Also, in the nematode
19
Caenorhabditis, somatic muscles are striated, while the short intestine and rectum
20
visceral myocytes are only one sarcomere-long and thus hard to classify (White 1988;
21
Corsi et al. 2000).
3
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
The evolutionary origin of smooth versus striated myocytes in bilaterians accordingly
2
remains unsolved. Ultrastructural studies have consistently documented the presence of
3
striated somatic myocytes in virtually every bilaterian group (Schmidt-Rhaesa 2007) and
4
in line with this, the comparison of Z-disc proteins supports homology of striated
5
myocytes across bilaterians (Steinmetz et al. 2012). The origin of smooth myocyte types
6
however is less clear. Given the absence of smooth muscles from fly and nematode, it has
7
been proposed that visceral smooth myocytes represent a vertebrate novelty, which
8
evolved independently from non-muscle cells in the vertebrate stem line (Goodson &
9
Spudich 1993; OOta & Saitou 1999). However, smooth muscles are present in many
10
other bilaterian groups, suggesting instead their possible presence in urbilaterians and
11
secondary loss in arthropods and nematodes. Complicating the matter further,
12
intermediate ultrastructures between smooth and striated myocytes have been reported,
13
suggesting interconversions (reviewed in (Schmidt-Rhaesa 2007))
14
Besides ultrastructure, the comparative molecular characterization of cell types can be
15
used to build cell type trees (Arendt 2003; Arendt 2008; Wagner 2014; Musser & Wagner
16
2015). Cell type identity is established via the activity of transcription factors acting as
17
terminal selectors (Hobert 2016) and forming “core regulatory complexes” (CRCs;
18
(Arendt et al. 2016; Wagner 2014)), which directly activate downstream effector genes.
19
This is exemplified for vertebrate myocytes in Figure 1B. In all vertebrate myocytes,
20
transcription factors of the Myocardin family (MASTR in skeletal muscles, Myocardin in
21
smooth and cardiac muscles) directly activate effector genes encoding contractility
22
proteins (Fig. 1B) (Creemers et al. 2006; Meadows et al. 2008; Wang & Olson 2004;
23
Wang et al. 2003)). They heterodimerize with MADS-domain factors of the Myocyte
4
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Enhancer Factor-2 (Mef2) (Black & Olson 1998; Blais et al. 2005; Molkentin et al. 1995;
2
Wales et al. 2014) and Serum Response Factor (SRF) families (Carson et al. 1996;
3
Nishida et al. 2002). Other myogenic transcription factors are specific for different types
4
of striated and smooth myocytes. Myogenic Regulatory Factors (MRF) family members,
5
including MyoD and its paralogs Myf5, Myogenin, and Mrf4/Myf6 (Shi & Garry 2006),
6
directly control contractility effector genes in skeletal (and esophageal) striated
7
myocytes, cooperatively with Mef2 (Blais et al. 2005; Molkentin et al. 1995) – but are
8
absent from smooth and cardiac muscles. In smooth and cardiac myocytes, this function
9
is ensured by NK transcription factors (Nkx3.2/Bapx and Nkx2.5/Tinman, respectively),
10
GATA4/5/6, and Fox transcription factors (FoxF1 and FoxC1, respectively), which bind
11
to SRF and Mef2 to form CRCs directly activating contractility effector genes (Durocher
12
et al. 1997; Hoggatt et al. 2013; Lee et al. 1998; Morin et al. 2000; Nishida et al. 2002;
13
Phiel et al. 2001) (Figure 1B).
14
Regarding effector proteins (Figure 1B), (Kierszenbaum & Tres 2015), all myocytes
15
express distinct isoforms of the myosin heavy chain: the striated myosin heavy chain ST-
16
MHC (which duplicated into cardiac, fast skeletal, and slow skeletal isoforms in
17
vertebrates) and the smooth/non-muscle myosin heavy chain SM-MHC (which duplicated
18
in vertebrates into smooth myh10, myh11 and myh14, and non-muscle myh9) (Steinmetz
19
et al. 2012). The different contraction speeds of smooth and striated muscles are due to
20
the distinct kinetic properties of these molecular motors (Bárány 1967). In both myocyte
21
types, contraction occurs in response to calcium, but the responsive proteins differ
22
(Alberts et al. 2014): the Troponin complex (composed of Troponin C, Troponin T and
23
Troponin I) for striated muscles, Calponin and Caldesmon for smooth muscles. In both
5
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
myocyte types, calcium also activates the Calmodulin/Myosin Light Chain Kinase
2
pathway (Kamm & Stull 1985; Sweeney et al. 1993). Striation itself is implemented by
3
specific effectors, including the long elastic protein Titin (Labeit & Kolmerer 1995)
4
(which spans the entire sarcomere and confers it elasticity and resistance) and
5
ZASP/LBD3 (Z-band Alternatively Spliced PDZ Motif/LIM-Binding Domain 3), which
6
binds actin and stabilizes sarcomeres during contraction (Au et al. 2004; Zhou et al.
7
2001). The molecular study of Drosophila and Caenorhabditis striated myocytes
8
revealed important commonalities with their vertebrate counterparts, including the
9
Troponin complex (Beall & Fyrberg 1991; Fyrberg et al. 1994; Fyrberg et al. 1990;
10
Marín et al. 2004; Myers et al. 1996), and a conserved role for Titin (Zhang et al. 2000)
11
and ZASP/LBD3 (Katzemich et al. 2011; McKeown et al. 2006) in the striated
12
architecture.
13
Finally, smooth and striated myocytes also differ physiologically. All known striated
14
myocyte types (apart from the myocardium) strictly depend on nervous stimulations for
15
contraction, exerted by innervating motor neurons. In contrast, gut smooth myocytes are
16
able to generate and propagate automatic (or “myogenic”) contraction waves responsible
17
for digestive peristalsis in the absence of nervous inputs (Faussone‐Pellegrini &
18
Thuneberg 1999; Sanders et al. 2006). These autonomous contraction waves are
19
modulated by the autonomic nervous system (Silverthorn 2015). Regarding overall
20
contraction speed, striated myocytes have been measured to contract 10 to 100 times
21
faster than their smooth counterparts (Bárány 1967).
22
6
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
To elucidate the evolutionary origin and diversification of bilaterian smooth and striated
2
myocytes,
3
characterization of the myocyte complement in the marine annelid Platynereis dumerilii,
4
which belongs to the Lophotrochozoa. Strikingly, as of now, no invertebrate smooth
5
visceral muscle has been investigated on a molecular level (Hooper & Thuma 2005;
6
Hooper et al. 2008). Platynereis has retained more ancestral features than flies or
7
nematodes and is thus especially suited for long-range comparisons (Raible et al. 2005;
8
Denes et al. 2007). Also, other annelids such as earthworms have been reported to
9
possess both striated somatic and midgut smooth visceral myocytes based on electron
10
microscopy (Anderson & Ellis 1967). Our study reveals the parallel presence of smooth
11
myocytes in the musculature of midgut, hindgut, and pulsatile dorsal vessel and of
12
striated myocytes in the somatic musculature and the foregut. Platynereis smooth and
13
striated myocytes closely parallel their vertebrate counterparts in ultrastructure, molecular
14
profile, contraction speed, and reliance on nervous inputs, thus supporting the ancient
15
existence of a smooth-striated duality in protostome/deuterostome ancestors.
we
provide
an
in-depth
ultrastructural,
molecular
and
functional
16
17
18
Results
19
Platynereis midgut and hindgut muscles are smooth, while foregut and somatic muscles
20
are striated
21
Differentiation of the Platynereis somatic musculature has been documented in much
22
detail (Fischer et al. 2010) and, in five days old young worms, consists of ventral and
23
dorsal longitudinal muscles, oblique and parapodial muscles, head muscles and the
7
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
axochord (Lauri et al. 2014). At this stage, the first Platynereis visceral myocytes become
2
detectable around the developing tripartite gut, which is subdivided into foregut, midgut
3
and hindgut (based on the conserved regional expression of foxA, brachyury and hnf4 gut
4
specification factors (Martín-Durán & Hejnol 2015); Figure 2-supplement 1). At 7 days
5
post fertilization (dpf), visceral myocytes form circular myofibres around the foregut, and
6
scattered longitudinal and circular fibres around midgut and hindgut (Figure 2A), which
7
expand by continuous addition of circular and longitudinal fibres to completely cover the
8
dorsal midgut at 11dpf (Figure 2A) and finally form a continuous muscular orthogon
9
around the entire midgut and hindgut in the 1.5 months-old juvenile (Figure 2A).
10
We then proceeded to characterize the ultrastructure of Platynereis visceral and somatic
11
musculature by transmission electron microscopy (Figure 2C-M). All somatic muscles
12
and anterior foregut muscles display prominent oblique striation with discontinuous Z-
13
elements (Figure 2C-H; compare Figure 1A), as typical for protostomes (Burr & Gans
14
1998; Mill & Knapp 1970; Rosenbluth 1972). To the contrary, visceral muscles of the
15
posterior foregut, midgut and hindgut are smooth with scattered dense bodies (Figure 2I-
16
M). The visceral muscular orthogon is partitioned into an external longitudinal layer and
17
an internal circular layer (Figure 2J), as in vertebrates (Marieb & Hoehn 2015) and
18
arthropods (Lee et al. 2006). Thus, according to ultrastructural appearance, Platynereis
19
has both somatic (and anterior foregut) striated muscles and visceral smooth muscles.
20
The molecular profile of smooth and striated myocytes
21
We then set out to molecularly characterize annelid smooth and striated myocytes via a
22
candidate gene approach. As a starting point, we investigated, in the Platynereis genome,
23
the presence of regulatory and effector genes specific for smooth and/or striated
8
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
myocytes in the vertebrates. We found striated muscle-specific and smooth muscle/non-
2
muscle isoforms of both myosin heavy chain (consistently with published phylogenies
3
(Steinmetz et al. 2012)) and myosin regulatory light chain. We also identified homologs
4
of genes encoding calcium transducers (calponin for smooth muscles; troponin I and
5
troponin T for striated muscles), striation structural proteins (zasp/lbd3 and titin), and
6
terminal selectors for the smooth (nk3, foxF, and gata456) and striated phenotypes
7
(myoD).
8
We investigated expression of these markers by whole-mount in situ hybridization
9
(WMISH). Striated effectors are expressed in both somatic and foregut musculature
10
(Figure 3A,C; Figure 3—supplement 1). Expression of all striated effectors was observed
11
in every somatic myocyte group by confocal imaging with cellular resolution (Figure 3—
12
supplement 2). Interestingly, myoD is exclusively expressed in longitudinal striated
13
muscles, but not in other muscle groups (Figure 3—supplement 2).
14
The expression of smooth markers is first detectable at 3 dpf in a small triangle-shaped
15
group of mesodermal cells posteriorly abutting the macromeres (which will form the
16
future gut) (Figure 3B, Figure 3—supplement 3A-D,H). Double WMISH shows that
17
these cells coexpress smooth markers and the muscle marker actin (Figure 3—
18
supplement 3I-K), supporting the idea that they are forming gut myocytes. At this stage,
19
smooth markers are also expressed in the foregut mesoderm (Figure 3B, Figure 3—
20
supplement 3A-D,H, yellow arrows). At 6 dpf, expression of all smooth markers (apart
21
from nk3) is maintained in the midgut and hindgut differentiating myocytes (Figure 3D,
22
Figure 3—supplement 3E-G, Figure 3—supplement 4A-E) but smooth effectors
23
disappear from the foregut, which turns on striated markers instead (Figure 3—
9
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
supplement 1R-W) – reminiscent of the replacement of smooth fibres by striated fibres
2
during development of the vertebrate anterior esophageal muscles (Gopalakrishnan et al.
3
2015). Finally, in 2 months-old juvenile worms, smooth markers are also detected in the
4
dorsal pulsatile vessel (Figure 3—supplement 3L-Q) – considered equivalent to the
5
vertebrate heart (Saudemont et al. 2008) but, importantly, of smooth ultrastructure in
6
polychaetes (Spies 1973; Jensen 1974). None of the striated markers is expressed around
7
the midgut or the hindgut (Figure 3—supplement 4F-K), or in the dorsal vessel (Figure
8
3—supplement 3P). Taken together, these results strongly support conservation of the
9
molecular fingerprint of both smooth and striated myocytes between annelids and
10
vertebrates.
11
We finally investigated general muscle markers that are shared between smooth and
12
striated muscles. These include actin, mef2 and myocardin – which duplicated into
13
muscle type-specific paralogs in vertebrates, but are still present as single-copy genes in
14
Platynereis. We found them to be expressed in the forming musculature throughout larval
15
development (Figure 3—supplement 5A-F), and confocal imaging at 6 dpf confirmed
16
expression of all 3 markers in both visceral (Figure 3—supplement 5G-L) and somatic
17
muscles (Figure 3—supplement 5M).
18
Smooth and striated muscles differ in contraction speed
19
We then characterized the contraction speed of the two myocyte types in Platynereis by
20
measuring myofibre length before and after contraction. Live confocal imaging of
21
contractions in Platynereis larvae with fluorescently labeled musculature (Movie 1,
22
Movie 2) gave a striated contraction rate of 0.55±0.27 s-1 (Figure 4A-E) and a smooth
23
myocyte contraction rate of 0.07±0.05 s-1 (Figure 4G). As in vertebrates, annelid striated
10
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
myocytes thus contract nearly one order of magnitude faster than smooth myocytes
2
(Figure 4F).
3
Striated, but not smooth, muscle contraction depends on nervous inputs
4
Finally, we investigated the nervous control of contraction of both types of muscle cells.
5
In vertebrates, somatic muscle contraction is strictly dependent on neuronal inputs. By
6
contrast, gut peristalsis is automatic (or myogenic – i.e., does not require nervous inputs)
7
in vertebrates, cockroaches (Nagai & Brown 1969), squids (Wood 1969), snails (Roach
8
1968), holothurians, and sea urchins (Prosser et al. 1965). The only exceptions appear to
9
be bivalves and malacostracans (crabs, lobster and crayfish), in which gut motility is
10
neurogenic (Prosser et al. 1965). Regardless of the existence of an automatic component,
11
the gut is usually innervated by nervous fibres modulating peristalsis movements (Wood
12
1969; Wu 1939).
13
Gut peristalsis takes place in Platynereis larvae and juveniles from 6 dpf onwards (Movie
14
3), and we set out to test whether nervous inputs were necessary for it to take place. We
15
treated 2 months-old juveniles with 180 µM Brefeldin A, an inhibitor of vesicular traffic
16
which prevents polarized secretion (Misumi et al. 1986) and neurotransmission (Malo et
17
al. 2000). Treatment stopped locomotion in all treated individuals, confirming that
18
neurotransmitter release by motor neurons is required for somatic muscles contraction,
19
while DMSO-treated controls were unaffected. On the other hand, vigorous gut
20
peristalsis movements were maintained in Brefeldin A-treated animals (Movie 4).
21
Quantification of the propagation speed of the peristalsis wave (Figure 5A-D; see
22
Material and Methods) indicated that contractions propagated significantly faster in
23
Brefeldin A-treated individuals than in controls. The frequency of wave initiation and
11
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
their recurrence (the number of repeated contraction waves occurring in one
2
uninterrupted sequence) did not differ significantly in Brefeldin A-treated animals
3
(Figure 5E,F). These results indicate that, as in vertebrates, visceral smooth muscle
4
contraction and gut peristalsis do not require nervous (or secretory) inputs in Platynereis.
5
An enteric nervous system is present in Platynereis
6
In vertebrates, peristaltic contraction waves are initiated by self-excitable myocytes
7
(Interstitial Cajal Cells) and propagate across other smooth muscles by gap junctions
8
ensuring direct electrical coupling (Faussone‐Pellegrini & Thuneberg 1999; Sanders et
9
al. 2006). We tested the role of gap junctions in Platynereis gut peristalsis by treating
10
animals with 2.5 mM 2-octanol, which inhibits gap junction function in both insects
11
(Bohrmann & Haas-Assenbaum 1993; Gho 1994) and vertebrates (Finkbeiner 1992). 2-
12
octanol abolishes gut peristalsis, both in the absence and in the presence of Brefeldin A
13
(Figure 5G), indicating that propagation of the peristalsis wave relies on direct coupling
14
between smooth myocytes via gap junctions.
15
The acceleration of peristalsis upon Brefeldin A treatment suggests that secretory (and
16
possibly neural) inputs modulate gut contractions (with a net inhibitory effect). We thus
17
investigated the innervation of the Platynereis gut. Immunostainings of juvenile worms
18
for acetylated tubulin revealed a dense, near-orthogonal nerve net around the entire gut
19
(Figure 6A), which is tightly apposed to the visceral muscle layer (Figure 6C) and
20
includes serotonergic neurites (Figure 6B,C) and cell bodies (Figure 6D). Interestingly,
21
some enteric serotonergic cell bodies are devoid of neurites, thus resembling the
22
vertebrate (non-neuronal) enterochromaffine cells – endocrine serotonergic cells residing
12
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
around the gut and activating gut peristalsis by direct serotonin secretion upon
2
mechanical stretch (Bulbring & Crema 1959).
3
Discussion
4
Smooth and striated myocyte coexisted in bilaterian ancestors
5
Our study represents the first molecular characterization of protostome visceral smooth
6
musculature (Hooper & Thuma 2005; Hooper et al. 2008). The conservation of molecular
7
signatures for both smooth and striated myocytes indicates that a dual musculature
8
already existed in bilaterian ancestors: a fast striated somatic musculature (possibly also
9
present around the foregut – as in Platynereis, vertebrates (Gopalakrishnan et al. 2015)
10
and sea urchins (Andrikou et al. 2013; Burke 1981)), under strict nervous control; and a
11
slow smooth visceral musculature around the midgut and hindgut, able to undergo
12
automatic peristalsis thanks to self-excitable myocytes directly coupled by gap junctions,
13
and undergoing modulation by nervous and paracrine inputs. In striated myocytes, a core
14
regulatory complex (CRC) involving Mef2 and Myocardin directly activated striated
15
contractile effector genes such as ST-MHC, ST-MRLC and the Troponin genes (Figure
16
7—supplement 1). Notably, myoD might have been part of the CRC in only part of the
17
striated myocytes, as it is only detected in longitudinal muscles in Platynereis. The
18
absence of myoD expression in other annelid muscle groups is in line with the “chordate
19
bottleneck” concept (Thor & Thomas 2002), according to which specialization for
20
undulatory swimming during early chordate evolution would have fostered exclusive
21
reliance on trunk longitudinal muscles, and loss of other muscle types. In smooth
22
myocytes, a CRC composed of NK3, FoxF and GATA4/5/6 together with Mef2 and
13
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Myocardin activated the smooth contractile effectors SM-MHC, SM-MRLC and calponin
2
(Figure 7—supplement 1). In spite of their absence in flies and nematodes, gut myocytes
3
of smooth ultrastructure are widespread in other bilaterians, and an ancestral state
4
reconstruction retrieves them as present in the last common protostome/deuterostome
5
ancestor with high confidence (Figure 7—supplement 2), supporting our homology
6
hypothesis.
7
immunoreactivity in intestinal muscles of earthworms (Royuela et al. 1997) and snails
8
(which also lack immunoreactivity for Troponin T) (Royuela et al. 2000).
Our
results
are
consistent
with
previous
reports
of
Calponin
9
10
Origin of smooth and striated myocytes by cell type individuation
11
How did smooth and striated myocytes diverge in evolution? Figure 7 presents a
12
comprehensive cell type tree for the evolution of myocytes, with a focus on Bilateria.
13
This tree illustrates the divergence of the two muscle cell types by progressive
14
partitioning of genetic information in evolution – a process called individuation (Wagner
15
2014; Arendt et al. 2016). The individuation of fast and slow contractile cells involved
16
two complementary processes: (1) changes in CRC (black circles, Figure 7) and (2)
17
emergence of novel genes encoding new cellular modules, or apomeres (Arendt et al.
18
2016) (grey squares, Figure 7).
19
Around a common core formed by the Myocardin:Mef2 complex (both representing
20
transcription factors of pre-metazoan ancestry (Steinmetz et al. 2012)), smooth and
21
striated CRCs incorporated different transcription factors implementing the expression of
22
distinct effectors (Figure 1B; Figure 7—supplement 1) – notably the bilaterian-specific
14
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
bHLH factor MyoD (Steinmetz et al. 2012) and GATA4/5/6, which arose by bilaterian-
2
specific duplication of a single ancient pan-endomesodermal GATA transcription factor
3
(Martindale et al. 2004; Leininger et al. 2014).
4
Regarding the evolution of myocyte-specific apomeres, one prominent mechanism of
5
divergence has been gene duplications. While the MHC duplication predated metazoans,
6
other smooth and striated-specific paralogs only diverged in bilaterians. Smooth and
7
striated MRLC most likely arose by gene duplication in the bilaterian stem-line
8
(Supplementary Figure 1). Myosin essential light chain, actin and myocardin paralogs
9
split even later, in the vertebrate stem-line (Figure 7). Similarly, smooth and non-muscle
10
mhc and mrlc paralogs only diverged in vertebrates. The calponin-encoding gene
11
underwent parallel duplication and subfunctionalization in both annelids and chordates,
12
giving rise to both specialized smooth muscle paralogs and more broadly expressed
13
copies with a different domain structure (Figure 7—supplement 3). This slow and
14
stepwise nature of the individuation process is consistent with studies showing that
15
recently evolved paralogs can acquire differential expression between tissues that
16
diverged long before in evolution (Force et al. 1999; Lan & Pritchard 2016).
17
Complementing gene duplication, the evolution and selective expression of entirely new
18
apomeres also supported individuation: for example, Titin and all components of the
19
Troponin complex are bilaterian novelties (Steinmetz et al. 2012). In vertebrates, the new
20
gene caldesmon was incorporated in the smooth contractile module (Steinmetz et al.
21
2012).
22
15
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Smooth to striated myocyte conversion
2
Strikingly, visceral smooth myocytes were previously assumed to be a vertebrate
3
innovation, as they are absent in fruit flies and nematodes (two groups which are in fact
4
exceptions in this respect, at least from ultrastructural criteria (Figure 7—supplement
5
2A)). This view received apparent support from the fact that the vertebrate smooth and
6
non-muscle myosin heavy chains (MHC) arose by vertebrate-specific duplication of a
7
unique ancestral bilaterian gene, orthologous to Drosophila non-muscle MHC (Goodson
8
& Spudich 1993) – which our results suggest reflects instead gradual individuation of
9
pre-existing cell types (see above). Strikingly, the striated gut muscles of Drosophila
10
closely resemble vertebrate and annelid smooth gut muscles by transcription factors
11
(nk3/bagpipe (Azpiazu & Frasch 1993), foxF/biniou (Jakobsen et al. 2007a; Zaffran et al.
12
2001)), even though they express the fast/striated contractility module (Fyrberg et al.
13
1994; Fyrberg et al. 1990; Marín et al. 2004). If smooth gut muscles are ancestral for
14
protostomes, as our results indicate, this suggests that the smooth contractile module was
15
replaced by the fast/striated module in visceral myocytes during insect evolution.
16
Interestingly, chromatin immunoprecipitation assays (Jakobsen et al. 2007a) show that
17
the conserved visceral transcription factors foxF/biniou and nk3/bagpipe do not directly
18
control contractility genes in Drosophila gut muscles (which are downstream mef2
19
instead), but establish the morphogenesis and innervation of the visceral muscles, and
20
control non-contractile effectors such as gap junctions – which are the properties these
21
muscles seem to have conserved from their smooth ancestors. The striated gut myocytes
22
of insects would thus represent a case of co-option of an effector module from another
16
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
cell type, which happened at an unknown time during ecdysozoan evolution (Figure 7;
2
Figure 7—supplement 1).
3
Another likely example of co-option is the vertebrate heart: vertebrate cardiomyocytes
4
are striated and express fast myosin and troponin, but resemble smooth myocytes by
5
developmental origin (from the splanchnopleura), function (automatic contraction and
6
coupling by gap junctions) and terminal selector profile (Figure 1B). These similarities
7
suggest that cardiomyocytes might stem from smooth myocytes that likewise co-opted
8
the fast/striation module. Indicative of this possible ancestral state, the Platynereis dorsal
9
pulsatile vessel (considered homologous to the vertebrate heart based on comparative
10
anatomy and shared expression of NK4/tinman (Saudemont et al. 2008)) expresses the
11
smooth, but not the striated, myosin heavy chain (Figure 3—supplement 3O-Q). An
12
ancestral state reconstruction based on ultrastructural data further supports the notion that
13
heart myocytes were smooth in the last common protostome/deuterostome ancestor, and
14
independently acquired striation in at least 5 descendant lineages (Figure 7—supplement
15
2B) – usually in species with large body size and/or fast metabolism.
16
17
Striated to smooth conversions
18
Smooth somatic muscles are occasionally found in bilaterians with slow or sessile
19
lifestyles – for example in the snail foot (Faccioni-Heuser et al. 1999; Rogers 1969), the
20
ascidian siphon (Meedel & Hastings 1993), and the sea cucumber body wall (Kawaguti &
21
Ikemoto 1965). As an extreme (and isolated) example, flatworms lost striated muscles
22
altogether, and their body wall musculature is entirely smooth (Rieger et al. 1991).
17
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Interestingly, in all cases that have been molecularly characterized, smooth somatic
2
muscles express the same fast contractility module as their striated counterparts,
3
including ST-MHC and the Troponin complex – in ascidians (Endo & Obinata 1981;
4
Obinata et al. 1983), flatworms (Kobayashi et al. 1998; Witchley et al. 2013; Sulbarán et
5
al. 2015), and the smooth myofibres of the bivalve catch muscle (Nyitray et al. 1994;
6
Ojima & Nishita 1986). (It is unknown whether these also express zasp and titin in spite
7
of the lack of striation). This suggests that these are somatic muscles having secondarily
8
lost striation (in line with the sessile lifestyle of ascidians and bivalves, and with the
9
complete loss of striated muscles in flatworms). Alternatively, they might represent
10
remnants of ancestral smooth somatic fibres that would have coexisted alongside striated
11
somatic fibres in the last common protostome/deuterostome ancestor. Interestingly, the
12
fast contractile module is also expressed in acoel body wall smooth muscles (Chiodin et
13
al. 2011); since acoels belong to a clade that might have branched off before all other
14
bilaterians (Cannon et al. 2016) (but see (Telford & Copley 2016)), these could represent
15
fast-contracting myocytes that never evolved striation in the first place, similar to those
16
found in cnidarians. In all cases, the fast contractility module appears to represent a
17
consistent synexpression group (i.e. its components are reliably expressed together), and
18
a stable molecular profile of all bilaterian somatic muscles, regardless of the presence of
19
morphologically overt striation. This confirms the notion that, even in cases of
20
ambiguous morphology or ultrastructure, the molecular fingerprint of cell types holds
21
clue to their evolutionary affinities.
18
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Implications for cell type evolution
2
In the above, genetically well-documented cases of cell type conversion (smooth to
3
striated conversion in insect visceral myocytes and vertebrate cardiomyocytes), cells kept
4
their ancestral CRC of terminal selector transcription factors, while changing the
5
downstream effector modules. This supports the recent notion that CRCs confer an
6
abstract identity to cell types, which remains stable in spite of turnover in downstream
7
effectors (Wagner 2014) – just as hox genes impart conserved abstract identity to
8
segments of vastly diverging morphologies (Deutsch 2005). Tracking cell type-specific
9
CRCs through animal phylogeny thus represents a powerful means to decipher the
10
evolution of cell types.
11
Pre-bilaterian origins
12
If the existence of fast-contracting striated and slow-contracting smooth myocytes
13
predated bilaterians – when and how did these cell types first split in evolution? The first
14
evolutionary event that paved the way for the diversification of the smooth and striated
15
contractility modules was the duplication of the striated myosin heavy chain-encoding
16
gene into the striated isoform ST-MHC and the smooth/non-muscle isoform SM-MHC.
17
This duplication occurred in single-celled ancestors of animals, before the divergence of
18
filastereans and choanoflagellates (Steinmetz et al. 2012). Consistently, both sm-mhc and
19
st-mhc are present in the genome of the filasterean Ministeria (though st-mhc was lost in
20
other single-celled holozoans) (Sebé-Pedrós et al. 2014). Interestingly, st-mhc and sm-
21
mhc expression appears to be segregated into distinct cell types in sponges, cnidarians
22
(Steinmetz et al. 2012), and ctenophores (Dayraud et al. 2012), suggesting that a cell type
19
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
split between slow and fast contractile cells is a common feature across early-branching
2
metazoans (Figure 7). Given the possibility of MHC isoform co-option (as outlined
3
above), it is yet unclear whether this split happened once or several times. The affinities
4
of bilaterians and non-bilaterians contractile cells remain to be tested from data on the
5
CRCs establishing contractile cell types in non-bilaterians.
6
7
Conclusions
8
Our results indicate that the split between visceral smooth myocytes and somatic striated
9
myocytes is the result of a long individuation process, initiated before the last common
10
protostome/deuterostome ancestor. Fast- and slow-contracting cells expressing distinct
11
variants of myosin II heavy chain (ST-MHC versus SM-MHC) acquired increasingly
12
contrasted molecular profiles in a gradual fashion – and this divergence process continues
13
to this day in individual bilaterian phyla. Blurring this picture of divergence, co-option
14
events have led to the replacement of the slow contractile module by the fast one, leading
15
to smooth-to-striated myocyte conversions. Our study showcases the power of molecular
16
fingerprint comparisons centering on effector and selector genes to reconstruct cell type
17
evolution (Arendt 2008). In the bifurcating phylogenetic tree of animal cell types (Liang
18
et al. 2015), it remains an open question how the two types of contractile cells relate to
19
other cell types, such as neurons (Mackie 1970) or cartilage (Lauri et al. 2014).
20
21
22
20
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Material and Methods
2
3
Immunostainings and in situ hybridizations
4
Immunostaining, rhodamine-phalloidin staining, and simple and double WMISH were
5
performed according to previously published protocols (Lauri et al. 2014). For all
6
stainings not involving phalloidin, the following modifications: animals were mounted in
7
97% TDE/3% PTw for imaging following (Asadulina et al. 2012). Phalloidin-stained
8
larvae were mounted in 1% DABCO/glycerol instead, as TDE was found to quickly
9
disrupt phalloidin binding to F-actin. Confocal imaging of stained larvae was performed
10
using a Leica SPE and a Leica SP8 microscope. Stacks were visualized and processed
11
with ImageJ 1.49v. 3D renderings were performed with Imaris 8.1. Bright field Nomarski
12
microscopy was performed on a Zeiss M1 microscope. Z-projections of Nomarski stacks
13
were performed using Helicon Focus 6.7.1. 14
15
Transmission electron microscopy
16
TEM was performed as previously published (Lauri et al. 2014).
17
18
Pharmacological treatments
19
Brefeldin A was purchased from Sigma Aldrich (B7561) and dissolved in DMSO to a
20
final concentration of 5 mg/mL. Animals were treated with 50 µg/mL Brefeldin A in 6-
21
well plates filled with 5 mL filtered natural sea water (FNSW). Controls were treated
22
with 1% DMSO (which is compatible with Platynereis development and survival without
23
noticeable effect). Other neurotransmission inhibitors had no effect on Platynereis (as
21
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
they elicited no impairment of locomotion): tetanus toxic (Sigma Aldrich T3194; 100
2
µg/mL stock in distilled water) up to 5 µg/mL; TTX (Latoxan, L8503; 1 mM stock) up to
3
10 µM; Myobloc (rimabotulinum toxin B; Solstice Neurosciences) up to 1%; saxitoxin 2
4
HCl (Sigma Aldrich NRCCRMSTXF) up to 1%; and neosaxitoxin HCl (Sigma Aldrich
5
NRCCRMNEOC) up to 1%. (±)-2-Octanol was purchased from Sigma Aldrich and
6
diluted to a final concentration of 2.5 mM (2 µL in 5 mL FNSW). (±)-2-Octanol
7
treatment inhibited both locomotion and gut peristalsis, in line with the importance of gap
8
junctions in motor neural circuits (Kawano et al. 2011; Kiehn & Tresch 2002).
9
10
11
Live imaging of contractions
12
13
Animals were mounted in 3% low melting point agarose in FNSW (2-
14
Hydroxyethylagarose, Sigma Aldrich A9414) between a slide and a cover slip (using 5
15
layers of adhesive tape for spacing) and imaged with a Leica SP8 confocal microscope.
16
Fluorescent labeling of musculature was achieved either by microinjection of mRNAs
17
encoding GCaMP6s, LifeAct-EGFP or H2B-RFP, or by incubation in 3 µM 0.1% FM-
18
464FX (ThermoFisher Scientific, F34653). Contraction speed was calculated as (l2-
19
l1)/(l1*t), where l1 is the initial length, l2 the length after contraction, and t the duration
20
of the contraction. Kymographs and wave speed quantifications were performed with the
21
ImageJ Kymograph plugin: http://www.embl.de/eamnet/html/kymograph.html
22
23
22
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Ancestral state reconstruction
2
Ancestral state reconstructions were performed with Mesquite 3.04 using the Maximum
3
Likelihood and Parsimony methods.
4
5
Cloning
6
The following primers were used for cloning Platynereis genes using a mixed stages
7
Platynereis cDNA library (obtained from 1, 2, 3, 5, 6, 10, and 14-days old larvae) and
8
either the HotStart Taq Polymerase from Qiagen or the Phusion polymerase from New
9
England BioLabs (for GC-rich primers):
10
Gene name
Forward primer
Reverse primer
foxF
CCCAGTGTCTGCATCCTTGT
CATGGGCATTGAAGGGGAGT
zasp
CATACCAGCCATCCCGTCC
AAATCAGCGAACTCCAGCGT
troponin T
TTCTGCAGGGCGCAAAGTCA
CGCTGCTGTTCCTTGAAGCG
SM-MRLC
TGGTGTTTGCAGGGCGGTCA
GGTCCATACCGTTACGGAAGCTTTT
calponin
ACGTGCGGTTTACGATTGGA
GCTGGCTCCTTGGTTTGTTC
transgelin1
GCTGCCAAGGGAGCTGACGC ACAAAGAGCTTGTACCACCTCACCC
myocardin
GACACCAGTCCGAAGCTTGA
CGTGGTAGTAGTCGTGGTCG
11
12
13
The following genes were retrieved from an EST plasmid stock: SM-MHC (as two
14
independent clones that gave identical expression patterns) and ST-MRLC.
23
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Other genes were previously published: NK3 (Saudemont et al. 2008), actin and ST-MHC
2
(under the name mhc1-4) (Lauri et al. 2014) and GATA456 (Gillis et al. 2007).
3
Acknowledgements
4
5
We thank Kaia Achim for the hnf4 plasmid, Pedro Machado (Electron Microscopy Core
6
Facility, EMBL Heidelberg) for embedding and sectioning samples for TEM, and the
7
Arendt lab for feedback on the project. The work was supported by the European
8
Research Council “Brain Evo-Devo” grant (TB, PB and DA), European Union’s Seventh
9
Framework Programme project “Evolution of gene regulatory networks in animal
10
development (EVONET)” [215781-2] (AL), the Zoonet EU-Marie Curie early training
11
network [005624] (AF), the European Molecular Biology Laboratory (AL, AHLF, and
12
PRHS) and the EMBL International Ph.D. Programme (TB, AHLF, PRHS, AL).
13
14
15
24
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
References
Alberts, B. et al., 2014. Molecular Biology of the Cell 6 edition., New York, NY: Garland Science.
Anderson, W.A. & Ellis, R.A., 1967. A comparative electron microscope study of visceral muscle fibers in
Cambarus, Drosophila and Lumbricus. Zeitschrift für Zellforschung und Mikroskopische
Anatomie, 79(4), pp.581–591.
Andrikou, C. et al., 2013. Myogenesis in the sea urchin embryo: the molecular fingerprint of the myoblast
precursors. EvoDevo, 4(1), p.33.
Arendt, D., 2003. Evolution of eyes and photoreceptor cell types. International Journal of Developmental
Biology, 47(7/8), pp.563–572.
Arendt, D. et al., 2016. Evolution of sister cell types by individuation. Nature Reviews Genetics, submitted.
Arendt, D., 2008. The evolution of cell types in animals: emerging principles from molecular studies.
Nature Reviews. Genetics, 9(11), pp.868–882.
Asadulina, A. et al., 2012. Whole-body gene expression pattern registration in Platynereis larvae. EvoDevo,
3, p.27.
Au, Y. et al., 2004. Solution Structure of ZASP PDZ Domain: Implications for Sarcomere Ultrastructure
and Enigma Family Redundancy. Structure, 12(4), pp.611–622.
Ayme-Southgate, A. et al., 1989. Characterization of the gene for mp20: a Drosophila muscle protein that is
not found in asynchronous oscillatory flight muscle. The Journal of cell biology, 108(2), pp.521–
531.
Azpiazu, N. & Frasch, M., 1993. tinman and bagpipe: two homeo box genes that determine cell fates in the
dorsal mesoderm of Drosophila. Genes & Development, 7(7b), pp.1325–1340.
Bárány, M., 1967. ATPase activity of myosin correlated with speed of muscle shortening. The Journal of
general physiology, 50(6), pp.197–218.
Beall, C.J. & Fyrberg, E., 1991. Muscle abnormalities in Drosophila melanogaster heldup mutants are
caused by missing or aberrant troponin-I isoforms. The Journal of Cell Biology, 114(5), pp.941–
951.
Black, B.L. & Olson, E.N., 1998. Transcriptional control of muscle development by myocyte enhancer
factor-2 (MEF2) proteins. Annual Review of Cell and Developmental Biology, 14, pp.167–196.
Blais, A. et al., 2005. An initial blueprint for myogenic differentiation. Genes & Development, 19(5),
pp.553–569.
Bohrmann, J. & Haas-Assenbaum, A., 1993. Gap junctions in ovarian follicles of Drosophila melanogaster:
inhibition and promotion of dye-coupling between oocyte and follicle cells. Cell and Tissue
Research, 273(1), pp.163–173.
Brunet, T. & Arendt, D., 2016. From damage response to action potentials: early evolution of neural and
contractile modules in stem eukaryotes. Philosophical Transactions of the Royal Society of
London. Series B, Biological Sciences, 371(1685), p.20150043.
Bulbring, E. & Crema, A., 1959. The release of 5-hydroxytryptamine in relation to pressure exerted on the
intestinal mucosa. The Journal of Physiology, 146(1), pp.18–28.
Burke, R.D., 1981. Structure of the digestive tract of the pluteus larva of Dendraster excentricus
(Echinodermata: Echinoida). Zoomorphology, 98(3), pp.209–225.
Burr, A. & Gans, C., 1998. Mechanical Significance of Obliquely Striated Architecture in Nematode
Muscle. The Biological Bulletin, 194(1), pp.1–6.
Cannon, J.T. et al., 2016. Xenacoelomorpha is the sister group to Nephrozoa. Nature, 530(7588), pp.89–93.
Carnevali, M.D.C. & Ferraguti, M., 1979. Structure and ultrastructure of muscles in the priapulid
Halicryptus spinulosus: functional and phylogenetic remarks. Journal of the Marine Biological
Association of the United Kingdom, 59(3), pp.737–748.
Carson, J.A., Schwartz, R.J. & Booth, F.W., 1996. SRF and TEF-1 control of chicken skeletal alpha-actin
gene during slow-muscle hypertrophy. American Journal of Physiology - Cell Physiology, 270(6),
pp.C1624–C1633.
Chiodin, M. et al., 2011. Molecular architecture of muscles in an acoel and its evolutionary implications.
Journal of Experimental Zoology Part B: Molecular and Developmental Evolution, 316B(6),
pp.427–439.
Corsi, A.K. et al., 2000. Caenorhabditis elegans twist plays an essential role in non-striated muscle
development. Development, 127(10), pp.2041–2051.
25
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
Creemers, E.E. et al., 2006. Coactivation of MEF2 by the SAP Domain Proteins Myocardin and MASTR.
Molecular Cell, 23(1), pp.83–96.
Dayraud, C. et al., 2012. Independent specialisation of myosin II paralogues in muscle vs. non-muscle
functions during early animal evolution: a ctenophore perspective. BMC evolutionary biology,
12(1), p.107.
Denes, A.S. et al., 2007. Molecular architecture of annelid nerve cord supports common origin of nervous
system centralization in bilateria. Cell, 129(2), pp.277–288.
Deutsch, J., 2005. Hox and wings. BioEssays: News and Reviews in Molecular, Cellular and
Developmental Biology, 27(7), pp.673–675.
Durocher, D. et al., 1997. The cardiac transcription factors Nkx2-5 and GATA-4 are mutual cofactors. The
EMBO journal, 16(18), pp.5687–5696.
Duvert, M. & Salat, C., 1995. Ultrastructural Studies of the Visceral Muscles of Chaetognaths. Acta
Zoologica, 76(1), pp.75–87.
Endo, T. & Obinata, T., 1981. Troponin and Its Components from Ascidian Smooth Muscle. Journal of
Biochemistry, 89(5), pp.1599–1608.
Faccioni-Heuser, M.C. et al., 1999. The pedal muscle of the land snail Megalobulimus oblongus
(Gastropoda, Pulmonata): an ultrastructure approach. Acta Zoologica, 80(4), pp.325–337.
Faussone‐Pellegrini, M.-S. & Thuneberg, L., 1999. Guide to the identification of interstitial cells of Cajal.
Microscopy research and technique, 47(4), pp.248–266.
Feral, J.-P. & Massin, C., 1982. Digestive systems: Holothuroidea. Echinoderm nutrition, pp.191–212.
Finkbeiner, S., 1992. Calcium waves in astrocytes-filling in the gaps. Neuron, 8(6), pp.1101–1108.
Fischer, A.H., Henrich, T. & Arendt, D., 2010. The normal development of Platynereis dumerilii
(Nereididae, Annelida). Frontiers in zoology, 7(1), p.1.
Force, A. et al., 1999. Preservation of Duplicate Genes by Complementary, Degenerative Mutations.
Genetics, 151(4), pp.1531–1545.
Francés, R. et al., 2006. B-1 cells express transgelin 2: unexpected lymphocyte expression of a smooth
muscle protein identified by proteomic analysis of peritoneal B-1 cells. Molecular immunology,
43(13), pp.2124–2129.
Fyrberg, C. et al., 1994. Drosophila melanogaster genes encoding three troponin-C isoforms and a
calmodulin-related protein. Biochemical Genetics, 32(3–4), pp.119–135.
Fyrberg, E. et al., 1990. Drosophila melanogaster troponin-T mutations engender three distinct syndromes
of myofibrillar abnormalities. Journal of Molecular Biology, 216(3), pp.657–675.
Genikhovich, G. & Technau, U., 2011. Complex functions of Mef2 splice variants in the differentiation of
endoderm and of a neuronal cell type in a sea anemone. Development, 138(22), pp.4911–4919.
Gho, M., 1994. Voltage-clamp analysis of gap junctions between embryonic muscles in Drosophila. The
Journal of Physiology, 481(Pt 2), pp.371–383.
Gillis, W.J., Bowerman, B. & Schneider, S.Q., 2007. Ectoderm‐and endomesoderm‐specific GATA
transcription factors in the marine annelid Platynereis dumerilli. Evolution & development, 9(1),
pp.39–50.
Gimona, M. et al., 2003. Calponin repeats regulate actin filament stability and formation of podosomes in
smooth muscle cells. Molecular biology of the cell, 14(6), pp.2482–2491.
Goldstein, M.A. & Burdette, W.J., 1971. Striated visceral muscle of Drosophila melanogaster. Journal of
Morphology, 134(3), pp.315–334.
Goodson, H.V. & Spudich, J.A., 1993. Molecular evolution of the myosin family: relationships derived
from comparisons of amino acid sequences. Proceedings of the National Academy of Sciences of
the United States of America, 90(2), pp.659–663.
Gopalakrishnan, S. et al., 2015. A cranial mesoderm origin for esophagus striated muscles. Developmental
cell, 34(6), pp.694–704.
Han, Z. et al., 2004. A myocardin-related transcription factor regulates activity of serum response factor in
Drosophila. Proceedings of the National Academy of Sciences of the United States of America,
101(34), pp.12567–12572.
Hirakow, R., 1985. The vertebrate heart in relation to the protochordates. Fortschritt der Zoologie, 30,
pp.367–369.
Hobert, O., 2016. Chapter Twenty-Five - Terminal Selectors of Neuronal Identity. In P. M. Wassarman, ed.
Current Topics in Developmental Biology. Essays on Developmental Biology, Part A. Academic
26
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
Press, pp. 455–475. Available at:
http://www.sciencedirect.com/science/article/pii/S0070215315002148 [Accessed June 23, 2016].
Hoggatt, A.M. et al., 2013. The Transcription Factor Foxf1 Binds to Serum Response Factor and
Myocardin to Regulate Gene Transcription in Visceral Smooth Muscle Cells. Journal of
Biological Chemistry, 288(40), pp.28477–28487.
Hooper, S.L., Hobbs, K.H. & Thuma, J.B., 2008. Invertebrate muscles: thin and thick filament structure;
molecular basis of contraction and its regulation, catch and asynchronous muscle. Progress in
neurobiology, 86(2), pp.72–127.
Hooper, S.L. & Thuma, J.B., 2005. Invertebrate muscles: muscle specific genes and proteins. Physiological
reviews, 85(3), pp.1001–1060.
Jakobsen, J.S. et al., 2007a. Temporal ChIP-on-chip reveals Biniou as a universal regulator of the visceral
muscle transcriptional network. Genes & Development, 21(19), pp.2448–2460.
Jakobsen, J.S. et al., 2007b. Temporal ChIP-on-chip reveals Biniou as a universal regulator of the visceral
muscle transcriptional network. Genes & Development, 21(19), pp.2448–2460.
Jensen, H., 1974. Ultrastructural studies of the hearts in Arenicola marina L. (Annelida: Polychaeta). Cell
and Tissue Research, 156(1), pp.127–144.
Jensen, H. & Myklebust, R., 1975. Ultrastructure of muscle cells in Siboglinum fiordicum (Pogonophora).
Cell and tissue research, 163(2), pp.185–197.
Kamm, K.E. & Stull, J.T., 1985. The function of myosin and myosin light chain kinase phosphorylation in
smooth muscle. Annual review of pharmacology and toxicology, 25(1), pp.593–620.
Katzemich, A. et al., 2011. Muscle type-specific expression of Zasp52 isoforms in Drosophila. Gene
Expression Patterns, 11(8), pp.484–490.
Kawaguti, S. & Ikemoto, N., 1965. Electron microscopy on the longitudinal muscle of the sea-cucumber
(Stichopus japonicus). S. Ebashi et al., Edits. Molecular biology of muscular contraction. Tokyo:
Igaku Shoin Ltd. XII, pp.124–131.
Kawano, T. et al., 2011. An imbalancing act: gap junctions reduce the backward motor circuit activity to
bias C. elegans for forward locomotion. Neuron, 72(4), pp.572–586.
Kiehn, O. & Tresch, M.C., 2002. Gap junctions and motor behavior. Trends in Neurosciences, 25(2),
pp.108–115.
Kierszenbaum, A.L. & Tres, L., 2015. Histology and Cell Biology: An Introduction to Pathology, 4e 4
edition., Philadelphia, PA: Saunders.
Kobayashi, C. et al., 1998. Identification of Two Distinct Muscles in the Planarian Dugesia japonica by
their Expression of Myosin Heavy Chain Genes. Zoological Science, 15(6), pp.861–869.
Labeit, S. & Kolmerer, B., 1995. Titins: Giant Proteins in Charge of Muscle Ultrastructure and Elasticity.
Science, 270(5234), pp.293–296.
Lan, X. & Pritchard, J.K., 2016. Coregulation of tandem duplicate genes slows evolution of
subfunctionalization in mammals. Science, 352(6288), pp.1009–1013.
Lauri, A. et al., 2014. Development of the annelid axochord: insights into notochord evolution. Science
(New York, N.Y.), 345(6202), pp.1365–1368.
Lee, H.-H., Zaffran, S. & Frasch, M., 2006. Development of the larval visceral musculature. In Muscle
Development in Drosophila. Springer, pp. 62–78.
Lee, Y. et al., 1998. The Cardiac Tissue-Restricted Homeobox Protein Csx/Nkx2.5 Physically Associates
with the Zinc Finger Protein GATA4 and Cooperatively Activates Atrial Natriuretic Factor Gene
Expression. Molecular and Cellular Biology, 18(6), pp.3120–3129.
Leininger, S. et al., 2014. Developmental gene expression provides clues to relationships between sponge
and eumetazoan body plans. Nature Communications, 5, p.3905.
Li, L. et al., 1996. SM22α, a Marker of Adult Smooth Muscle, Is Expressed in Multiple Myogenic Lineages
During Embryogenesis. Circulation Research, 78(2), pp.188–195.
Liang, C. et al., 2015. The statistical geometry of transcriptome divergence in cell-type evolution and
cancer. Nature Communications, 6, p.6066.
Mackie, G.O., 1970. Neuroid conduction and the evolution of conducting tissues. The Quarterly Review of
Biology, 45(4), pp.319–332.
Malo, M. et al., 2000. Effect of brefeldin A on acetylcholine release from glioma C6BU-1 cells.
Neuropharmacology, 39(11), pp.2214–2221.
Marieb, E.N. & Hoehn, K.N., 2015. Human Anatomy & Physiology 10 edition., Boston: Pearson.
27
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
Marín, M.-C., Rodríguez, J.-R. & Ferrús, A., 2004. Transcription of Drosophila Troponin I Gene Is
Regulated by Two Conserved, Functionally Identical, Synergistic Elements. Molecular Biology of
the Cell, 15(3), pp.1185–1196.
Martindale, M.Q., Pang, K. & Finnerty, J.R., 2004. Investigating the origins of
triploblasty:mesodermal’gene expression in a diploblastic animal, the sea anemone Nematostella
vectensis (phylum, Cnidaria; class, Anthozoa). Development, 131(10), pp.2463–2474.
Martín-Durán, J.M. & Hejnol, A., 2015. The study of Priapulus caudatus reveals conserved molecular
patterning underlying different gut morphogenesis in the Ecdysozoa. BMC biology, 13, p.29.
Martynova, M.G., 1995. Possible Cellular Mechanisms of Heart Muscle Growth in Invertebratea. Annals of
the New York Academy of Sciences, 752(1), pp.149–157.
Martynova, M.G., 2004. Proliferation and Differentiation Processes in the Heart Muscle Elements in
Different Phylogenetic Groups. In B.-I. R. of Cytology, ed. Academic Press, pp. 215–250.
Available at: http://www.sciencedirect.com/science/article/pii/S0074769604350059 [Accessed
June 17, 2016].
McKeown, C.R., Han, H.-F. & Beckerle, M.C., 2006. Molecular characterization of the Caenorhabditis
elegans ALP/Enigma gene alp-1. Developmental Dynamics, 235(2), pp.530–538.
Meadows, S.M. et al., 2008. The myocardin-related transcription factor, MASTR, cooperates with MyoD to
activate skeletal muscle gene expression. Proceedings of the National Academy of Sciences,
105(5), pp.1545–1550.
Meedel, T.H. & Hastings, K.E., 1993. Striated muscle-type tropomyosin in a chordate smooth muscle,
ascidian body-wall muscle. Journal of Biological Chemistry, 268(9), pp.6755–6764.
Mill, P.J. & Knapp, M.F., 1970. The Fine Structure of Obliquely Striated Body Wall Muscles in the
Earthworm, Lumbricus Terrestris LINN. Journal of Cell Science, 7(1), pp.233–261.
Misumi, Y. et al., 1986. Novel blockade by brefeldin A of intracellular transport of secretory proteins in
cultured rat hepatocytes. Journal of Biological Chemistry, 261(24), pp.11398–11403.
Molkentin, J.D. et al., 1995. Cooperative activation of muscle gene expression by MEF2 and myogenic
bHLH proteins. Cell, 83(7), pp.1125–1136.
Morin, S. et al., 2000. GATA-dependent recruitment of MEF2 proteins to target promoters. The EMBO
journal, 19(9), pp.2046–2055.
Musser, J.M. & Wagner, G.P., 2015. Character trees from transcriptome data: Origin and individuation of
morphological characters and the so-called “species signal.” Journal of Experimental Zoology.
Part B, Molecular and Developmental Evolution, 324(7), pp.588–604.
Myers, C.D. et al., 1996. Developmental genetic analysis of troponin T mutations in striated and
nonstriated muscle cells of Caenorhabditis elegans. The Journal of Cell Biology, 132(6), pp.1061–
1077.
Nagai, T. & Brown, B.E., 1969. Insect visceral muscle. Electrical potentials and contraction in fibres of the
cockroach proctodeum. Journal of Insect Physiology, 15(11), pp.2151–2167.
Nishida, W. et al., 2002. A Triad of Serum Response Factor and the GATA and NK Families Governs the
Transcription of Smooth and Cardiac Muscle Genes. Journal of Biological Chemistry, 277(9),
pp.7308–7317.
Nyitray, L. et al., 1994. Scallop striated and smooth muscle myosin heavy-chain isoforms are produced by
alternative RNA splicing from a single gene. Proceedings of the National Academy of Sciences,
91(26), pp.12686–12690.
Nylund, A. et al., 1988. Heart ultrastructure in four species of Onychophora (Peripatopsidae and
Peripatidae) and phylogenetic implications. Zool Beitr, 32, pp.17–30.
Obinata, T., Ooi, A. & Takano-Ohmuro, H., 1983. Myosin and actin from ascidian smooth muscle and their
interaction. Comparative Biochemistry and Physiology Part B: Comparative Biochemistry, 76(3),
pp.437–442.
Ojima, T. & Nishita, K., 1986. Isolation of Troponins from Striated and Smooth Adductor Muscles of
Akazara Scallop. Journal of Biochemistry, 100(3), pp.821–824.
Økland, S., 1980. The heart ultrastructure of Lepidopleurus asellus (Spengler) and Tonicella marmorea
(Fabricius)(Mollusca: Polyplacophora). Zoomorphology, 96(1–2), pp.1–19.
OOta, S. & Saitou, N., 1999. Phylogenetic relationship of muscle tissues deduced from superimposition of
gene trees. Molecular Biology and Evolution, 16(6), pp.856–867.
Paniagua, R. et al., 1996. Ultrastructure of invertebrate muscle cell types. Histology and Histopathology,
11(1), pp.181–201.
28
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
Phiel, C.J. et al., 2001. Differential Binding of an SRF/NK-2/MEF2 Transcription Factor Complex in
Normal Versus Neoplastic Smooth Muscle Tissues. Journal of Biological Chemistry, 276(37),
pp.34637–34650.
Prosser, C.L., Nystrom, R.A. & Nagai, T., 1965. Electrical and mechanical activity in intestinal muscles of
several invertebrate animals. Comparative biochemistry and physiology, 14(1), pp.53–70.
Raible, F. et al., 2005. Vertebrate-type intron-rich genes in the marine annelid Platynereis dumerilii.
Science, 310(5752), pp.1325–1326.
Reynolds, P.D., Morse, M.P. & Norenburg, J., 1993. Ultrastructure of the Heart and Pericardium of an
Aplacophoran Mollusc (Neomeniomorpha): Evidence for Ultrafiltration of Blood. Proceedings of
the Royal Society of London B: Biological Sciences, 254(1340), pp.147–152.
Rieger, R. et al., 1991. Organization and differentiation of the body-wall musculature in Macrostomum
(Turbellaria, Macrostomidae). Hydrobiologia, 227(1), pp.119–129.
Roach, D.K., 1968. Rhythmic muscular activity in the alimentary tract of Arion ater (L.)(Gastropoda:
Pulmonata). Comparative biochemistry and physiology, 24(3), pp.865–878.
Rogers, D.C., 1969. Fine structure of smooth muscle and neuromuscular junctions in the foot of Helix
aspersa. Zeitschrift für Zellforschung und Mikroskopische Anatomie, 99(3), pp.315–335.
Rosenbluth, J., 1972. Obliquely striated muscle. In The structure and function of muscle. New York and
London: Academic Press, pp. 389–420.
Royuela, M. et al., 2000. Characterization of several invertebrate muscle cell types: a comparison with
vertebrate muscles. Microscopy Research and Technique, 48(2), pp.107–115.
Royuela, M. et al., 1997. Immunocytochemical electron microscopic study and Western blot analysis of
caldesmon and calponin in striated muscle of the fruit fly Drosophila melanogaster and in several
muscle cell types of the earthworm Eisenia foetida. European journal of cell biology, 72(1),
pp.90–94.
Sanders, K.M., Koh, S.D. & Ward, S.M., 2006. Interstitial cells of Cajal as pacemakers in the
gastrointestinal tract. Annu. Rev. Physiol., 68, pp.307–343.
Saudemont, A. et al., 2008. Complementary striped expression patterns of NK homeobox genes during
segment formation in the annelid Platynereis. Developmental biology, 317(2), pp.430–443.
Schlesinger, J. et al., 2011. The Cardiac Transcription Network Modulated by Gata4, Mef2a, Nkx2.5, Srf,
Histone Modifications, and MicroRNAs. PLOS Genet, 7(2), p.e1001313.
Schmidt-Rhaesa, A., 2007. The Evolution of Organ Systems 1 edition., Oxford ; New York: Oxford
University Press.
Sebé-Pedrós, A. et al., 2014. Evolution and classification of myosins, a paneukaryotic whole-genome
approach. Genome biology and evolution, 6(2), pp.290–305.
Shaw, K., 1974. The fine structure of muscle cells and their attachments in the tardigrade Macrobiotus
hufelandi. Tissue and Cell, 6(3), pp.431–445.
Shi, X. & Garry, D.J., 2006. Muscle stem cells in development, regeneration, and disease. Genes &
Development, 20(13), pp.1692–1708.
Shimeld, S.M. et al., 2010. Clustered Fox genes in lophotrochozoans and the evolution of the bilaterian Fox
gene cluster. Developmental biology, 340(2), pp.234–248.
Silverthorn, D.U., 2015. Human Physiology: An Integrated Approach 7 edition., San Francisco: Pearson.
Spies, R.B., 1973. The blood system of the flabelligerid polychaete Flabelliderma commensalis (Moore).
Journal of morphology, 139(4), pp.465–490.
Steinmetz, P.R.H. et al., 2012. Independent evolution of striated muscles in cnidarians and bilaterians.
Nature, 487(7406), pp.231–234.
Sulbarán, G. et al., 2015. An invertebrate smooth muscle with striated muscle myosin filaments.
Proceedings of the National Academy of Sciences, 112(42), pp.E5660–E5668.
Susic-Jung, L. et al., 2012. Multinucleated smooth muscles and mononucleated as well as multinucleated
striated muscles develop during establishment of the male reproductive organs of Drosophila
melanogaster. Developmental Biology, 370(1), pp.86–97.
Sweeney, H.L., Bowman, B.F. & Stull, J.T., 1993. Myosin light chain phosphorylation in vertebrate
striated muscle: regulation and function. American Journal of Physiology-Cell Physiology, 264(5),
pp.C1085–C1095.
Telford, M.J. & Copley, R.R., 2016. Zoology: War of the Worms. Current Biology, 26(8), pp.R335–R337.
Thor, S. & Thomas, J.B., 2002. Motor neuron specification in worms, flies and mice: conserved and “lost”
mechanisms. Current Opinion in Genetics & Development, 12(5), pp.558–564.
29
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Tjønneland, A., Økland, S. & Nylund, A., 1987. Evolutionary aspects of the arthropod heart. Zoologica
Scripta, 16(2), pp.167–175.
Wagner, G.P., 2014. Homology, Genes, and Evolutionary Innovation, Princeton ; Oxford: Princeton
University Press.
Wales, S. et al., 2014. Global MEF2 target gene analysis in cardiac and skeletal muscle reveals novel
regulation of DUSP6 by p38MAPK-MEF2 signaling. Nucleic Acids Research, p.gku813.
Wang, D.-Z. & Olson, E.N., 2004. Control of smooth muscle development by the myocardin family of
transcriptional coactivators. Current Opinion in Genetics & Development, 14(5), pp.558–566.
Wang, Z. et al., 2003. Myocardin is a master regulator of smooth muscle gene expression. Proceedings of
the National Academy of Sciences, 100(12), pp.7129–7134.
White, J., 1988. 4 The Anatomy. The nematode Caenorhabditis elegans - Cold Spring Harbor Monograph
Archive, 17, pp.81–122.
Witchley, J.N. et al., 2013. Muscle Cells Provide Instructions for Planarian Regeneration. Cell Reports,
4(4), pp.633–641.
Wood, J.D., 1969. Electrophysiological and pharmacological properties of the stomach of the squid Loligo
pealii (Lesueur). Comparative biochemistry and physiology, 30(5), pp.813–824.
Wu, K.S., 1939. On the physiology and pharmacology of the earthworm gut. J. exp. Biol., 16, pp.184–197.
Zaffran, S. et al., 2001. biniou (FoxF), a central component in a regulatory network controlling visceral
mesoderm development and midgut morphogenesis in Drosophila. Genes & Development, 15(21),
pp.2900–2915.
Zhang, Y. et al., 2000. Drosophila D-titin is required for myoblast fusion and skeletal muscle striation.
Journal of cell science, 113(17), pp.3103–3115.
Zhou, Q. et al., 2001. Ablation of Cypher, a PDZ-LIM domain Z-line protein, causes a severe form of
congenital myopathy. The Journal of Cell Biology, 155(4), pp.605–612.
30
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
A
Cross-striation, solid Z-discs
B
Skeletal myocyte
core regulatory complex
MyoD E12
MASTR
SRF/MEF2
Cross-striation, discontinuous Z-elements
C
Striated effectors
skeletal actin
skeletal troponin C
skeletal troponin I
ZASP/LDB3
Transcription factor
families
bHLH
SRF or
MEF2
(MADS
domain)
Myocardin
(SAP
domain)
Smooth myocyte
core regulatory complex
Fox
FoxF1
Oblique striation, discontinuous Z-elements
Nkx3.2
Myocardin
SRF/MEF2
GATA6
Smooth ultrastructure
D
Brunet et al. Figure 1
GATA
NK
direct
enhancer
binding
demonstrated
regulation,
direct binding
untested
Cardiac myocyte
core regulatory complex
FoxC1
Thin myofilament: actin
Thick myofilament: myosin
Dense bodies
Smooth effectors
smooth actin
smooth MLCK
SM22alpha
telokin
caldesmon
Nkx2.5
?
Myocardin
SRF/MEF2
GATA4
?
?
Cardiac effectors
cardiac actin
cardiac MHC
cardiac troponin T
?
demonstrated
co-regulation,
direct binding
untested
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
2
Figure 1. Ultrastructure and core regulatory complexes of myocyte types.
3
(A) Schematic smooth and striated ultrastructures. Electron-dense granules called “dense
4
bodies” separate adjacent myofibrils. Dense bodies are scattered in smooth muscles, but
5
aligned in striated muscles to form Z lines. (B-D) Core regulatory complexes (CRC) of
6
transcription factors for the differentiation of different types of myocytes in vertebrates.
7
Complexes composition from (Creemers et al. 2006; Meadows et al. 2008; Molkentin et
8
al. 1995) for skeletal myocytes, (Hoggatt et al. 2013; Nishida et al. 2002; Phiel et al.
9
2001) for smooth myocytes, and (Durocher et al. 1997; Lee et al. 1998) for
10
cardiomyocytes. Target genes from (Blais et al. 2005) for skeletal myocytes, (Nishida et
11
al. 2002) from smooth myocytes, and (Schlesinger et al. 2011) for cardiomyocytes.
12
31
posted online Jul. 20, 2016;
doi: http://dx.doi.org/10.1101/064881
copyright holder for this preprint (which was not
dpf
11 dpf
1.5 months . TheB
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND
4.0
International license.
AbioRxiv preprint 7first
Somatic
Visceral
muscles
Ventral half
foregut
A
D
muscles
striated
ventral longitudinal
muscles
smooth/striated
foregut musculature
C
midgut
G H
hindgut
ventral oblique
muscles
V
I
3
E
4
J K L
posterior
foregut
muscles
midgut
musculature
P
axochord
D
Dorsal half
M
parapodial
muscles
F
hindgut
musculature
Somatic muscles
Visceral muscles
E
C
G
mitochondrion
foregut
lumen
Z-lines
F
Zlines
Zlines
D
H
I
Z-lines
dense
bodies
foregut
muscles
Zlines
J
K
coelom
dense bodies
longitudinal muscles
circular
muscles
circular muscles
L
gut epithelium
Brunet et al. Figure 2
longitudinal muscles
dense bodies
M
dense
bodies
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 2. Development and ultrastrcture of visceral and somatic musculature in
2
Platynereis larvae and juveniles. (A) Development of visceral musculature. All panels
3
are 3D renderings of rhodamine-phalloidin staining imaged by confocal microscopy.
4
Visceral muscles have been manually colored green and somatic muscle red. Scale bar:
5
50 µm. (B) Schematic of the musculature of a late nectochaete (6 dpf) larva. Body outline
6
modified from (Fischer et al. 2010). Ventral view, anterior is up. (C-M) Electron
7
micrographs of the main muscle groups depicted in B. Each muscle group is shown
8
sectioned parallel to its long axis, so in the plane of its myofilaments. Scale bar: 2 µm.
9
(H) is a close-up of the region in the yellow box in G. (J) shows the dorsal midgut in
10
cross-section, dorsal side up.
11
32
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
72 hpf
Mef2
Myocardin
KLF5
Actin
GATA456
FoxF
NK3
Calponin*
SM-MRLC
Smooth molecular profile
SM-MHC
MyoD
Titin
Zasp/LBD3
Troponin T
*
Troponin I
B
Striated molecular profile
ST-MRLC
*
SM-MHC
ST-MHC
A
ST-MHC
Annelid longitudinal
striated muscles
Annelid striated
foregut muscles
*
D
*
6 dpf
C
Annelid
oblique/parapodial
striated muscles
Annelid gut
smooth muscles
Vertebrate striated
trunk muscles
Vertebrate striated
esophageal muscles
Vertebrate gut
smooth muscles
Acto- Calcium Striation Transc. Actomyosin response
factors myosin
Expressed
Brunet et al. Figure 3
Not expressed
Transcription Acto- Transcription
factors
myosin factors
Expressed in a subset
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 3. Expression of smooth and striated muscle markers in Platynereis larvae.
2
Animals have been stained by WMISH and observed in bright field Nomarski
3
microscopy. Ventral views, anterior side up. Scale bar: 25 µm. (A-D) Expression patterns
4
of the striated marker ST-MHC and the smooth marker SM-MHC. These expression
5
patterns are representative of the entire striated and smooth effector module (see Figure
6
3-supplements 1 and 3). (E) Table summarizing the expression patterns of smooth and
7
striated markers in Platynereis and vertebrate muscles. (*) indicates that Platynereis and
8
vertebrate Calponin are mutually most resembling by domain structure, but not one-to-
9
one orthologs, as independent duplications in both lineages have given rise to more
10
broadly expressed paralogs with a different domain structure (Figure 7—supplement 3).
11
12
33
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
A
c. b.
Lifeact-EGFP
H2B-RFP
B
c. b.
Lifeact-EGFP
H2B-RFP
GCaMP6s
D
E
52 μm
61 μm
t0
F
t0+280ms
Brunet et al. Figure 4
t0+436 ms
myocytes
1.2
H2B-RFP
50 μm
midgut
t0
0.8
B’
t0+280ms 3 dpf
0.4
H2B-RFP
foregut
51
μm
Contraction speed (s-1)
t0
telotroch
p = 8.5*10-8
0.0
20 μm
A’
FM-464FX
t0
telotroch
6 dpf
G
60
μm
longitudinal
muscle
c. b.
ventral
longitudinal
muscles
C
Somatic striated Visceral smooth
muscles
muscles
t0+1.3s
1
1
31
μm
27
μm
2
2
hindgut
50 μm
6 dpf
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 4. Contraction speed quantifications of smooth and striated muscles. (A-
2
B) Snapshots of a time lapse live confocal imaging of a late nectochaete larva expressing
3
fluorescent markers. Ventral view of the 2 posterior-most segments, anterior is up.
4
(C) Snapshots of a time lapse live confocal imaging of a 3 dpf larva expressing
5
GCaMP6s. Dorsal view, anterior is up. (D-E) Two consecutive snapshots on the left
6
dorsal longitudinal muscle of the larva shown in C, showing muscle contraction.
7
(F) Quantification of smooth and striated muscle contraction speeds (see Experimental
8
procedures). (G) Snapshot of a time lapse live confocal imaging of a late nectochaete
9
larva. Ventral view, anterior is up. Optical longitudinal section at the midgut level.
10
11
34
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
A
B
head
pharynx
C Kymograph
line of
interest
longitudinal
fibres
circular
fibres
parapodia
space (x)
Δx
contraction
waves
outline of
the gut
gut
wave propagation
=
speed
7
2
3
4
5
6
n. s.
1
F
Control Brefeldin Atreated
+Brefeldin-A
+2-octanol
Brefeldin Atreated
time (t)
Recurrence of contraction events
n. s.
Number of contractions
per minute
1 2 3 4 5 6
300
200
100
Control
Control
G
FM464-FX
E
p = 5*10-3
0
Wave propagation speed ( m/s)
D
100 μm
Brunet et al. Figure 5
Control Brefeldin Atreated
+Brefeldin-A +2-octanol
2 months
400 μm
Δx
Δt
Δt
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 5. Platynereis gut peristalsis is independent of nervous inputs and dependent
2
on gap junctions. (A) 2 months-old juvenile mounted in 3% low-melting point (LMP)
3
agarose for live imaging. (B) Snapshots of a confocal live time lapse imaging of the
4
animal shown in A. Gut is observed by detecting fluorescence of the vital membrane dye
5
FM-464FX. (C) Kymograph of gut peristalsis along the line of interest in (B).
6
Contraction waves appear as dark stripes. A series of consecutive contraction waves is
7
called a contraction event: here, two contraction waves are visible, which make up one
8
contraction event with a recurrence of 2. (D) Quantification of the propagation speed of
9
peristaltic contraction waves in mock (DMSO)-treated individuals and Brefeldin A-
10
treated individuals (inhibiting neurotransmission). Speed is calculated from kymographs
11
(see Material and Methods). (E,F) Same as in E, but showing respectively the frequency
12
of initiation and the recurrence of contraction events. (G) Representative kymographs of
13
controls, animals treated with Brefeldin A (inhibiting neurotransmission), animals treated
14
with 2-octanol (inhibiting gap junctions), and animals treated with both (N=10 for each
15
condition). 2-octanol entirely abolishes peristaltic waves with or without Brefeldin A.
16
17
35
A outline of the gut B
C
D
serotonin
Brunet et al. Figure 6
ac-tubulin
serotonin
combined
neurites
50 μm ac-tubulin
10 μm
ac-tubulin
ac-tub
phalloidin
serotonin
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 6. The enteric nerve net of Platynereis. (A) Immunostaining for acetylated
2
tubulin, visualizing neurites of the enteric nerve plexus. Z-projection of a confocal stack
3
at the level of the midgut. Anterior side up. (B) Same individual as in A, immunostaining
4
for serotonin (5-HT). Note serotonergic neurites (double arrow), serotonergic neuronal
5
cell bodies (arrow, see D), and serotonergic cell bodies without neurites (arrowhead).
6
(C) Same individual as in A showing both acetylated tubulin and 5-HT immunostainings.
7
Snapshot in the top right corner: same individual, showing both neurites (acetylated
8
tubulin, yellow) and visceral myofibres (rhodamine-phalloidin, red). The acetylated
9
tubulin appears yellow due to fluorescence leaking in the rhodamine channel. (D) 3D
10
rendering of the serotonergic neuron shown by arrow in B.
11
12
36
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
Slow-contracting group NM/SM-MHC+
Cardiomyocytes
Porifera
Cnidaria
Deuterostomia
striation,
fast
module
Somatic striated
myocytes
Gut myocytes
Protostomia
Protostomia
Vertebrate
Sponge
Nematostella
Annelid
Drosophila Drosophila gut
exopinacocyte slow myocytes cardiomyocyte cardiomyocyte cardiomyocyte
striated
myocytes
SM-MHC
loss
Fast-contracting group ST-MHC+
SM-MHC
loss
SM-MHC
loss
striation,
fast module striation,
fast module
Deuterostomia Deuterostomia
Annelid gut
smooth
myocytes
Vertebrate gut
smooth
myocytes
sm-actin
sm-MELC
Vertebrate
striated
myocytes
Annelid
striated
myocytes
Cnidaria
Planarian
Nematostella
somatic
fast
smooth myocytes myocytes
st-actin
st-MELC
actin
MELC
myocardin
Protostomia
striation
loss
Porifera
Jellyfish
Sponge
striated reticuloapopylocyte
myocytes
striation:
convergent
evolution
MASTR
myocardin
connexin
gap junctions
SM-MRLC
GATA456
?
?
ST-MRLC
MRLC
innexin
gap junctions
SM-MHC
striation:
titin, troponin complex
MyoD
ST-MHC
?
?
change in core regulatory complex
apomere
SM-MHC
module co-option
module loss
species split
cell type split
paralog 1
paralog 2
reciprocal loss of expression
(division of labor)
paralog 2
ancestral
gene
MHC
ST-MHC
paralog 1
gene duplication
Brunet et al. Figure 7
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 7. The evolutionary tree of animal contractile cell types. Bilaterian smooth and
2
striated muscles split before the last common protostome/deuterostome ancestor.
3
Bilaterian myocytes are split into two monophyletic cell type clades: an ancestrally SM-
4
MHC+ slow-contracting clade (green) and an ancestrally ST-MHC+ fast-contracting
5
clade (orange). Hypothetical relationships of the bilaterian myocytes to the SM-MHC+
6
and ST-MHC+ contractile cells of non-bilaterians are indicated by dotted lines (Steinmetz
7
et al. 2012). Apomere: derived set of effector genes common to a monophyletic group of
8
cell types (Arendt et al. 2016). Note that ultrastructure only partially reflects evolutionary
9
relationships, as striation can evolve convergently (as in medusozoans), be co-opted (as
10
in insect gut myocytes or in vertebrate and insect cardiomyocytes), be blurred, or be lost
11
(as in planarians). Conversion of smooth to striated myocytes took place by co-option of
12
striation proteins (Titin, Zasp/LDB3) and of the fast contractile module (ST-MHC, ST-
13
MRLC, Troponin complex) in insect cardiomyocytes and gut myocytes, as well as in
14
vertebrate cardiomyocytes.
15
16
37
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
foxA
B
foxA
*
C
bra
*
D
hnf4
*
E
*
foregut
A
bra
6 dpf
6 dpf
6 dpf
6 dpf
Brunet et al. Figure 2 - supplement 1
hindgut
midgut
foxA
bra+foxA
hnf4
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 2—figure supplement 1. Gut patterning in Platynereis 6 dpf larvae. (A-D)
2
WMISH for gut markers. (A) Ventral view, anterior is up. (B-D) Left lateral views,
3
anterior is up, ventral is right. (E) Schematic of gut patterning in Platynereis late
4
nectochaete larvae. Scale bar: 50 µm.
5
6
38
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
B
zasp
C
titin
D
troponin I
E
troponin T
48 hpf
A
ST-MRLC
G
H
I
J
K
L
M
N
O
6 dpf
72 hpf
EF
P
Q
Q
R
l
U
ST-MHC
S
ST-MRLC
T
troponin I
V
troponin T
W
Brunet et al. Figure 3 - supplement 1
titin
zasp
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 3—supplement 1. Expression of striated muscle markers in Platynereis
2
larvae. (A-O) Larvae stained by WMISH and observed in bright field Nomarski
3
microscopy. Ventral views, anterior side up. Scale bar is 20 µm for 48 hpf and 25 µm for
4
the two other stages. (P) Foregut musculature visualized by rhodamine-phalloidin
5
fluorescent staining. Z-projection of confocal planes. Ventral view, anterior side up.
6
Scale bar: 20 µm. (Q) TEM micrograph of a cross-section of the foregut. Foregut muscles
7
are colored green, axochord orange, ventral oblique muscles pink, ventral nerve cord
8
yellow. Inset: zoom on the area in the red dashed box with enhanced contrast to visualize
9
oblique striation. Scale bar: 10 µm. (R-W) WMISH for striated muscle markers
10
expression in the foregut observed in Nomarski bright field microscopy, ventral views,
11
anterior side up. Scale bar: 20 µm.
12
13
39
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
Axochord
titin
zasp
ST-MHC
ST-MRLC
DAPI
**
**
* *
*
* *
*
**
*
** *
*
*
DAPI
*
*
*
*
*
*
*
*
** *
*
*
Oblique muscles
DAPI NBT/BCIP
*
*
*
*
*
*
*
*
*
troponin I
*
*
*
*
*
*
myoD
troponin T
*
*
*
*
**
*
*
*
*
*
*
*
*
*
*
*
* *
*
*
*
*
*
*
*
*
*
*
*
*
**
*
** *
*
*
**
*
*
* *
*
* *
** *
*
* *
*
* *
** *
*
** *
*
* **
*
**
* **
*
**
*
*
*
*
*
*
** * *
*
* * *
*
*
**
*
* **
*
*
*
*
*
*
* *
* **
*
*
*
**
*
*
* *
**
*
*
*
* *
*
*
*
*
*
*
* *
*
*
*
*
*
* *
* *
* **
*
*
*
*
*
**
*
*
*
*
*
*
* * *
*
*
*
**
*
*
*
** * *
*
*
*
*
*
*
*
**
* **
**
*
*
*
*
*
**
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
* *
*
*
*
*
*
*
*
DAPI NBT/BCIP
*
*
*
**
*
*
*
DAPI
*
*
*
Parapodial muscles
*
*
*
*
*
*
*
DAPI NBT/BCIP
*
*
*
*
DAPI
*
*
* *
* *
*
*
*
*
*
*
*
Antennal muscles
DAPI NBT/BCIP
*
**
*
*
*
DAPI
*
*
*
Longitudinal muscles
*
*
*
*
*
*
*
*
*
* *
* *
*
*
DAPI NBT/BCIP
*
** *
*
Brunet et al. Figure 3 - supplement 2
*
*
**
*
*
*
**
*
*
** *
*
*
*
**
*
*
*
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 3—supplement 2. Expression of striated muscle markers in the 6 dpf
2
Platynereis larva. Animals have been stained by WMISH and observed by confocal
3
microscopy (DAPI fluorescence and NBT/BCIP 633 nm reflection). All striated effector
4
genes are expressed in all somatic muscles examined. The transcription factor myoD is
5
detectable in the axochord and in ventral longitudinal muscles, but not in other muscles.
6
7
40
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
SM-MRLC
72 hpf
A
B
GATA456
C
l. d.
E
F
*
D
foxF
*
*
*
*
6 dpf
calponin
*
l. d.
G
*
72 hpf
H
*
forming
NK3 gut muscles
*
M
N
dlm
dlm
dlm
DAPI FM-464FX
phalloidin
heart
O
SM-MHC coloc.
t.f.
l.f.
72 hpf
K foxF
GATA456 calponin
72 hpf
heart
coelom
gut
muscles
SM-MHC
DAPI
heart
intrinsic
muscles
gut
muscles
gut
gut
P
heart
dlm
gut
NK3
actin coloc.
gut
muscles
gut
heart
tube
actin
Q
dorsal longitudinal
muscles
coelom
intrisic
muscles
heart tube
2 months - heart tube
L
SM-MHC
DAPI
t.f.
l.f.
72 hpf
DAPI FM-464FX
phalloidin
J
I
ST-MHC
DAPI
intrisic
muscles
striated musculature ST-MHC+
gut
smooth musculature SM-MHC+
blood
epidermis
coelomic lining
Brunet et al. Figure 3 - supplement 3
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 3—supplement 3. Expression of smooth muscle markers in Platynereis
2
larvae. (A-H) Animals are stained by WMISH and observed by Nomarski bright field
3
microscopy. Ventral views, anterior side up. Yellow arrows: expression in the foregut
4
mesoderm. White dashed lines: outline of the midgut and hindgut (or their anlage at 3
5
dpf). Asterisk: stomodeum. (I) Schematic drawing of a 3 dpf larva (ventral view, anterior
6
is up) showing the forming midgut and hindgut (dotted line) and the forming visceral
7
myofibres in green (compared to double WMISH panels J and K). (J,K) Animals are
8
stained by double WMISH and observed by confocal microscopy (cyanine 3 precipitate
9
fluorescence for actin, NBT/BCIP precipitate reflection for other genes). Ventral views,
10
anterior side up. Abbreviations: t.f.: transverse fibres; l.f.: longitudinal fibres. (L-Q)
11
Molecular profile of the pulsatile dorsal vessel. All panels show 2 months-old juvenile
12
worms. (L,M) Maximal Z-projections of confocal stacks. Dorsal view, anterior is up. (L)
13
Dorsal musculature of a juvenile Platynereis dumerilii individual visualized by
14
phalloidin-rhodamine (green) together with nuclear (DAPI, blue) and membrane (FM-
15
464FX, red) stainings. The heart tube lies on the dorsal side, bordered by the somatic
16
dorsal longitudinal muscles (dlm). (M) Expression of SM-MHC in the heart tube
17
visualized by WMISH. (N) Virtual cross-section of the individual shown in A. Dorsal
18
side up. Note the continuity of the muscular heart tube with gut musculature. (O) Virtual
19
cross-section of the individual shown in B, showing continuous expression of SM-MHC
20
in the heart and the midgut smooth musculature. Note the similarity to the NK4/tinman
21
expression pattern documented in (Saudemont et al. 2008). (P) Virtual cross-section on
22
an individual stained by WMISH for ST-MHC expression. Note the lack of expression in
23
the heart, while expression is detected in intrinsic muscles that cross the internal cavity to
41
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
suspend internal organs, and in dorsal longitudinal muscles (dlm). (Q) Schematic cross-
2
section of a juvenile worm (dorsal side up) showing the shape, connections and molecular
3
profile of the main muscle groups. Scale bar: 30 µm in all panels.
4
5
42
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
B
SM-MRLC
Smooth muscle markers
A
C
calponin
D GATA456
E
foxF
G
ST-MHC
H
J
troponin I
K
SM-MHC
foregut
*
midgut
6 dpf
Striated muscle markers
F
I
ST-MRLC
titin
zasp
troponin T
Brunet et al. Figure 3 - supplement 4
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 3—supplement 4. Molecular profile of midgut muscles in the 6 dpf larva. All
2
panels are Z-projections of confocal planes, ventral views, anterior side up. Blue: DAPI,
3
red: NBT/BCIP precipitate. White dashed line: midgut/hindgut, yellow dashed ellipse:
4
stomodeum. (A-E) Smooth markers expression. White arrows indicate somatic
5
expression of GATA456 in the ventral oblique muscles. (F-K) Striated markers
6
expression; none of them is detected in any gut cell. White arrows: somatic expression.
7
Scale bar: 20 µm in all panels.
8
9
43
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
*
actin
48 hpf
*
A
dlm
24 hpf B
dlm
24 hpf C
D dlmvlm
vlm
48 hpf
E
am
a
72 hpf
F
dlm
a vom
lom
vom
lom
vc
*
vc
G
pm
lom
myocardin
*
actin
mef2
6 dpf H
6 dpf J
I
f.m.
f.m.
6 dpf L
K
f.m.
f.m.
f.m.
midgut
midgut
M Oblique muscles
DAPI
midgut
Longitudinal muscles
DAPI NBT/BCIP
DAPI
DAPI NBT/BCIP
Oblique muscles
DAPI
DAPI NBT/BCIP
*
*
*
*
*
*
*
*
*
*
*
*
vlm
*
*
* **
*
Longitudinal muscles
DAPI
DAPI NBT/BCIP
r.a.
Oblique muscles
DAPI
DAPI NBT/BCIP
ach
*
*
vlm
*
* **
midgut
midgut
midgut
ach
*
72 hpf
*
*
**
*
*
**
*
Brunet et al. Figure 3 - supplement 5
* *
*
* *
*
Longitudinal muscles
DAPI
*
*
*
*
*
*
DAPI NBT/BCIP
*
*
*
*
*
*
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 3—supplement 5. General muscle markers are expressed in both smooth and
2
striated muscles. All panels show gene expression visualized by WMISH. (A-F) actin
3
expression. (A-B) bright field micrographs in Nomarski optics. (A) is an apical view, (B)
4
is a ventral view. Abbreviations: dlm, dorsal longitudinal muscles; vc, ventral
5
mesodermal cells, likely representing future ventral musculature. (C-F) 3D rendering of
6
confocal imaging of NBT/BCIP precipitate. (C,E) ventral views, anterior side up. (D,F)
7
ventrolateral views, anterior side up. Abbreviations are as in Figure 1. (G-M) Z-
8
projections of confocal stacks. Blue is DAPI, red is reflection signal of NBT/BCIP
9
precipitate. (G-L) Ventral views, anterior side up. White dashed line: midgut, yellow
10
ellipse: foregut. White arrows: somatic muscle expression. Abbreviations: f.m.: foregut
11
muscles; r.a.: reflection artifact. (M) Expression in individual somatic muscles. Scale bar:
12
20 µm in all panels.
13
14
44
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
Evolution of somatic striated CRC
Vertebrate trunk
striated myocytes
Platynereis trunk striated
longitudinal myocytes
MyoD E12
MyoD E12
MASTR
Nau
Myocardin
SRF/MEF2
Drosophila trunk
striated myocytes
Da
hlh-1 hnd-1
Striated
effectors
MEF2
MEF2
Caenorhabditis trunk
striated myocytes
Striated
effectors
UNC-120
Striated
effectors
Striated
effectors
MyoD E12
Myocardin
SRF/MEF2
Ancestral somatic
striated CRC
Striated
effectors
Gut myocytes branch
Vertebrate gut
smooth myocyte
Platynereis gut
smooth myocyte
FoxF1
Nkx3.2
GATA6
Drosophila gut
striated myocyte
FoxF
Myocardin
NK3
SRF/MEF2
GATA456
MEF2
Smooth contractile and
differentiation effectors
Platynereis
cardiomyocyte
GATA4
MEF2
Gut myocyte
differentiation
effectors
NK4
Myocardin
Nkx2.5
SRF/MEF2
Striated
effectors
Striated contractile
effectors
Myocardin
FoxC
Ancestral gut
smooth CRC
NK4
Myocardin
SRF/MEF2
GATA4/5/6
Smooth contractile effectors
Differentiation effectors
(morphogenesis, innervation, adhesion)
Ancestral
cardiac CRC
Smooth
effectors
Fox
NK
Myocardin
SRF/MEF2
GATA4/5/6
Ancestral visceral
smooth CRC
Smooth contractile effectors
Differentiation effectors
Brunet et al. Figure 7 - supplement 1
MEF2
Striated
effectors
striation
striation
SRF/MEF2
Pannier
Smooth
effectors
FoxF
NK3
tinman
Myocardin
?GATA4/5/6 SRF/MEF2
striation
GATA4/5/6
Drosophila
cardiomyocyte
FoxC1
bagpipe
SRF/MEF2
Cardiomyocytes branch
Vertebrate
cardiomyocyte
biniou
Myocardin
Smooth contractile and
differentiation effectors
Evolution of visceral smooth CRC
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 7—supplement 1. Evolution of myogenic Core Regulatory Complexes (CRC)
2
in Bilateria. Transcription factor families are depicted as in Figure 1. Direct contact
3
indicates proven binding. Co-option of the fast/striated module happened on three
4
occasions: in Drosophila gut myocytes, and in cardiomyocytes of both vertebrates and
5
Drosophila. Note that in both cases, composition of the CRC was maintained in spite of
6
change in the effector module. In insect gut myocytes, replacement of the smooth by the
7
striated module entailed a split of the CRC, with the ancient smooth CRC still controlling
8
conserved differentiation genes (involved in adhesion, morphogenesis, axonal guidance,
9
or formation of innexin gap junctions), while the striated contractile cassette is
10
downstream Mef2 alone (Jakobsen et al. 2007b). It is less clear whether a similar split of
11
CRC took place in striated cardiomyocytes.
12
13
45
A
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
Deuterostomia
Protostomia
peer-reviewed)
is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International
license.
Tardigrada
Nematoda
Priapulida
Annelida
Arthropoda
Ecdysozoa
Mollusca
Chaetognatha
Chordata
Echinodermata
Spiralia
1.00
1.00
0.74
0.73
0.97
0.97
1.00
Ultrastructure of midgut muscles
Smooth
One-sarcomere
Striated
0.82
0.83
0.54
0.82
0.89
Insecta
0.90
0.77
0.77
0.69
Peripatopsis
Arenicola
Siboglinum
0.80
Panarthropoda
0.66
0.61
0.53
Helobdella
Eisenia
0.62
0.88
Ecdysozoa
Annelida
Mytilus
Helix
Rossia
Lepidopleurus
Meiomenia
Mollusca
Hemithris
Cephalodiscus
Rhabdopleura
Hemichordata
Saccoglossus
Cephalochordata
Urochordata
Vertebrata
Chordata
Spiralia
Crustacea
Deuterostomia
Flabelliderma
B
Chelicerata
0.98
0.62
0.79
0.73
0.63
0.56
Brunet et al. Figure 7 - supplement 2
Striated heart
myocytes
Smooth heart
myocytes
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 7—supplement 2. Ancestral state reconstructions of the ultrastructure of
2
midgut/hindgut
3
reconstruction, of midgut smooth muscles in Bilateria. Ancestral states were inferred
4
using Parsimony and Maximum Likelihood (ML) (posterior probabilities indicated on
5
nodes). Character states from: Chordata (Marieb & Hoehn 2015), Echinodermata (Feral
6
& Massin 1982), Chaetognatha (Duvert & Salat 1995), Mollusca (Royuela et al. 2000),
7
Annelida (Anderson & Ellis 1967), Priapulida (Carnevali & Ferraguti 1979), Nematoda
8
(White 1988), Arthropoda (Goldstein & Burdette 1971), and Tardigrada (Shaw 1974).
9
(B) Distribution, and ancestral state reconstruction, of cardiomyocyte ultrastructure in
10
Bilateria. Ancestral states were inferred using Parsimony and ML (posterior probabilities
11
indicated on nodes). Note that, due to the widespread presence of striated cardiomyocytes
12
in bilaterians, the support value for an ancestral smooth ultrastructure in the ML method
13
remain modest (0.56). This hypothesis receives independent support from the comparison
14
of CRCs (Figure 1B, Figure 7—supplement 1) and developmental data (see Discussion).
15
Character states follow the review by Martynova (Martynova 2004; Martynova 1995) and
16
additional references for Siboglinum (Jensen & Myklebust 1975), chordates (Hirakow
17
1985), Peripatopsis (Nylund et al. 1988), arthropods (Tjønneland et al. 1987), Meiomenia
18
(Reynolds et al. 1993), and Lepidopleurus (Økland 1980). In Peripatopsis and
19
Lepidopleurus, some degree of alignment of dense bodies was detected (without being
20
considered regular enough to constitute striation), suggesting these might represent
21
intermediate configurations.
and
heart
myocytes.
(A) Distribution,
and
ancestral
state
22
46
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
A
Homo sapiens
Platynereis dumerilii
Hsa-calponin1
Pdu-calponin
Hsa-calponin2
Pdu-transgelin1
Hsa-calponin3
Pdu-transgelin2
C
am
auhm
Drosophila melanogaster
Hsa-transgelin1
mp20
Hsa-transgelin2
Hsa-transgelin3
Caenorhabditis elegans
vlm
ach
chd64
vom
Domains
Calponin homology
unc-87
Calponin repeats
B
Branchiostoma-floridae-calponin
Danio-transgelin
0,9907
Danio-transgelin2
Homo-transgelin2
1
Rattus-transgelin2
1
Rattus-transgelin
1
Homo-transgelin
Homo-transgelin3
Mus-calponin2
1
Homo-calponin2
1
Xla-calponin2B
Latimeria-calponin3-isoformX1
1
Latimeria-calponin3-isoformX2
0,5786
Homo-calponin3
1
Oryzias-calponin3-like
Danio-calponin3b
0,99730,9787
0,9973Danio-calponin3a
Astyanax-calponin3-like
Danio-calponin2
0,5846
Xiphophorus-calponin1-like
1
Takifugu-calponin1-like
Meleagris-calponin1
1
0,5959
Gallus-calponin1
0,52
Chrysemys-calponin1
0,5959
Mus-calponin1
1 Sus-calponin2
0,9794
Sus-calponin3
Homo-calponin1
Ceratitis-capitata-myophilin-likeX1
0,998
Drosophila-chd64
1
Musca-domestica-myophilinX1
1
Riptortus-pedestris-calponin-transgelin
0,9734
Daphnia-pulex-DAPPUDRAFT_308120
0,8702
Nasonia-vitripennis-myophilin
0,98
Bombyx-mori-transgelin
0,9927
Pdu-transgelin1
1
Pdu-transgelin2
0,5473
Echinococcus-myophilin
0,9953
1
Drosophila-mp20
Metaseiulus-occidentalis-mp20-like
Solen-grandis-calponin-like-protein
0,5839
Mytilus-galloprovincialis-calponin-like
0,7603
Helobdella-robusta-HELRODRAFT_186180
1
Capitella-teleta-CAPTEDRAFT_17776
1
1
Helobdella-robusta-HELRODRAFT_185231
Cerebratulus-lacteus-calponin
Capitella-myophilin
0,715
Helobdella-robusta-HELRODRAFT_184984
Pdu-calponin
Lottia-gigantea-LOTGIDRAFT_149649
Aplysia-californica-myophilin-like
Crassostrea-gigas-Mp20
Transgelin
clade
0,8256
1
Deuterostomia
Calponin
clade
0,7643
1
1
0,7716
Protostomia
transgelin1
DAPI
D
midgut
transgelin1
DAPI
0.2
Brunet et al. - Figure 7 supplement 3
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Figure 7—supplement 3. Domain structure, phylogeny and expression patterns of
2
members of the calponin gene family. (A) Domain structure of calponin-related
3
proteins in bilaterians. Calponin is characterized by a Calponin Homology (CH) domain
4
with several calponin repeats, while Transgelin is characterized by a CH domain and a
5
single calponin repeat. In vertebrates, Calponin proteins are specific smooth muscle
6
markers, and the presence of multiple CH repeats allows them to stabilize actomyosin
7
(Gimona et al. 2003). Transgelins, with a single calponin repeat, destabilize actomyosin
8
(Gimona et al. 2003), and are expressed in other cell types such as podocytes (Gimona et
9
al. 2003), lymphocytes (Francés et al. 2006), and striated muscles (transiently in mice (Li
10
et al. 1996) and permanently in fruit flies (Ayme-Southgate et al. 1989)). (B) Maximum
11
Likelihood phylogeny of the calponin/transgelin family based on alignment of the CH
12
domain. Paralogs with calponin and transgelin structures evolved independently in
13
vertebrates and Platynereis (Pdu, in red squares). (C) Expression patterns of Pdu-
14
transgelin1. Scale bar: 25 µm. As in vertebrates and insects, transgelin1 is not smooth
15
myocyte-specific, but also detected in striated myocytes.
16
17
47
A
CapsaCalmo
bioRxiv
preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
BosCalmo
0.419
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
DmeCalmo
ThtrCalmo
0.763
Fungi
Fungi MRLC
Fungi MRLC
FalbaMRLC
0.747 Sponge MRLC
0.99
Sponge MRLC
OscMRLC
1
0.87
0.97 Cnidarian MRLC
0.997
0.291
0
Porifera
Cnidaria
BflMRLCb
Striated MRLC
0.718
CteMRLCa
0.99
Platynereis-ST-MRLC
0.77
0.628
0.99
RypaRLC
Mollusc striated MRLC
0.923
0.48
0.74
Arthropod muscle MRLC
1
0.99
0.88
1
Arthropod/nematode muscle MRLC
0.006
Vertebrate striated MRLC
0.86
1
0.8
SpuMRLCa
0.678
SkoMRLCb
SpuMRLCb
0.715
SpuMRLCc
0.83
1 Nematode non-muscle MRLC
Smooth MRLC
0.77
SkoMRLCa
0.86
0.96 Insect non-muscle MRLC
Amphioxus MRLCa
0.71
0.87
Vertebrate-smooth/non-muscle MRLC
0.98
Platynereis-SM-MRLC
0.833
HelobdMRLC
Mollusc smooth MRLC
0.9
BatdenMRLC
Fungi
0.606
MonveMRLC
0.4
Dme-biniou
B
Pdu-FoxF
Cgi-LBD3
C
1
Aca-LBD3
0.9956
Cte-FoxF
0.9878
1
1
Patvu-FoxF
1
Hro-LBD3
1
Brafl-FoxF
Cte-LBD3
1
Mus-FoxF2
1
1
Hsa-FoxF2
1
1
Pdu-ZASP/LBD3
Mus-FoxF1
GgaLBD3X12
Xenla-FoxF1B
1
1
1
Xenla-FoxF1A
HSALBD3X14
0.9612
1
Hsa-FoxF1
DreCyphX3a
Mus-FoxQ1
1
Cte-FoxQ1
0.2
Brunet et al. Supplementary Figure 1
Nve-LBD3
0.2
Bilateria
0.704
D
bioRxiv preprint first posted online Jul. 20,
2016; doi: http://dx.doi.org/10.1101/064881
. The copyright holder for this preprint
(which was not
Cte-Titin
Cel-MEF2
HSA-MYH7
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
0
DANRE-AMHC
0
DANRE-MYH1
1
F
Tt-Mef2
0.99
1
HSA-MYH1
Dme-MEF2
1
1
DME-ST-MHC
0.992
G
1
Dre-TitinA
1
1
Hsa-MEF2D
Dre-TitinB
1
1
Xenla-MEF2D
BGANMMHCX3
SkoTitin
NemvMef2X1
Bfl-TropT
Cint-TropT-X5
0.986
Hsa-Titin
1
DanreMEF2D
SKOMYH10X2
0.2
1
Danre-MEF2C
0.94
ACANMMHCX1
1
Mus-Titin
0.5876
MUS-MYH11
Platynereis-SM-MHC
0.984
Cel-Titin
Xenla-MEF2C
1
LPO-NMMHC
0.989
PolisTitin
0.79
0.94
HSA-MYH11
DME-NM-MHC
0.999
1
Hsa-MEF2C
GGA-MYH11
1
Limu-Titin
DanreMEF2A
1
DANREMYH11
0.999
PrcauTitin
XenlaMEF2A
MUS-MYH9
1
1
0.91
HSA-MYH9
Aca-titin
Cgig-Titin
1
Octo-Titin
Hsa-MEF2A
1
MUS-MYH10
0
1
DanreMEF2B
1
0
0.6208
0.6086
1
0.85
HSA-MYH10
1
Lgi-Titin
1
Hsa-MEF2B
DME-M-MHC
1
Hro-Titin
Pdu-titin
Platynereis-ST-MHC
0.994
1
1
Pdu-mef2
HSA-MYH8
1
1
E
SpurTwitch
0.2
H
Onchocerca-calponin-like
1
0.2
Calponin/transgelin clade
0.9996
Hsa-TropI2
0.9997
Oryctolagus-TropI2
Dre-TropT-cardiac
0.9
0.9982
0.34
0.8288
Hsa-TropT-skeletal
0.738
0.962
Xla-TropI3
Mus-TropT-cardiac
0.6167
0.886
Rattus-TropI3
0.7829
Drome-TropT
0.9999
Hsa-TropI3
Gga-TropI3
0.9998
Cte-TropT
0.849
Mus-TropI3
0.9999
Hsa-TropT-cardiac
0.986
Gga-TropI2
Hsa-TropI1
Coturnix-TropI3
Helro-TropT
1
Dm-TropI
1
Astacus-TropI
0.227
Pdu-TropT
0.872
Ce-TropI3
1
1
Cgi-TropT-X8
0.9998
0.79
0.845
0.9991
Linan-TropT
Bbe-MyoD2
0.3803
I
Bbe-MyoD1
Mus-MYKL2
0.8647
Hsa-MKL1
0.6341
1
Hsa-MYOD1
Mus-MYKL1
Xenla-MRTFA
1
0.2914
Hsa-Myf5
Gga-MYCD
Xenla-Myf5
Hsa-MYCD
1
0.5938
Dre-Myf5
0.4319
1
Mus-MYCD
0.6741
Mus-Myog
Xenla-MYCD
0.2719
HSA-Myf6
0.6148
Pdu-TropI
1
0.2224
Xenla-MyoD
0.3348
Chlamys-TropI
0.4
Hsa-MYLK2
J
DanreMyoD
0.5046
Ce-TropI1
Ce-TropI4
Sma-TropT
0.4
Ce-TropI2
0.9678
1
Pdu-MYCD
Spur-MyoD1
0.9302
0.5143
Sacko-MyoD
Lottia-gigantea-LOTGIDRAFT_111577
0.7073
Crassostrea-gigas-MKL-myocardin-like-protein-2
Tt-MyoD
1
0.5314
Daphnia-pulex-DAPPUDRAFT_309984
Pdu-MyoD
0.9989
0.5938
Spur-MyoD2
0.5326
0.8282
Drosophila-melanogaster-Myocardin-related-transcription-factor-isoform-H
Stegodyphus-mimosarum-MKL-myocardin-like-protein-2
Dme-nau
0.07
0.2
Brunet et al. Supplementary Figure 1
CapsaCalmo
BosCalmo
DmeCalmo
ThtrCalmo
0,291
MucorMRLC
LichcoMRLC
0,988
0,482
0,006
0,878
0,987
0.4
0,715
SpuMRLCa
0,678
SkoMRLCb
SpuMRLCb
SpuMRLCc
AscMRLC
WuchMRLC
1
CelMLC4
AncyMRLC
0,77
SkoMRLCa
HarpeMRLCn
0,83
0,87
AnopMRLCnm
BombyxMRLC
DanaMRLCsm
DmeMRLCnm
CeraMRLCnm
BflMRLC
1 BflMRLCa
Callorhinc
CalMRLC12B
LepiMRLC
1
GgaMRLCsm
0,868
RatMRLCsm
HsaMRLCsm
PterMRLCsm
BosMRLCsm
XentrMRLCs
TtdnMRLC
FuguMRLC9
AstyMRLC9
IctaMRLCsm
OryzMRLCsm
OreoMRLCsm
0,833
Pdu-SM-MRLC
HelobdMRLC
CrassosMRL
0,343
LgiMRLCb
0,901
ApcalMRLC
0,965
0,791
0,718
0,768
LeucMRLC6
SyconMRLCa
0,747
LeucMRLC19
LeucMRLC21
1
SyconMRLCb
LeucMRLC23
EphmuMRLC
0,988AmquMRLC
OscMRLC
0,87
MontaMRLC3
0,971
1
NveMRLCb
MetseMRLCb
0,971
0,997
FalbaMRLC
0,738
BatdenMRLC
MonveMRLC
AllomaMRLC
0,763 SalpunMRLC
0,704
1
0,606
Brunet et al. Supplementary Figure 1
K
0,864
0,923
0,628
0,991
Pdu-ST-MRLC
1 PlmaMRLC
AirMRLCm
ScolopMRLC
TriboMRLC
0,797
0,997
ArtfraMRLC
0,422
DaphMRLC
1
1
DaphMRLCb
BmorMRLCb
0,596
DmeMRLCm
0,98
AnopMRLCb
TrichiMRLC
1
LoaMRLC1
Necator-am
0,922
CelMLC1
CelMLC2
MusMRLCat
GgaMRLCsk
0,989
HsaMRLCsk
RbMRLCsk
0,904
RatMRLCsk
GgaMRLChe
0,595
HsaMRLCve
RatMRLCve
0,951
BflMRLCb
LgiMRLCd
LgiMRLCa
PfuMRLC
RypaRLC
CteMRLCa
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Supplementary Figure 1. Phylogenetic trees of the markers investigated.
2
(A) Simplified Maximum Likelihood (ML) tree for Myosin Regulatory Light Chain (full
3
tree in panel M), rooted with Calmodulin, which shares an EF-hand calcium-binding
4
domain with MRLC. (B) ML tree for FoxF, rooted with FoxQ1, the probable closest
5
relative of the FoxF family (Shimeld et al. 2010). (C) MrBayes tree for bilaterian
6
ZASP/LBD3, rooted with the cnidarian ortholog (Steinmetz et al. 2012). (D) ML tree for
7
bilaterian Myosin Heavy Chain, rooted at the (pre-bilaterian) duplication between smooth
8
and striated MHC (Steinmetz et al. 2012). (E) MrBayes tree for Mef2, rooted by the first
9
splice isoform of the cnidarian ortholog (Genikhovich & Technau 2011). (F) MrBayes
10
tree for Titin, rooted at the protostome/deuterostome bifurcation (Titin is a bilaterian
11
novelty). (G) MrBayes tree for Troponin T, rooted at the protostome/deuterostome
12
bifurcation (Troponin T is a bilaterian novelty). (H) MrBayes tree for Troponin I, rooted
13
by the Calponin/Transgelin family, which shares an EF-hand calcium-binding domain
14
with Troponin I. (I) MrBayes tree for MyoD, rooted at the protostome/deuterostome
15
bifurcation (MyoD is a bilaterian novelty). (J) MrBayes tree for Myocardin, rooted at the
16
protostome/deuterostome bifurcation (the Drosophila myocardin ortholog is established
17
(Han et al. 2004)). (K) Complete MRLC tree.
18
Species names abbreviations: Pdu: Platynereis dumerilii; Xenla: Xenopus laevis; Mus:
19
Mus musculus; Hsa: Homo sapiens; Dre: Danio rerio; Gga: Gallus gallus; Dme:
20
Drosophila melanogaster; Cte: Capitella teleta; Patvu: Patella vulgata; Brafl:
21
Branchiostoma floridae; Nve or Nemv: Nematostella vectensis; Acdi: Acropora
22
digitifera; Expal: Exaiptasia pallida; Rat: Rattus norvegicus; Sko: Saccoglossus
23
kowalevskii; Limu or Lpo: Limulus polyphemus; Trib or Trca: Tribolium castaneum;
48
bioRxiv preprint first posted online Jul. 20, 2016; doi: http://dx.doi.org/10.1101/064881. The copyright holder for this preprint (which was not
peer-reviewed) is the author/funder. It is made available under a CC-BY-NC-ND 4.0 International license.
1
Daph: Daphnia pulex; Prcau: Priapulus caudatus; Cgi or Cgig: Crassostrea gigas; Ling
2
or Linan: Lingula anatina; Hdiv: Haliotis diversicolor; Apcal or Aca: Aplysia
3
californica; Spu: Strongylocentrotus purpuratus; Poli or Polis: Polistes dominula; Cin or
4
Cint: Ciona intestinalis; Hro: Helobdella robusta; Bos: Bos taurus; Capsa: Capsaspora
5
owczarzaki; Thtr: Thecamonas trahens; Lpo: ; Bga: Biomphalaria glabrata; Cel:
6
Caenorhabditis elegans; Tt: Terebratalia transversa; Octo: Octopus vulgaris; Sma:
7
Schmidtea mediterranea; Bbe: Branchiostoma belcheri, Batden: Batrachochytrium
8
dendrobaditis; Monve: Mortierella verticillata; Alloma: Allomyces macrogynus; Salpun:
9
Spizellomyces punctatus; Mucor: Mucor racemosus; Lichco: Lichtheimia corymbifera;
10
Ephmu: Ephydatia muelleri; Sycon: Sycon ciliatum; Amqu: Amphimedon queenslandica;
11
Osc: Oscarella lobularis; Metse: Metridium senile; Pfu: Pinctada fucata; Rypa: Riftia
12
pachyptila; Plma: Placopecten magellanicus; Air: Argopecten irradians; Scolop:
13
Scolopendra gigantea; Artfra: Artemia franciscana; Bmor: Bombyx mori; Loa: Loa loa;
14
Necator-am: Necator americanus; Trichi: Trichinella spiralis; Asc: Ascaris lumbricoides;
15
Wuch: Wucheria bancrofti; Ancy: Ancylostoma duodenale; Callorinc: Callorhinchus
16
milii; Dana: Danaus plexippus; Anop: Anopheles gambiae; Asty: Astyanax mexicanus;
17
Oreo: Oreochromis niloticus; Icta: Ictalurus punctatus.
18
19
49