Invertebrate Muscles: Muscle Specific Genes and

Transcription

Invertebrate Muscles: Muscle Specific Genes and
Physiol Rev 85: 1001–1060, 2005;
doi:10.1152/physrev.00019.2004.
Invertebrate Muscles: Muscle Specific Genes and Proteins
SCOTT L. HOOPER AND JEFFREY B. THUMA
Neuroscience Program, Department of Biological Sciences, Ohio University, Athens, Ohio
I. Introduction
A. Why study invertebrate muscles, genes, and proteins?
B. Invertebrate and vertebrate phylogeny
C. Scope of review and literature database
II. Review of Vertebrate Muscle Specific Proteins
III. Invertebrate Muscle Proteins
A. Thin filament proteins
B. Thick filament proteins
C. ␣-Actinin and other Z line proteins
D. Ca2⫹ binding proteins and their targets
E. Giant sarcomere-associated proteins
F. Miscellaneous other proteins
G. Summary
1001
1001
1001
1002
1003
1005
1006
1014
1020
1020
1021
1026
1028
Hooper, Scott L., and Jeffrey B. Thuma. Invertebrate Muscles: Muscle Specific Genes and Proteins. Physiol Rev
85: 1001–1060, 2005; doi:10.1152/physrev.00019.2004.—This is the first of a projected series of canonic reviews
covering all invertebrate muscle literature prior to 2005 and covers muscle genes and proteins except those involved
in excitation-contraction coupling (e.g., the ryanodine receptor) and those forming ligand- and voltage-dependent
channels. Two themes are of primary importance. The first is the evolutionary antiquity of muscle proteins. Actin,
myosin, and tropomyosin (at least, the presence of other muscle proteins in these organisms has not been examined)
exist in muscle-like cells in Radiata, and almost all muscle proteins are present across Bilateria, implying that the
first Bilaterian had a complete, or near-complete, complement of present-day muscle proteins. The second is the
extraordinary diversity of protein isoforms and genetic mechanisms for producing them. This rich diversity suggests
that studying invertebrate muscle proteins and genes can be usefully applied to resolve phylogenetic relationships
and to understand protein assembly coevolution. Fully achieving these goals, however, will require examination of
a much broader range of species than has been heretofore performed.
I. INTRODUCTION
A. Why Study Invertebrate Muscles, Genes,
and Proteins?
Although pure research has its defenders, science is
generally justified by perceived human benefit. Two arguments suggest that studying invertebrate muscle genes and
proteins can reveal generally applicable principles that
could benefit humans. First, the last common ancestor of
vertebrates and invertebrates had muscle, and most vertebrate muscle genes and proteins have invertebrate homologs. The experimental advantages of invertebrate preparations often allow these genes and proteins to be investigated more easily, or at a greater level of detail, than is
possible in vertebrates. Many human diseases result from
errors in muscle protein structure, and thus invertebrate
www.prv.org
studies have the possibility of improving human health. Second, invertebrate muscle genes and proteins show great
variation. Despite this variety, in all cases the proteins must
functionally interact correctly. Comparative studies therefore provide a rich arena in which to investigate the relationship between protein assembly structure and activity,
again an area with clear relevance to human well-being.
B. Invertebrate and Vertebrate Phylogeny
Figures 1–3 show a contemporary, molecular biology-based, tree of life (4 and sources listed in the legend
to Fig. 1). Three things are of particular importance. First,
Cnidaria (corals, jellyfish) are separate and equal to Bilatera. All Bilatera are thus equally distant from all Cnidaria. Second (as has been long known), echinoderms,
tunicates, and amphioxus are more closely related to
0031-9333/05 $18.00 Copyright © 2005 the American Physiological Society
1001
1002
SCOTT L. HOOPER AND JEFFREY B. THUMA
FIG. 1. Phylogeny of Animalia (except Ecdysozoa, see Figs. 2 and 3). This phylogeny and those in Figs. 2 and 3 are from Tree of Life
(http://tolweb.org/tree/phylogeny.html; Coordinator, David Maddison, Univ. of Arizona), the University of California Museum of Paleontology
Phylogeny exhibit (http://www.ucmp.berkeley.edu/exhibit/phylogeny.html), and Systema Naturae 2000 [SJ Brands (comp.) 1989 –2002. Systema
Naturae 2000. Amsterdam, The Netherlands; http://sn2000.taxonomy.nl/]. In all three figures, gray lines represent unresolved relationships, and
differing horizontal line lengths (e.g., Platyzoa) are simply for figure composition, and in particular do not indicate evolutionary distance.
vertebrates than they are to other invertebrates. Third,
Ecdysozoa, Lophotrochozoa, and Deuterostomia, each of
which contains invertebrates, are presently coequal
branches. This tripartite split will presumably be eventually resolved into two bipartite branchings, but whether
the Ecdysozoa and Lophotrochozoa, or one of them and
the Deuterostomia, will end up being most closely related
is as yet unclear. This issue has profound implications for
comparative research since, depending on the ultimate
resolution of bilaterian relationships, it may be that lobsters are more closely related to humans than they are to
leeches. Although research on muscle genes may help
resolve this issue, it is so far insufficient to do so. We have
therefore organized the data presented here in simple
concordance with the relationships shown in Figures 1–3.
C. Scope of Review and Literature Database
A comprehensive review of invertebrate muscle is
unavailable. However, our invertebrate muscle database
Physiol Rev • VOL
contains over 6,700 references, and this massive literature
cannot be covered in a single review. We therefore intend
to produce a series of canonic reviews covering all journal
articles (due to their limited availability, books and book
chapters are not included) on invertebrate muscle written
before 2005. To that end, the field has been divided into
subsets, of which this review covers the first. Subsequent
reviews will cover 1) thick filament, thin filament, and
sarcomere structure; the molecular basis of contraction
and its regulation; asynchronous muscle and catch; 2)
muscle and synaptic ultrastructure and excitation/contraction coupling; 3) voltage- and ligand-gated ionotropic
channels; 4) metabotropic channels (modulation); and 5)
integrative properties and production of behavior. Even
with this broad net, some boundaries had to be drawn. In
particular, papers dealing with metabolic pathways are
generally not included, and no attempt to cover molting
and muscle proteases, regeneration, or muscle development has been made (papers that identify regulatory regions in muscle protein genes are included, but papers
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
FIG.
1003
2. Phylogeny of Ecdysozoa (except Insecta, see Fig. 3).
further “upstream” are not). The database is available at
http:\\crab-lab.zool.ohiou.edu\invert. Every effort will be
made to maintain this site for at least 10 years from
publication date.
II. REVIEW OF VERTEBRATE MUSCLE
SPECIFIC PROTEINS
Due to the number of references in this review, this
section is only sparsely referenced; References 6 and 128
are excellent reviews of vertebrate muscle. All muscles
contain thick and thin filaments. Thick filaments are composed of myosin. Myosin is composed of three pairs of
proteins: the heavy chain and the essential and regulatory
light chains (Fig. 4). The tails of the heavy chains form an
␣-helical coiled-coil tail. The other end of each heavy
chain and one essential and one regulatory chain form
one of the combined molecule’s two heads, each of which
contains an ATPase activity and can independently bind
to the thin (actin) filament. Trypsin severs the myosin tail,
resulting in light meromyosin, which contains only tail
Physiol Rev • VOL
(rod) sequences and heavy meromyosin (HMM), which
contains part of the tail and the two head regions (30, 42,
525, 1135, 1138). Further digestion of HMM results in the
S1 and S2 fragments, S1 consisting of the heads and S2 of
the HMM tail portion. Thin filaments are a double helix of
polymerized actin monomers. Tropomyosin and troponin
are two thin filament-associated proteins involved in contraction regulation in striated muscle (muscles with wellorganized sarcomeres, Fig. 5). The other type of vertebrate muscle, smooth muscle, does not have well-organized sarcomeres. Vertebrate smooth muscle contraction
is regulated both by myosin light-chain phosphorylation
by myosin light-chain kinase and a thin filament-based
regulatory system based on the actin-binding proteins
calponin and caldesmon.
All vertebrate striated muscle sarcomeres are very
similar (Fig. 5). Two Z lines, composed largely of ␣-actinin, define the sarcomere edges. The thin filaments attach
to each Z line and extend toward the center of the sarcomere. The region adjacent to each Z line containing only
thin filaments is the I band. One-half of each I band
85 • JULY 2005 •
www.prv.org
1004
SCOTT L. HOOPER AND JEFFREY B. THUMA
FIG. 3. Phylogeny of Insecta. Insects in red have asynchronous muscles, insects in green do not, and insects in black have not been examined
for this characteristic.
therefore belongs to one sarcomere and the other half to
the adjacent sarcomere. The thick filaments are located at
the center of the sarcomere. The region of the sarcomere
with only thick filaments is the H band, and the region
defined by their extent is the A band. At the very center of
the sarcomere there is often also a line (due to the presence of additional proteins) called the M line.
Sarcomeres consisting of only thick and thin filaments and Z lines would be inherently unstable. To appreciate this, consider a muscle fiber stimulated to contract while maintained in an isometric (constant length)
condition. If the thick filaments were exactly centered in
the sarcomere, equal numbers of myosin heads would
engage the thin filaments on the two sides of the M line,
the thick filament would feel equal force in both the right
and left directions, and the thick filaments would therefore remain centered in the sarcomere. However, if a
thick filament was even slightly uncentered, it would
experience greater force in one direction and would
therefore slide in that direction. The force imbalance on
the thick filament would now be even greater, and it
would thus continue to slide until it reached the Z line.
Solving this difficulty requires a mechanism that develops
Physiol Rev • VOL
FIG. 4. Myosin is composed of three paired molecules: the heavy
chain and the essential and regulatory light chains. Part of the heavy
chains form a coiled coil tail; the remainder of the heavy chains and the
two light chains form two globular heads. HMM, heavy meromyosin; S1,
S2, two HMM subfragments; LMM, light meromyosin (see text). [Modified from Rayment and Holden (972).]
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
1005
FIG. 5. Thick and thin filament arrangement in vertebrate striated muscle
sarcomere. Also shown are the sarcomere associated proteins ␣-actinin, titin,
myomesin, M-protein, and myosin binding proteins C and H. Red, actin; blue,
myosin; yellow, ␣-actinin; triangle, titin
serine/threonine protein kinase activity;
ellipse and circle in M line, myomesin
and M-protein, respectively.
a centering force if the thick filaments become uncentered. For example, if the thick filaments were attached to
the Z lines by springs, any thick filament movement away
from the center would decrease force in the shortened
spring, and increase force in the stretched spring, which
would recenter the thick filament.
Electron microscopic evidence of filaments linking
the thick filaments to the Z line (and which could thus
function as springs) was early obtained in both vertebrates and invertebrates (781). However, these filaments
are not composed of polymerized smaller subunits but are
instead enormous single proteins, and it took almost 20
years to characterize them chemically. Intriguingly, many
of them contain large numbers of immunoglobulin and
fibronectin III repeats. The largest of these proteins is titin
(3 MDa, ⬃30,000 amino acids, Fig. 5). Titin can be 1 ␮m in
length, and single titin molecules connect the M and Z
lines. Titin’s NH2 terminus extends through the Z line in
close association with the thin filament. At the A/I junction it leaves the thin filament and joins the thick filament,
with which it runs until reaching the M line, where it
overlaps the titin filaments from the other half-sarcomere.
The M line-associated proteins M-protein, myomesin, and
skelemin (which is generated from the myomesin gene by
alternative splicing) are other members of this family, as
are the A band myosin binding proteins C and H (also
called C-protein and 86-kDa protein) (Fig. 5). These proteins are much smaller than titin (a few hundred kiloDaltons), and unlike titin lack a serine/threonine kinase activity. Myosin light-chain kinase and telokin are two other
members of this protein family, found in smooth muscle.
III. INVERTEBRATE MUSCLE PROTEINS
We have attempted to order this review hierarchically. However, reviews, papers describing large numbers
of mutants with a common phenotype (e.g., flightlessness), and treatments describing the global advantages of
particular organisms do not nicely fit into this hierarchy
because they cover a large number of topics. Rather than
cite them repeatedly below, we instead summarize these
papers here.
Physiol Rev • VOL
For a recent general review of sarcomere structure
and proteins (primarily vertebrate, but includes some invertebrate work), see Reference 182; Reference 933 reviews all aspects of both vertebrate and invertebrate muscle. Reference 1136 reviews the early history of muscle
protein isolation and analysis. Reference 188 is a dated,
primarily vertebrate, review of calcium binding proteins,
troponin C, and myosin light chains. References 112, 276,
330 –332, 451, 1223; 17, 276, 1261; and 437 review, respectively, Drosophila, Caenorhabditis elegans, and amphioxus muscle. References 236, 237, 440 – 443, 571, 815,
869 and 43, 378, 673, 987, 1266, 1282, respectively, describe phenotypic mutant derivations in Drosophila and
C. elegans. Reference 333 reviews basic methods in Drosophila muscle biology. Gene expression profiles that
included muscle specific proteins have been performed in
jellyfish (Cyanea capillata) tentacle (1307); oyster (Crassostrea gigas) mantle (799); Mytilus galloprovincialis
(1210); the platyhelminths Clonorchis sinensis (652) and
Schistosoma japonicum (291, 1214, 1215); the nematodes
Brugia malayi (118, 119), C. elegans (719, 769, 774, 798,
1259), Globodera species (951), Haemonchus contortus
(436), Meloidogyne incognita (773), Onchocerca volvulus
(697), and Strongyloides stercoralis (798) [for nematodes,
multiple expressed sequence tag databases are now available on-line (926, 927, 1293)]; the mites Psoroptes ovis
(555) and Boophilus microplus (220); and amphioxus
notochord (1124) and Ciona intestinalis embryo (1048,
1049), larva (632, 1048), and adult (173, 1048), and cDNA
clones covering nearly 85% of C. intestinalis mRNA species are available (1050). References 1060 and 281, 814,
1159 show two-dimensional electrophoresis profiles of,
respectively, C. elegans and Drosophila proteins. References 814 and 1159 show that anatomically different Drosophila muscles (fibrillar vs. tubular) have different protein compositions. Also uncategorized are papers describing muscle post mortem changes, food-related properties,
and calorimetric measurements of muscle proteins (this
list is not comprehensive) (18, 19, 65, 170, 267, 269, 308,
367–372, 462, 463, 467, 493, 518 –520, 542, 543, 546, 565,
588, 589, 620, 687, 701, 753, 764, 765, 770, 807, 808, 839,
85 • JULY 2005 •
www.prv.org
1006
SCOTT L. HOOPER AND JEFFREY B. THUMA
840, 856, 865, 878, 899, 900, 907, 908, 910 –923, 1043, 1044,
1068, 1093, 1191, 1192, 1199, 1200, 1300, 1313–1316).
Actin and myosin heavy chain have been extensively
studied, and these sections are therefore subordered by
phylogenetic group. For some groups not all the articles
associated with it are explicitly covered in the text (e.g.,
descriptions of isolation techniques). For completeness,
in these cases all references dealing with that group are
listed immediately after the relevant subtitle. References
about purification of myosin as an oligomer (i.e., heavy
and light chains together) are listed in this manner in the
myosin heavy chain section. Table 1 provides actin and
myosin heavy chain data for groups for which only limited
information is available (for actin, Pterobranchia, Annelida, Gastropoda, Brachiopoda, Chaetognatha, Chelicerata; for myosin heavy chain, Cephalochordata, Urochordata, Annelida, Chaetognatha, Chelicerata).
A. Thin Filament Proteins
1. Actin
Mammals have six (two striated muscle, two smooth
muscle, and two cytoplasmic) (1203) and teleost fish have
nine (1211) actin isoforms. References 303 and 418 are
general (vertebrate and invertebrate) reviews of actin
molecular genetics, References 557 and 1025 review vertebrate and invertebrate actin isoforms, and Reference
499 reviews ascidian actin. Reference 238 shows that
TABLE
1.
muscle actins isolated from a variety of invertebrates all
have the same molecular weight and coelectrofocus with
the ␤-form of vertebrate smooth muscle actin, but are
immunologically distinct from each another.
A) CNIDARIA. Two coral actin cDNA clones have been
identified, one of which is expressed only in adults, the
only stage with muscles. Although not verified by in situ
hybridization, cnidaria may thus have a muscle specific
actin. The putative muscle actin gene showed greatest
homology to metazoan cytoplasmic actins (which metazoa are not clear from the article) (321). However, invertebrate muscle actins are typically most similar to vertebrate cytoplasmic actins (see below), and thus, depending
on which metazoans the authors used for comparison,
this similarity is not strong evidence that the adult actin
gene is not a muscle gene. Hydra has three or more
transcribed genes, but whether any are muscle specific is
unknown (304).
B) CEPHALOCHORDATA. Cephalochordate (Branchiostoma
only) actin gene number and expression are confusing.
Work using antibodies specific for vertebrate smooth
muscle, striated muscle, and cytoplasmic actins shows
that B. lanceolatum has at least three actin isoforms, with
the antismooth and antistriated antibodies staining separate sets of muscles and the anticytoplasmic antibody
staining most other cells (note that Ref. 1204 is wrong in
stating cephalochordates express only one muscle actin
isoform). Despite early confusion on this point (373, 946,
1234), in these animals the notochord is an innervated
Actin and myosin information for relatively little studied groups
Group/Protein
Pterobranchia actin
Annelida actin
Gastropod actin
Species
Muscle Versus
Cytoplasmic
Isoforms?
Identification Method
Muscle or
Developmental Stage
Specific Expression?
Reference Nos.
Saccoglossus kowalevskii
Glycera, Nereis, species
unreported
Helobdella triserialis
Two cDNAs
Protein purification
Unknown
Unknown
Unknown
Unknown
127
658
One cDNA, apparent
family
Unknown
1269
Aplysia californica
Aplysia californica
Protein purification
One cDNA, 3–5 gene
family
One cDNA
Cloning of 5⬘-flanking
and part of coding
region of one gene
Protein purification
Many tissues
including
muscle, but not
segmental
ganglia
Unknown
cDNA clone
muscle specific
Unknown
Unknown
Unknown
Unknown
401
246
Unknown
Unknown
643
366
Unknown
Unknown
658
Immunohistochemistry
Three genes
Unknown
Two genes muscle
specific
1011
1308
Protein purification or
immunohistochemistry
Expressed sequence tag
Unknown
Unknown
Differentially
expressed in
different muscles
Unknown
Unknown
Unknown
Zebra mussel
Haliotis rufescens
Lophophorata, Brachiopoda
actin
Chaetognatha actin
Glottidea, unreported
species
Sagitta friderici
Paraspadella gotoi
Chelicerata actin
Various
Cephalochordata myosin
Amphioxus (notochord)
Physiol Rev • VOL
85 • JULY 2005 •
www.prv.org
247, 658,
1021, 1204
1124
INVERTEBRATE MUSCLE GENES AND PROTEINS
muscle (310, 311, 511, 831, 1124, 1273). Interestingly, this
tissue nonetheless stains with the anticytoplasmic antibody (285, 1096).
Cloning in B. lanceolatum identified a cytoplasmic
and a muscle actin on the basis of “diagnostic” amino
acids (see below) (127). However, no tests for tissue
specific expression to verify these assignments were
made. Cloning in B. floridae identified one cytoplasmic
and two muscle actins on the basis of diagnostic amino
acids (623). In situ hybridization showed that staining for
one muscle actin was less strong in gill slit muscle and
that the cytoplasmic form was weakly expressed in axial
muscles in embryos, which suggests muscle and stage
specific variation in actin gene expression. Cytoplasmic
actin was present in notochord during early development,
but none of the three forms was present in adult notochord. Cloning in B. belcheri identified only one cytoplasmic and one muscle actin gene (on the basis of diagnostic
amino acids alone), but Southern blot analysis suggested
the presence of “a number” of additional actin genes
(624).
Expressed sequence tag analysis of a B. belcheri
notochord cDNA library suggests that notochord has
three actin genes, each of which codes for an identical
actin that differs from the cytoplasmic and muscle clones
already identified in this species (1124). The notochord,
muscle, and cytoplasmic actin amino acid sequences are
identical in the regions used to construct the oligonucleotide primers used to clone the muscle and cytoplasmic
actins. It is thus unclear why the notochord genes were
not identified in the earlier work (one possibility being a
difference in codon usage, although codon variation is
small for all amino acids in question, two of the seven
amino acids in each primer have unique codons, and for
the other five, only third nucleotide use varies). PCR
analysis of notochord, muscle, and ovary libraries shows
that one notochord actin gene is expressed only in notochord but some of the others are also expressed in muscle
(1124). The most conservative interpretation of these data
is that in Cephalochordata there are three actin groups
(cytoplasmic, muscle, and notochord), each of which may
contain more than one actin gene, but the number of
these genes, and which tissues each is expressed in, is not
clear.
C) UROCHORDATA. Urochordate (87, 173, 176, 428, 429,
500, 605, 623, 625– 632, 858, 875, 1047, 1049, 1096, 1126,
1181–1183, 1204, 1280, 1281) actin genes have been studied almost exclusively in Phlebobranchia (Ciona intestinalis, Ascidia ceradotes) and Stolidobranchia (Styela
clava, S. plicata, Halocynthia roretzi, Molgula oculata,
M. occulta). Ciona has eight nonmuscle and six muscle
actin genes (classified by diagnostic amino acid position,
not verified by in situ hybridization) (note Ref. 1204 is
wrong in stating urochordates express only one muscle
actin). Three muscle actin genes encode identical proPhysiol Rev • VOL
1007
teins (173). All six muscle actins have amino acid positions diagnostic of vertebrate striated muscle actin. Ciona
thus appears to have no counterpart to vertebrate smooth
muscle actin, a conclusion supported by the failure of
antismooth muscle antibodies to stain Ascidia muscles
(1096). Gene expression profiles show that muscle actin is
expressed in Ciona embryos (1049), larvae (632), and
adults (173). Five muscle actins are expressed only in
embryos and larvae, while the sixth is expressed solely in
adults (173). In neither embryos nor larvae was an actin
gene showing notochord specific expression mentioned
(632, 1049).
Two-dimensional gel electrophoresis indicates that
S. clava embryos, tadpoles, and adults contain three major and two minor actin isoforms. Two of the major
isoforms are likely cytoplasmic and the third a muscle
actin (1181). Cloning indicates four to seven muscle actin
genes (according to diagnostic amino acid position). The
four well-described genes encode identical proteins, but
show different temporal and spatial expression patterns.
In particular, one is expressed in a wide variety of tissues
(including nonmuscle), but only in larva and younger
animals, whereas another is expressed primarily in muscle cells in embryos but in nonmuscle tissue in adults.
None of the genes has been shown to be expressed at high
levels in adult muscle, which may suggest an adult-specific actin gene remains to be found. None of the clones
stained the notochord (87). A muscle actin gene (by diagnostic amino acid position) that shows some amino acid
variation (5.6%) from the S. clava muscle actins (and
which might thus be the “missing” adult muscle actin
gene) has been isolated from an adult S. plicata muscle
cDNA library (605). In embryo and larva, this gene is
expressed only in muscle cell lineages or functional muscle cells. Although the clone’s origin shows the gene is
expressed in adult muscle, whether it is expressed only in
muscle is unknown (1182).
H. roretzi adult body wall muscle contains two actins
differing in isoelectric point (875). The gene(s) coding for
these actins are not identified, but differ from those coding for larval muscle actin (625, 626, 628 – 630, 1047). H.
roretzi has seven larval muscle [verified by in situ hybridization (625, 630)] actin genes. The genes are arranged in
two clusters, one of which has five tandemly repeated
genes in the same orientation and the other of which has
two genes, arranged head to head on opposite DNA
strands, that share a common interposed promoter (626,
628, 629). Two of the actins are identical, and all are very
similar. Genes in each cluster have similar regulatory
elements and are believed to be coordinately controlled
(626, 628, 629, 1047). The sequences responsible for this
regulation are beginning to be identified (428). When reporter genes with a H. roretzi 5⬘-upstream muscle actin
gene (from the 5 gene cluster) flanking region are introduced into C. savignyi, reporter gene function is ob-
85 • JULY 2005 •
www.prv.org
1008
SCOTT L. HOOPER AND JEFFREY B. THUMA
served only in larval muscle cells, suggesting that larval
muscle actin regulatory processes have been conserved in
the two groups (429).
Two works in Molgula species show that changing
actin gene expression can affect morphological development. The first (627, 631) involves M. oculata, which has
typical tailed larvae, and M. occulta, which has tailless
larvae. M. oculata has two actin muscle genes (classified
by diagnostic amino acids, but verified in the larva by in
situ hybridization), one expressed in larva and the other
in adults. M. occulta has two actin genes orthologous to
the M. oculata larva gene, but in situ hybridization with M.
oculata-derived probes do not detect any muscle actin
production in M. occulta. However, when M. occulta gene
5⬘-flanking regions are attached to a reporter gene, reporter gene product is observed in M. occulta embryo
vestigial muscle cells. Thus both actin gene promoter
functionality and proper spatial production of the transacting factors that activate the genes are present in M.
occulta. The coding regions of the M. occulta genes, however, contain insertions, deletions, and codon substitutions that would result in their producing nonfunctional
actin. M. occulta taillessness thus appears to be due, at
least in part, not to changes muscle actin gene activation,
but to changes in gene coding regions such that the
activated genes produce no functional actin.
The second work involves precocious development
in M. citrina larvae (1126). In most Molgula species,
mesenchyme cells, believed to be adult muscle progenitors, remain undifferentiated in larvae, but in M. citrina
they begin to differentiate during the larval stage. A M.
citrina actin gene has been identified that is expressed in
juveniles and adults (it is not known if it is exclusively
expressed in muscle at these stages), but not in larva tail
muscle, suggesting that it codes for an adult muscle actin.
In situ hybridization shows that the gene is expressed in
the precociously differentiating mesenchyme cells in
larva and (at least) early after larval metamorphosis. Precocious development of adult features in this species is
thus likely also associated with precocious expression of
an adult muscle actin gene.
Larvacea muscle (classification confirmed by in situ
hybridization) actin genes have been investigated only in
Oikopleura longicauda (858). This work identified one
actin gene expressed in larval and adult tail muscle (larvaceans retain their tails as adults), but not in adult heart.
Undiscovered muscle actin genes thus presumably exist
in this species.
D) ECHINODERMATA. Actin genes have been studied in
echinoids (Stronglyocentrotus purpuratus, S. franciscanus, Lytechinus pictus, Heliocidadris erythrogramma,
H. tuberculata) and asteroids (Pisaster ochraceus, Dermasterias imbricata) (199, 215, 218, 258, 265, 292, 293,
342, 508, 513, 568, 569, 602– 604, 606, 650, 651, 909, 948,
1055, 1062, 1063, 1085, 1204, 1280, 1281). Echinoids have
Physiol Rev • VOL
6 –10 cytoplasmic actins, but all possess only one muscle
actin gene [early studies suggesting as many as 20 actin
genes in Stronglyocentrotus and Lytechinus species apparently being mistaken (265, 508)] (215, 218, 265, 292,
293, 342, 568, 569, 650, 651, 909, 1055, 1085). In S. purpuratus, the muscle actin gene is expressed at high levels
only in postpluteus muscle (1085). Dermasterias has
eight actin genes, but which are cytoplasmic and which
muscle is unknown (603). Pisaster has five actin genes,
one believed to be cytoplasmic, two muscle specific, and
two unspecified (identifications on the basis of cDNA
library source tissue, not verified by in situ hybridization)
(602– 604, 606).
E) MOLLUSCA. For bivalves, see References 153, 230, 296,
556, 558, 658, 738 –740, 859, 930, 1000, 1125, 1137, 1204,
1210, 1281, 1309; for gastropods, see Refs. 246, 366, 401,
643, 826, 1280, 1281; for cephalopods, see Refs. 161, 448,
595, 1106, 1204. Although one of the first techniques to
isolate large quantities of invertebrate thin filaments was
developed in bivalves (1137), and regulation of bivalve
actomyosin has been extensively studied (see second review), relatively little is known about bivalve actin genes.
Early electrophoresis work showed no difference in actin
across a wide range of bivalve species (740) or between
different tissues within one examined species (Spisula)
(739). Actin cDNA clones have been isolated from scallop
(Placopecten magellanicus) (930) and oyster (Crassostrea gigas) (153). The scallop actin gene is likely a
muscle actin and appears to be the primary actin expressed in adductor muscle. Southern blot analysis suggests 12–15 actin genes in the genome. Scallop actin
polymerizes more slowly than rabbit actin, and once polymerized, the cleft between actin subdomains 2 and 4 is
larger in scallop than in rabbit actin (556, 558). The oyster
cDNA was used to locate the gene and its upstream
region, but nothing is known about temporal or tissue
expression. Actin gene polymorphisms can be used to
identify clam species muscle (to prevent fraudulent use of
less desirable species in consumer products) (296) and to
study interspecies hybridization (230). Bivalve actin levels
change in characteristic ways in response to various pollutants (1000).
The information available about gastropod muscle
actin is presented in Table 1. Cephalopods are only
slightly better investigated. Southern blot analysis suggests that coleoid (all cephalopods but Nautilus) have at
least three actin loci. Phylogenetic analysis suggests three
actin isoforms, two of which were identified as muscle or
cytoplasmic on the basis of being, respectively, “related to
the mollusk muscle type” and “clustered among the other
mollusk cytoplasmic” actins (161). However, almost all
these sequences were obtained from GenBank, not from
published work showing tissue specificity. As such, although the sequence comparisons showing three actin
isoforms are likely valid, the identification of the isoforms
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
as to type needs experimental confirmation. A capillary
sodium dodecyl sulfate gel electrophoresis technique for
actin separation in squid has been developed (1106).
F) PLATYHELMINTHS. Cestoda is represented by Taenia
solium (15, 155), Diphyllobothrium dendriticum (1237–
1241), and Echinococcus granulosus (229). Taenia has
seven actin isoforms (15). Two genes, with identical coding sequences, have been cloned (155). Diphyllobothrium
has five actin genes (1239, 1241), three muscle specific
(1237). Two actin genes have been isolated from Echinococcus, and Southern blotting indicates as many as eight
(229). Trematoda is represented by Schistosoma mansoni (1, 234, 706, 901, 1342) and Fasciola hepatica (1119).
Two-dimensional gel electrophoresis identifies seven
Schistosoma actin isoforms, and two actin genes have
been cloned. Fasciola has three actin isoforms, one of
which is specific to tegumental spines. Turbellaria is represented by Dugesia lugubris, which has at least two
major actin isoforms, one muscle specific (928, 929).
G) NEMATODA. All information comes from Caenorhabditis elegans and Onchocerca volvulus (300, 608, 609, 646,
718, 902, 1052, 1204, 1235, 1264, 1280, 1281). In C. elegans
four actin genes have been identified, two of which produce identical proteins and none which differs by more
than three amino acids (300, 609). Three of the genes are
clustered (10, 300). All four genes are transcribed, and
three acquire a 22-nucleotide leader sequence via RNA
trans-splicing (608). Genetic evidence suggests that two
of the genes are involved in muscle thin filaments (646,
1264). However, the actins are so similar that they migrate
as a single species by isoelectric focusing (1052), no
isoform-specific antibodies have been generated, and thus
other evidence of tissue specific isoform expression is
lacking. One reference (902) states that all four genes
code for muscle specific actins (and refers to a personal
communication about a putative cytoplasmic actin gene),
but on what basis is unclear. Expression of total actin
varies during development (718). C. elegans actin can be
efficiently extracted at high purity (902). Some biochemical differences between C. elegans and rabbit actin have
been identified (902), and the structure of C. elegans actin
resolved at 1.75 Å resolution (1235). There is suggestive
evidence that C. elegans actin folding may be chaperoned
(667). O. volvulus has four actin genes that can be divided
into two classes on the basis of EcoR1 digestions. The
genes are arranged in two clusters, each of which contains one copy of each class (1328). Nothing is known
about tissue or stage specific expression.
H) CRUSTACEA. Crustacean (202, 526, 720, 783, 826, 1206,
1280, 1281) actin genes have been relatively little studied.
Artemia (species unreported) has 8 –10 genes and 4 isoforms, one muscle specific (709, 905). cDNA clones have
been isolated from the shrimp Marsupenaeus japonicus
and crayfish Procumbarus clarkii, but nothing is known
about their expression (526, 720). Crab (Gecarcinus latePhysiol Rev • VOL
1009
ralis) has seven or eight actin genes, and immunocytochemistry suggests some tissue and stage specific expression (1206). In lobster (Homarus americanus), different
muscles contain different amounts of total actin (783).
I) INSECTA. The references for insecta are as follows:
Diptera: Drosophila (16, 20, 64, 77, 89, 92, 111, 115, 129,
130, 204, 222, 259 –262, 334, 336, 339, 340, 350, 419, 430,
431, 449, 512, 536, 540, 634, 647, 658, 702, 703, 722, 733,
826, 829, 868, 897, 925, 974, 1038, 1039, 1041, 1057, 1108,
1109, 1175, 1176, 1280, 1281, 1347), Aedes aegypti (466),
Mayetiola destructor (1086), Dacus dorsalis (also known
as Bactrocera dorsalis) (411), Phormia regina (591); Lepidoptera: Bombyx (826 – 829); Coleoptera: Heliocopris japetus (139); Heteroptera: Lethocerus cordofanus (139,
146, 658, 1037). D. melanogaster has six actin genes that
are widely dispersed throughout the genome and produce
three major mRNA size classes (339). The genes are similarly dispersed in other Drosophila species (702). Gene
coding sequences, but not intron positions, are highly
conserved (334). Four (cytogenetic positions 57B, 79B,
87E, and 88F) of these genes code muscle actin (64, 111,
340, 430, 733, 826, 1041, 1175), at least some of whose
expression is muscle or developmental stage specific (64,
77, 111, 204, 340, 430, 536, 634, 868, 1041, 1175, 1176). In
particular, Act88F is expressed primarily in indirect flight
muscles (64, 340, 430, 722) [although it is coexpressed
with other muscle actin genes, and its absence causes
behavioral defects, in a small number of other muscles
(868)], Act79B is primarily expressed in “tubular” muscles
(an anatomically specific muscle type, see third review)
(64, 204, 340, 883) [an early report (1347) that Act79B is
the larval muscle actin apparently being in error], and
Act57B and Act87E are expressed in embryonic and larval
muscle (340, 1175) and a variety of adult nontubular,
nonindirect flight muscles (340). Similar data are obtained
from D. virilis (703) and, for Act88F, in D. simulans (92),
except that in D. virilis gene coexpression occurs in
more muscles (although this difference may stem from
enhanced sensitivity of modern techniques). Regulatory
regions of the Act57B, Act79B, and Act88F genes have
been identified in D. melanogaster (204, 350, 431), and the
Act88F gene promoter has been used to drive green fluorescent protein expression (11).
The different actins are functionally nonequivalent
(336), and (although they differ by only 15 amino acids)
mammalian cytoplasmic actin cannot substitute for
Act88F (129). Drosophila indirect flight muscle actin requires posttranslational modification for normal polymerization (419, 722, 1057) (which presumably underlies the
indirect flight muscle “actin III” reported in Ref. 449; see
also Refs. 430, 647) and is the only known actin to have an
unacetylated, free NH2 terminus (1057).
Multiple mutants of Drosophila indirect flight muscle
actin that alter muscle force production, despite in some
cases assembling into seemingly normal thin filaments,
85 • JULY 2005 •
www.prv.org
1010
SCOTT L. HOOPER AND JEFFREY B. THUMA
have been obtained (16, 20, 222, 260 –262, 430, 815, 897,
974, 1038, 1108, 1109), as have mutants that disrupt myofibril structure (16, 222, 260, 536, 540, 722, 1109). In a
study comparing several Act88F mutants, in almost all
cases protein stability was similar (261). In one mutant in
which muscle force generation is altered but thin filament
structure appears normal, the mutation site is distant
from the myosin binding site, and actin mutations can
therefore have long-range effects on force generation
(262). Experiments measuring the effect of actin mutation
on profilin, ATP, and DNase I binding showed similar
distant effects for some mutants (259). Other mutants
identified Glu93 as part of a secondary myosin binding site
(974), electrostatic charge on actin domain two as critical
for thin filament regulation by tropomyosin (115), and the
binding site of Clostridridium toxins (512). Mutants that
produce no actin still produce relatively normal thick
filament arrays, and experiments in which the actin-tomyosin ratio is altered suggest that filament imbalances,
not lack of thin filaments per se, are responsible for the
observed defects (89). Act88F epitope tagging on the
COOH terminus results in flightlessness and disordered
indirect flight muscle sarcomeres, but NH2 terminus tagging gives relatively normal sarcomeres and partially restores flight ability (130). Many [but not all (430, 540)]
actin mutations induce heat shock protein production
(260, 430, 431, 897, 925). The molecular basis of this
induction is unknown but is independent of myofibril
degeneration (430, 540, 1038, 1039).
Asynchronous muscle (a special type of muscle in
which muscle contractions are not synchronized with
motor neuron activity, see second review) in all species
examined in Nepomorpha, and some species in Diptera,
contain a ubiquitinated actin, arthrin (64, 93, 138, 1058).
Arthrin and the tropomyosin/troponin complex are in
equimolar concentration, suggesting that they may colocalize on the thin filament. This suggestion has not been
verified, but it is known that the ubiquination site is on the
opposite side of the thin filament from where tropomyosin binds (341). Arthrin’s function is unknown, as arthrin
activates myosin ATPase, is regulated by troponin/tropomyosin in the same manner as actin, and its presence
does not alter actomyosin kinetics (138, 1058). Arthrin is
not required for asynchronous muscle, as mutant Drosophila in which actin ubiquination cannot occur still fly
(1058), and arthrin has not been found in any Hymenoptera species with asynchronous muscle, and is absent
from some Cicadellidae (a family of Cicadelloidea),
Diptera, Coleoptera, and Heteroptera species, even
though all animals in these groups have asynchronous
muscle (Fig. 3) (937, 1058). Arthrin evolved independently
at least twice, although in all cases the ubiquination site
(Lys-118) is identical (148, 1058).
Actin genes from three other Diptera (A. aegypti, M.
destructor, D. dorsalis) believed to code for muscle actin
Physiol Rev • VOL
(on the basis of nucleotide and amino acid similarity to D.
melanogaster genes) have been investigated (411, 466,
832, 1086, 1236). Southern blot analysis suggests Aedes
contains at least five actin-related sequences (466) with
differential expression in different muscles (832, 1236).
Four muscle genes (identity confirmed by hybridization of
gene specific clones to RNA extracted from different tissues) have been cloned in D. dorsalis (411). Developmental stage specific expression and differential expression in
different muscles are present. The 3⬘- and 5⬘-flanking
regions of these genes show very little sequence homology both among themselves and to other known actin
genes. In Lepidoptera (B. mori) three actin genes have
been identified, one cytoplasmic and two muscle, one of
which is expressed only in adult muscle and the other in
both larval and adult muscles (827, 828).
J) ACTIN AS A PHYLOGENETIC CHARACTER. Much of the above
work investigates phylogenetic relationships (92, 127, 161,
218, 285, 292, 300, 334, 366, 500, 568, 605, 623, 624, 626,
627, 826, 829, 858, 903, 1085, 1204, 1239, 1269). Until
recently, the generally accepted conclusion was that all
invertebrate muscle actins are most closely related to
vertebrate cytoplasmic actins, with insect muscle actins
diverging very early from those of other invertebrates.
Unfortunately, much of this work may be seriously
flawed. First, much of it is based on diagnostic amino
acids (623), a small number of differing amino acids that
were early and successfully used to classify mammalian
actins, and then (uncritically) applied to invertebrate actins. This approach has been strongly criticized because it
1) treats the other amino acids as carrying no phylogenetic information and 2) does not analyze gene evolution
as a dynamic process of “descent with modification” (i.e.,
it is insufficient to just compare differences between two
extant species; instead, enough information from multiple
species must be obtained to infer the history of changes
that resulted in the present differences) (1281).
Second, actin gene duplications and conversions (including between muscle and cytoplasmic forms), which
complicate phylogenetic comparisons, have occurred in
many lineages (218, 258, 624, 1280). Third, Drosophila
actin genes show pronounced codon bias, which could
bias phylogeny construction (410). Fourth, actin may not
be particularly suitable as a phylogenetic marker. In most
organisms actin is a relatively large proportion of total
protein. Growth rate changes might therefore disproportionately affect actin evolution, and even closely related
invertebrate species can grow at very different rates.
Indeed, the actin gene duplications noted above may have
arisen specifically to allow rapid growth in some lineages.
Furthermore, different actin isoforms presumably at least
partly reflect a need for muscles with different functional
properties, a need that in many cases may depend on
organism life-style. Closely related invertebrate species
with different life-styles could thus have experienced
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
higher rates of actin evolution, again complicating phylogenetic analysis.
Fifth, actin-based phylogenies often have untenable
relationships (Fig. 6). For instance, in phylogeny A, sea
urchin and ascidia cytoplasmic actin are more closely
related to mollusk cytoplasmic actin than to pufferfish
cytoplasmic actin, even though sea urchin, ascidia, and
pufferfish belong to Deuterostoma and mollusks to Lophotrochozoa (Fig. 1). Similarly, in phylogeny B, sea urchin cytoplasmic actin is more closely related to Cnidaria
actin than to either ascidia or sea star cytoplasmic actin,
even though Cnidaria belongs to Radiata and the other
species to Bilateria. Again, in phylogeny C, sea star cytoplasmic actin is more closely related to urochordate cytoplasmic actin than it is to that of another echinoderm,
sea urchin.
Muscle actin is part of a highly organized ensemble of
proteins that might be expected to coevolve. A more
fruitful approach therefore may be to build phylogenies
1011
for multiple muscle proteins (e.g., actin, troponin, myosin
light and heavy chains) and superimpose the trees to
arrive at the ensemble’s most likely phylogeny. When this
is done (Fig. 6D), an ancestral set of muscle proteins gives
rise to two branches, one of which leads to the vertebrate
smooth muscle ensemble and the arthropod and vertebrate cytoplasmic ensembles. The other branch gives rise
to urochordate smooth muscle and all striated muscles.
Note that this phylogeny agrees with those determined by
other methods (Figs. 1–3); in particular, in each branch all
vertebrates are more closely related to each other than
they are to nonvertebrates, and urochordates are more
closely related to vertebrates than they are to arthropods.
2. Tropomyosin
Vertebrate muscle tropomyosin is a dimer of tropomyosin molecules arranged in an in-parallel, in-register
coiled-coil. There are typically two isoforms, and tropo-
FIG. 6. Various molecular geneticbased phylogenies. A–C are based on actin alone. D is based on superimposition
of troponin C, myosin essential and regulatory light chains, myosin heavy chain,
actin, and (for the vertebrates) muscle
regulatory factor trees. [A modified from
Carlini et al. (161); B modified from
White and Crother (1280); C modified
from White and Crother (1281); D modified from Oota and Saitou (903).]
Physiol Rev • VOL
85 • JULY 2005 •
www.prv.org
1012
SCOTT L. HOOPER AND JEFFREY B. THUMA
myosin can form as a hetero- or homodimer of either.
Nonmuscle tropomyosin isoforms are also widely
present. Multiple tropomyosin isoforms are generally
present in invertebrates, and most work has concentrated
on distinguishing muscle and nonmuscle forms. Whether
invertebrate tropomyosin is a dimer, and if so, a hetero- or
homodimer, is less investigated. However, equilibrium
sedimentation work in crayfish, oyster, abalone, and
blowfly indicate that the tropomyosins self-associate
(1291), and optical rotary dispersion and hydrodynamic
measurements suggest blowfly tropomyosin is 100% ␣-helical and in a two-stranded configuration (591). Given
these data, invertebrate tropomyosin is presumably also a
coiled-coil dimer. Muscle-type tropomyosin is in some
cases expressed in nonmuscle tissues, and in some work
below these tissues were used as tropomyosin source
material. Reference 653 is a general (vertebrate and invertebrate) review covering muscle and nonmuscle tropomyosins. In the older literature, paramyosin is sometimes called tropomyosin A or even tropomyosin (for
references, see sect. IIIB5 and second review); in this
work tropomyosin is called tropomyosin B. In two very
early articles, we are unable to determine, for the invertebrate portions, whether tropomyosin or paramyosin
was being studied (938, 1174).
Tropomyosin has been identified by expressed sequence tagging, immunohistochemistry, protein isolation,
or cloning in Cnidaria (53, 320, 385, 386, 699); amphioxus
(1124); Ascidia (785, 1183); sea urchin (485, 486, 490 – 492,
1177); annelids (221, 658); mollusks: bivalves (179, 404,
469, 472, 481, 482, 484, 488, 494, 524, 658, 672, 738 –740,
795, 931, 1152, 1254, 1291), gastropods (401, 480, 1291),
cephalopods (448, 477, 483, 595, 806, 1197); brachiopods
(658); Chaetognatha (1011); platyhelminths (283, 706,
1277, 1295); C. elegans (21, 515–517, 718) and other nematodes (39, 318, 388, 461, 501, 842); Chelicerata: mite
(1021, 1285), horseshoe crabs (658, 659, 802, 804, 805),
scorpion (805); Crustacea: lobster (202, 487, 670, 790, 803,
805, 837, 1042), shrimp (233, 668, 803, 805, 960, 981, 982,
1073, 1285), crayfish (103, 801, 803, 805, 1291), crab (669,
803, 805), hermit crab (803, 805), isopod (803, 805); and
various insects (41, 84 – 86, 139, 146, 394, 395, 537–539,
591, 614, 658, 805, 1291).
Sea anemone has five isoforms (320) of unknown
origin, but some tissue specificity is observed (320). Jellyfish has two tropomyosin genes (Ref. 53 is mistaken in
stating only a single gene exists). One is expressed only in
striated muscle (385). One tropomyosin gene has been
found in Hydra (699), but is not expressed in epitheliomuscular cells. Unidentified epitheliomuscular cell specific genes may thus exist. Expressed sequence tag analysis shows that tropomyosin is expressed in Amphioxus
notochord (1124). Striated and smooth muscle in ascidia
has an identical tropomyosin, coded by a single gene and
more similar to vertebrate striated than vertebrate
Physiol Rev • VOL
smooth muscle tropomyosin. This common expression
may be due to ascidian smooth muscle being troponin
regulated (785). The isoform is expressed only at low
levels in larval tail muscle, so undiscovered tropomyosin
genes may exist. Sea urchin has at least two isoforms,
antigenically identified as muscle or nonmuscle types
(492). The muscle type binds actin (1177) and is located in
muscle (486). Earthworm muscle possesses two isoforms
of unknown origin that combine as a heterodimer (221).
Bivalves and gastropods have as many as six isoforms
(401, 404, 482, 489, 740), some expressed in a muscle
specific manner (488, 489, 494, 931), and multiple tropomyosin genes (404, 494, 931). NH2-terminal blocking is
important for head to tail polymerization, actin binding,
and Ca2⫹ regulation in Akazara scallop tropomyosin
(473). Bivalve tropomyosin levels change in characteristic
ways in response to various pollutants (1000).
Platyhelminths have multiple isoforms, with some
tissue-specific expression and likely multiple genes (283,
1277). C. elegans has one tropomyosin gene. Four isoforms arise by alternative splicing and show developmental stage (718) specific expression and are differentially
expressed in different muscles (21, 515, 516). Limulus
has four to six isoforms that are not expressed in a
tissue-specific manner (804). Scorpion has four muscle
isoforms that show some muscle specificity (805). Crustaceans have three or four muscle isoforms. One is
uniquely expressed in cardiac muscle, and the others
show considerable but not perfect muscle specific expression (487, 801, 803, 805, 837, 960, 1042). Two of the
isoforms may arise by alternative splicing (837). Beetles
and centipedes have three muscle tropomyosins, which
show some muscle specificity (805). Locust has multiple
isoforms with muscle specific expression. A cDNA clone
has been sequenced, and two mRNAs are present, but
whether they are from different genes is unknown (614).
Drosophila has two muscle tropomyosin genes. Each produces four or five tissue or developmentally specific isoforms (including nonmuscle forms) by a combination of
alternative splicing and multiple promoters (84 – 86, 381,
394, 395, 537–539, 1064) (two of these isoforms are the
flight muscle specific troponin H, see below). Considerable evidence has been obtained in Drosophila on the
regions of the tropomyosin gene that regulates its expression (686, 689, 690, 788, 1064).
Drosophila (538, 613, 817, 1164) and C. elegans (21,
516, 1282) tropomyosin mutations alter muscle structure
and function. Some Drosophila mutants can be rescued
by reproviding correct tropomyosin sequences (337, 1163,
1164). Despite the partial tissue specificity noted above, in
one case substitution of different isoforms can also rescue tropomyosin mutants (794).
Tropomyosin is an important component of immune
and allergic reactions (7, 25, 29, 33, 37, 38, 40, 41, 50, 51,
150, 157, 158, 179, 233, 255, 295, 362, 403, 479 – 484, 501–
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
506, 661– 663, 668 – 672, 726, 752, 806, 970, 978 –983, 1032,
1045, 1073, 1166, 1172, 1173, 1277, 1278, 1285, 1295, 1296,
1312).
3. Troponin
Vertebrate troponin contains three subunits: troponins T, I, and C. Ascidia (273, 882, 1183); Glycera,
Lumbricus, and Nereis (annelids) (658); scallop (363, 861,
885, 888 – 891) and squid (477, 595, 892, 1081, 1193); nematode (461, 564); Limulus (658, 659); shrimp (895, 960),
lobster (202, 790, 852, 860, 985, 1081), and crayfish (103);
and Drosophila (139, 143) and Lethocerus (658) troponin
also contains three subunits (although with considerable
molecular weight variation). Troponin T is present in
cross and obliquely striated Eisenia (earthworm) muscle
(1012, 1019), Helix heart (1012), mite (1021), and Antarctic mussel shrimp (1016); troponin C in amphioxus (1149),
sandworm (Perinereis) (1326), mussel (1081), and barnacle (12, 35, 192); troponin I in amphioxus (1124); and
“troponin” (unidentified subunit) in Chaetognatha (Sagitta frederici) (1011). A fourth troponin, troponin H
(called a tropomyosin isoform, not troponin H, by some
authors), which is a fusion product of tropomyosin and a
hydrophobic proline-rich sequence, is alternatively
spliced from one of the tropomyosin genes, at least partly
replaces troponin I, and exists in all indirect (asynchronous) and some direct (synchronous) flight muscles (143,
395, 537, 539, 937, 977). In mollusks, high (relative to
vertebrates, but physiological for the organisms in question) monovalent ion or Mg2⫹ levels are required for
troponin and tropomyosin to remain bound to the thin
filament (363, 655). Lack of recognition of this need presumably is the reason that early studies found troponin
was not present (552, 658), or was present at only low
levels (660), in molluscan muscle.
Lobster actin:tropomyosin:troponin (with the troponin consisting of equimolar amounts of troponin T, I, and
C) (790), and scallop actin:tropomyosin:[troponin I:troponin C], are present in molar ratios of 7:1:1 (657), as are
Limulus actin:tropomyosin:[troponin T:troponin C], but
in this species twice the amount of troponin I may be
present (659). Shrimp troponin C, I, and T are present in
a 1:1:1 molar ratio (895). Scallop troponin I and C are
associated with the I band (657). Troponin is absent from
surf clam foot muscle (175). Reference 656 compares the
amino acid composition of troponin C in rabbit and a
variety of invertebrates.
Cloning work has been performed for ascidia (184,
714, 1321, 1323), scallop (893, 1158), tick (1317), crayfish
(575), and Drosophila (68, 88) troponin I; Amphioxus
(1149), ascidia (1144, 1322), sandworm (1326), scallop
(863, 886, 887, 894, 1121, 1325), squid (892), nematode
(1168), Limulus (573), lobster (343), crayfish (576), barnacle (192), fire ant (945), and Drosophila (328, 423)
Physiol Rev • VOL
1013
troponin C; ascidia (272), C. elegans (833), Drosophila
(101, 143, 335), and scallop (470, 471) troponin T; and
Drosophila troponin H (395, 537, 539). Troponin-based
phylogenies agree with Figures 1–3 (184, 187, 471, 1149,
1320). Multiple isoforms, some developmental stage or
muscle specific, are present in ascidia (271, 272, 714, 882,
1321, 1323); scallop (888, 1325); C. elegans (843); shrimp,
crayfish, and lobster (202, 343, 581, 783, 790, 834 – 836,
838, 852, 860, 895, 960, 985, 1083, 1286); barnacle (35, 192);
Lethocerus (959), Anopheles (959), dragonfly (305, 735–
737); and Drosophila (68, 69, 101, 257, 328, 335, 395, 423,
538, 539, 959). Stage-specific isoform changes are associated with increased Ca2⫹ sensitivity and altered twitch
contraction kinetics in aging dragonflies (305, 736), and
expression of the asynchronous muscle isoform with
flight ability in bee (257).
Halocynthia (ascidia) has one adult (which produces
two isoforms by alternative splicing) and at least three
larval troponin I genes; sequences regulating gene expression have been identified (1321, 1323). Ciona, alternatively, has only one troponin I gene (which again produces two isoforms), and no homologs to the Halocynthia larval genes (714, 1324). Only one troponin I gene has
been identified in Drosophila (68, 69, 88, 244); sequences
regulating gene expression have been identified (741,
760). Ascidia, amphioxus, sandworm, and scallop have
one troponin C gene (1319, 1322, 1325, 1326), C. elegans
and barnacle two (192, 1168), lobster three (343), and
Drosophila five (328, 423). The fourth intron of invertebrate troponin C genes shows great variability, suggesting
that the ancestral gene may not have possessed it (1325,
1326). Ascidia has two troponin T genes (271, 272) and
Drosophila and dragonfly one (101, 735). Troponin I, C, T,
or H mutants can disrupt muscle function or development
(68, 69, 88, 335, 516, 538, 771, 833, 866, 1168), although
some polymorphism is tolerated (224). Drosophila troponin I mutants can be suppressed by tropomyosin (841) or
myosin heavy chain (299, 615, 866, 867) mutations or
troponin I second site mutations (952).
4. Calponin/caldesmon
Calponin-like proteins have been identified in Eisenia (1017); Mytilus (325); Helix (1012); Echinococcus
[called myophilin (748 –751)] and Schistosoma (510,
1306); and Onchocerca (where it is highly immunogenic)
(475), but not in crustacean or Drosophila muscle (1012,
1017). Calponin-like cDNAs in a nematode (Meloidogyne
incognita) (162), Echinococcus (751), and Schistosoma
(1306) have been sequenced. Echinococcus has only one
gene, but multiple isoforms are expressed due to posttranslational modification, including phosphorylation by
protein kinase C (750). In some Schistosoma species,
several copies are present, and multiple isoforms are
expressed (1306). Caldesmon-like proteins have been
85 • JULY 2005 •
www.prv.org
1014
SCOTT L. HOOPER AND JEFFREY B. THUMA
identified in Eisenia (1017) and mollusks (76, 100, 225,
1012).
5. C. elegans unc-87
The unc-87 mutants have almost normal embryonic
muscles that become severely disorganized as the animals
mature (1266). unc-87 codes for a 40-kDa protein that is
located in the I band, bundles actin filaments (but not
monomers) in the absence of tropomyosin or ␣-actinin,
and has the same actin binding sequence as calponin (361,
607). When unc-87 mutants are expressed in animals with
decreased myosin heavy chain activity, the age-dependent
disorganization is less, suggesting that it arises from contraction force. unc-87 is therefore not believed to be
required for sarcomere formation, but instead to serve a
structural role (360).
B. Thick Filament Proteins
1. Myosin heavy chain
There are at least 13 classes of myosin; all muscle
(and some nonmuscle) myosins belong to class II (61, 104,
172, 200, 374, 433, 1069, 1070, 1198, 1302). We cover here
only muscle myosins. References 26, 66, 400, 776, and
1134 review vertebrate and invertebrate muscle and nonmuscle myosins, and Reference 42 reviews vertebrate
(primarily) and invertebrate domain structure. Reference
275 reviews myosin in C. elegans. We also do not review
here how myosin aggregates to form the thick filament, as
these data are in the second review, but do note that
individual scallop (and presumably other invertebrate)
combined myosin heavy and light chain molecules have
the structure shown in Figure 4 (i.e., possess a long rod
and have two globular heads separated by a cleft and
connected to the rod by flexible necks) (1222).
References 172, 200, 374, and 433 examine myosin
heavy chain motor domain phylogenetic relationships.
These works are primarily concerned with the relationships among the different myosin classes, and in particular contain muscle myosins from too few invertebrate
species to shed light on invertebrate phylogenetic relationships. The available data suggest two important (from
the point of view of this review) myosin II branchings.
First, vertebrate smooth muscle myosin and some invertebrate nonmuscle (but still class II) myosins early split
from the myosins leading to vertebrate striated muscle
and all invertebrate muscle myosins. A second split resulted in two branches, one of which contains all striated
(both skeletal and cardiac) vertebrate and invertebrate
myosins and the other of which contains nematode muscle myosin (which has a special type of striation called
obliquely striated, see third review). However, given the
paucity of invertebrate species contained in this work and
Physiol Rev • VOL
in particular the lack of smooth invertebrate muscles in it,
these groupings must be considered preliminary.
A) CNIDARIA. A cDNA clone of a striated muscle specific
myosin heavy chain has been sequenced and is more
similar to striated muscle heavy chain isoforms than to
smooth or nonmuscle isoforms in either vertebrates or
invertebrates. Because cnidarians also have smooth muscle, at least one other muscle isoform is presumably
present. Southern blotting suggests this/these isoform(s)
would be from other genes, not alternative splicing (5,
1061).
B) MOLLUSCA. For bivalves, see References 67, 163–165,
405, 452, 453, 496, 567, 594, 597, 599, 622, 738, 740, 862,
864, 873, 874, 941, 1101, 1246); gastropods, References 32,
52; cephalopods, References 563, 580, 768. Myosin heavy
chain cDNAs or genes have been sequenced in scallops
Argopecten (also known as Aequipecten) (873, 874), Patinopecten (405), Pecten (496), and Placopectin (941) and
squid Loligo (768). Scallop (Placopectin, Argopectin) has
three or four isoforms alternatively spliced from a single
gene and expressed in a muscle-specific fashion (874,
941). Mollusk striated and catch (a special muscle type
that maintains contraction with very little energy expenditure, see second review) muscle myosin heavy chains
differ almost exclusively in only one region (surface loop
1). The two isoforms have different ATPase activities
(941) and ADP affinities and dissociation rates (622). The
sequence responsible for light-chain positioning on the
heavy chain has been identified in the scallop (452). Regulatory light-chain kinase phosphorylates scallop myosin
heavy chain (1101), and Mytilus myosin contains a tightly
bound endogenous kinase (163). Which residues were
phosphorylated in these works is unknown, but phosphorylation of specific serines in Mytilus tailpiece increases myosin solubility and favors molecular folding
(164, 165). Fish light chains bind to scallop heavy chain
(567, 1101). Squid has at least two isoforms, alternatively
spliced from one gene (768).
C) PLATYHELMINTHS. See References 14, 387, 572, 904,
1028, and 1276 for platyhelminth data. Planaria have two
muscle myosin heavy chain genes expressed in nonoverlapping sets of muscle (572, 904). Although all planarian
muscles are nonstriated, planaria heavy chain myosins
most closely resemble the striated muscle myosin heavy
chains of other organisms (572). Phylogenetic analysis
using myosin heavy chain genes suggests that the platyhelminths are polyphyletic and that two platyhelminth
groups, the Acoela and Nemertodermatida, are the earliest extant bilaterians (1028). Myosin may be useful in
platyhelminth vaccine development (13, 108, 114, 241,
254, 268, 282, 509, 682, 849, 935, 1103–1105, 1338 –1340).
D) NEMATODA. See References 24, 94, 250, 251, 275, 278,
280, 302, 402, 445– 447, 497, 541, 693, 711, 715–717, 792,
793, 812, 844, 855, 898, 956, 1053, 1161, 1260, 1265, 1274,
1275, 1279, and 1329 for nematode data. C. elegans has
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
four myosin heavy chain isoforms (MHC A, B, C, D),
coded by widely separated genes (myo-3, unc 54, myo-2,
myo-1, respectively), some on different chromosomes
(10, 251, 280, 541, 715, 717, 793, 1053). MHC A and B are
used only in body wall muscle and MCH C and D only in
pharyngeal muscle (24, 280, 711). The molecular basis of
this differential expression is beginning to be investigated
(497, 898). MHC A is present only in the center of the thick
filament and MHC B only at the ends (792), possibly
because MHC A’s more hydrophobic rod surface binds
better to paramyosin (which forms the core of nematode
thick filaments, see second review) (447). Depending on
the mutation, MHC B mutants (280, 716, 717) are paralyzed with a reduced number of thick filaments composed
entirely of MCH A (94, 250, 278) or very slowly moving
with normally assembled thick filaments (250, 812). MHC
B overexpression disrupts muscle structure (302). MHC A
mutations are embryonic lethal and contain no thick filaments, presumably because MHC A is necessary for the
initial steps of thick filament formation (1260). Two regions of the MHC rod sufficient to allow thick filament
initiation have been identified (445, 447), as has a region
for interaction with paramyosin (446). The terminal, nonhelical tailpiece of MHC A is not required for thick filament formation, although the thick filaments are disorganized relative to control animals (445). MHC B, but not
MHC A, synthesis and accumulation is reduced in mutants
lacking paramyosin (1279), and MHC A overexpression
can partially compensate for MHC B and paramyosin
mutations (754). Myosin heavy chain B pre-RNAs have
been used to examine intron splicing in C. elegans (881).
Partial or complete myosin heavy chain cDNAs have been
sequenced in Schistosoma mansoni (855), Toxocara canis (877), and Onchocerca volvulus (1275), and a myosin
heavy chain gene has been sequenced in Brugia malayi
(1274). The Onchocerca and Brugia data are for body wall
(putative MHC B homolog) myosins. Myosin may be useful in nematode vaccine development (877, 961, 962).
E) CRUSTACEA. See References 55, 202, 203, 439, 570, 638,
684, 721, 783, 834, 1016, 1040, 1090 –1092, and 1346 for
crustacean data. Lobster has at least two myosin heavy
chain isoforms (684, 783), and crayfish three to four (638,
1040). Heavy chain isoform expression differs in fast versus slow muscles in crayfish (638, 1040) and lobster (203,
684, 783) (although Ref. 834 reported no difference) when
the muscles are examined in bulk, but individual fibers
show a continuum of myosin heavy chain isoforms (782).
ATPase activity differences between slow and fast muscles are due solely to heavy chain isoform differences
(684). Lobster abdominal myosin is more susceptible to
tryptic proteolysis and is cleaved at different locations
(also unlike vertebrate myosin, in a Ca2⫹ concentrationindependent manner), than is rabbit myosin (734, 1090,
1092, 1346). A specific myosin exists in the Antarctic
isopod Glyptonotus, presumably to maintain muscle funcPhysiol Rev • VOL
1015
tion at very low temperature (439), a conclusion supported by evidence that lobster does not express temperature-dependent myosin heavy chain isoforms (unlike, for
instance, fish) (721).
F) INSECTA. See References 89, 109 –111, 139, 180, 185,
223, 235, 347, 406, 424, 434, 435, 550, 591, 616 – 618, 692,
791, 816, 823, 879, 880, 1013, 1023, 1024, 1095, 1115, 1116,
1127–1129, 1132, 1162, 1253, 1272, 1302, 1336, and 1344 for
insect data. Drosophila has at least 15 myosin heavy chain
isoforms (all alternatively spliced from a single gene)
(109 –111, 185, 347, 406, 434, 435, 550, 1023, 1024, 1253);
many show tissue or developmental specificity (109, 180,
185, 347, 406, 550, 616, 816, 880, 1024, 1336; Refs. 823, 1132
are reviews). The molecular basis of expression specificity is beginning to be investigated (235, 347, 424, 434, 435,
1115, 1116). Exon variation in the heavy chain head region
suggests the region where light chain binding occurs is
important for myosin function (110). Exon variation in the
rod suggests that variation in the rod hinge is associated
with contraction rate changes (185). ATP-induced dissociation of actin from indirect flight muscle isoform thick
filaments is fast, similar to mammalian skeletal muscle
values (1095). However, comparing different isoforms
shows that they all have similar, fast, cross-bridge detachment rates (791) and unitary step (single cross-bridge
cycle) displacements (692). Variation in these parameters
thus does not explain isoform differences in thin filament
displacement velocity in Drosophila.
Multiple myosin heavy chain mutants have been isolated in Drosophila (180, 185, 616 – 618, 816, 819, 879, 880),
for some of which rescue and overexpression data are
available (223). Single amino acid mutations in the rod of
Drosophila indirect flight muscle myosin cause flightmuscle degeneration with age (even though myofibril assembly appears normal) (618), and mutations in the head
dramatically decrease thick filament number (617). Mutants that produce no myosin heavy chain produce relatively normal thin filament arrays (89, 180), and altering
actin-to-myosin ratios suggests that filament imbalances,
not lack of thick filaments per se, are responsible for
observed sarcomere structure defects (89). In support of
this, mutants that reduce thick filament number disproportionately affect flight and jump muscles (which have
highly organized sarcomeres) compared with intersegmental muscles (which have more variable myofilament
arrays) (879). Drosophila expressing only embryonic myosin heavy chain isoforms are viable but flightless with
normal-appearing muscles, and the flight muscles show
progressive deterioration (1272). Drosophila embryonic
myosin has lower actin sliding rates and basal ATPase
activity than adult indirect flight muscle myosin, and replacing Drosophila indirect flight muscle myosin with the
embryonic isoform slows muscle kinetics, which presumably underlies the flightlessness of such mutants (1127,
1129). The kinetic rate constants of the two isoforms
85 • JULY 2005 •
www.prv.org
1016
SCOTT L. HOOPER AND JEFFREY B. THUMA
show different dependencies on the presence of phosphate, suggesting that the difference between them is that
in the embryonic form the rate-limiting step of the crossbridge cycle is ADP release, whereas in the indirect flight
muscle form the rate-limiting step is phosphate release
(during the cross-bridge cycle, myosin head rotation is
associated first with phosphate and then ADP release, see
second review) (1131). Substituting only the embryonic
S1 exon (one of four in this region that differ) decreases
ATPase activity, but does not affect actin sliding velocity,
flight muscle ultrastructure, or flight ability (1128). However, switching of indirect flight muscle exons into the
embryonic isoform (when the embryonic isoform is the
one being expressed in the flight muscles), or embryonic
exons into the flight muscle isoform, alters wingbeat frequency and increases or decreases, respectively, the muscle’s maximum power generation and optimal wingbeat
frequency for power generation (1130). Unlike Drosophila, two myosin heavy isoforms are present in fleshfly
(Phormia) asynchronous and locust synchronous flight
muscle (1344).
2. Catchin/myorod
An alternative splice of the myosin heavy chain gene
codes for catchin/myorod (1075–1077, 1299), a protein
found in smooth muscle thick filaments in a variety of
bivalves. The protein contains the COOH-terminal rod of
myosin heavy chain but has a unique NH2-terminal head
(1076, 1299) and, although most highly expressed in catch
muscles (1299), is not necessary for catch (1298). Despite
the identity of the tail portions, pure catchin/myorod polymerizes very differently than myosin or myosin rod.
Adding myosin heavy chain to myorod preparations dramatically alters myorod polymerization (1075). A similar
isoform, myosin rod protein, is encoded in Drosophila by
a gene internal to the myosin heavy chain gene. Myosin
rod protein is present in a variety of larval, embryonic,
and adult cardiac, visceral, and somatic (including direct
flight) muscles and localizes with myosin in the A band
(950, 1114). Myosin rod protein may be associated with
the relatively disordered thin and thick filament packing,
variable thin-to-thick filament ratio, and bent thin filaments observed in some direct flight muscles (1114). In
muscles expressing high levels of myosin rod protein,
thick filaments and sarcomeres still form in animals lacking myosin heavy chain, suggesting that sarcomere formation does not require actin-myosin heavy chain interaction (950).
3. Myosin light chains
Vertebrate myosin possesses two light chains, the
essential and regulatory, in equimolar amounts (one each
per myosin molecule, and hence two each per myosin
heavy chain dimer). All known invertebrate myosins simPhysiol Rev • VOL
ilarly contain two essential (also called SH, catalytic,
myosin light chain 1, and alkali light chain) and two
regulatory (also called myosin light chain 2) light chains,
both of which have an ellipsoidal shape (Fig. 4) in a
variety of species (1111). [Caution: in early scallop papers, the regulatory light chain is also called the EDTA
chain because EDTA exposure removes one or both (depending on concentration and temperature) regulatory
light chains (32, 169, 553, 596, 1139), presumably due to
the enhanced regulatory light chain binding to the heavy
chain that Ca2⫹ and Mg2⫹ induce (31, 1139). EDTA sensitivity is not present in many other species (553, 654),
and the term is not used in modern parlance, but could be
confusing when reading the older literature.]
Whether all invertebrate myosins contain equimolar
amounts of the two light chains, however, is not as clear.
Early work in the scallop suggested that there were two
moles of essential light chain but only one of regulatory
light chain per myosin heavy chain dimer (1139). More
recent work, however, shows that scallop myosin does
contain two moles of each light chain per dimer (553,
1294). The situation in Crustacea remains unclear. Lobster and crayfish myosin contains two light chain types, ␣
and ␤, each with multiple variants, but which type is
essential and which is regulatory is unknown (570, 834,
984). In some lobster muscle fibers, ␣-chain immunohistochemical staining is higher in the center of the thick
filaments, whereas ␤-chain staining is always uniformly
distributed (133). Although these data are suggestive, they
do not prove nonuniform chain distribution because they
could result from nonuniform masking of the ␣-chain in
different thick filament regions. No difference is seen by
two-dimensional gel electrophoresis in lobster light
chains from slow versus fast muscle fibers (684), although
different isoforms are present in the two muscle types in
crayfish (1040). In earthworms, different light chain isoforms are expressed in a muscle-specific fashion, but
which are regulatory and which are essential is unknown
(160). Ascaris has two light chains in an approximate 1:1
molar ratio, but which are regulatory and which are essential is unknown (844). In vertebrates the light chains
are associated with the myosin heads near the neck region (Fig. 4A). Scallop light chains are similarly situated,
and the regulatory and essential chains extensively interact (34, 217, 309, 399, 598, 1112, 1140, 1217, 1219, 1244,
1245, 1247, 1283, 1294). Bivalve light-chain levels show
characteristic changes in response to various pollutants
(1000).
Essential light chains have been investigated in amphioxus (438), ascidia (1148), annelid (249a), several bivalves (70, 190, 375, 496, 723, 739, 740, 862, 940, 1139,
1246), abalone (32), squid (595, 596, 1258), and Drosophila (181, 288 –290, 665, 1153, 1155, 1156). In amphioxus,
scallop, and Drosophila only one gene was found. Early
work with gel electrophoresis showed that essential light
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
chains vary across bivalve species (740) but were identical in all muscle types in the one species (Spisula) in
which this question was examined (739). Later radioimmunoassay work using scallop antibodies showed that
the essential and regulatory light chains differ immunologically, that the antibodies did not cross-react with
other invertebrate essential light chains, and that striated
and smooth, but not heart, essential light chains were
immunologically identical (1246). Sequencing work
showed that several essential light-chain transcripts are
present in scallop (375), and identical isoforms are
present in striated and catch muscle (940) (cardiac muscle was not examined in this work). Essential light-chain
isoforms, generated by alternative splicing and protein
phosphorylation, are expressed in a muscle-specific fashion (indirect flight muscles vs. all others) in Drosophila
(288, 289, 1155, 1156, 1160). The splice patterns are maintained across several Drosophila species (665).
Regulatory light chains have been investigated in
amphioxus (1124); ascidia (1148); earthworm (228, 1072)
and the pogonophore Riftia (971); several bivalves (70,
375, 496, 592, 593, 597, 724, 739, 740, 800, 824, 862, 940,
1139, 1157, 1246), abalone (32), and squid (595, 596, 725);
C. elegans (227, 1029, 1030); Limulus (1248) and tarantula
(425); and Drosophila (924, 1153, 1155, 1156, 1178). Early
gel electrophoresis work showed that regulatory light
chains vary across bivalve species (740). Later radioimmunoassay work showed that scallop regulatory light
chain antibodies did not cross-react with other invertebrate regulatory light chains, and that striated and
smooth, but not heart, regulatory light chains are immunologically identical (1246). Bivalves (375, 940) and Drosophila (924, 1178) have one gene, and C. elegans has two
(227). Muscle, developmental stage, or gender specific
regulatory light-chain isoforms exist in earthworm (228),
C. elegans (1030), and Drosophila (1156, 1160, 1178).
Muscle specific isoforms are apparently not present in all
bivalves, as only one isoform is present in clam (Spisula
sachalinensis) adductor mantle, leg elevator, and heart
muscle (739), but two (phasic vs. catch muscle) in scallop
(592, 593, 824, 940). However, the different phasic and
catch muscle ATPase activities in bivalves are due to
differences in the myosin heavy chains, not the regulatory
light chains (940). Blocking scallop heavy chain thiol
groups blocks regulatory chain binding (594, 597, 896).
In C. elegans one gene is redundant, but deletion of
the other causes lethal pharyngeal muscle defects in hermaphrodites. This likely occurs because the first gene is X
linked, and its expression is downregulated (dosage compensated) in hermaphrodites (which have two X chromosomes). Consequently, without the second gene, in hermaphrodites insufficient regulatory light chain is produced in the pharynx (1030). The Drosophila isoforms are
believed to arise by posttranslational modification (1155,
1178), although several transcripts are expressed (924).
Physiol Rev • VOL
1017
Drosophila null mutation heterozygotes in the regulatory
light chain have reduced adult thoracic musculature,
somewhat disorganized indirect flight muscle, and altered
wing beat frequency spectrum or are flightless (464, 1251,
1252); changing a single amino acid from serine to alanine
reduces flight muscle power output (252). Drosophila
regulatory light-chain NH2-terminal deletions reduce indirect flight muscle calcium sensitivity, maximum amplitude of oscillatory work, sarcomere length, dynamic stiffness, and elastic modulus (476, 820). The stiffness and
elastic modulus changes suggest that in these muscles
regulatory light chain could link the thick and thin filaments.
4. Myosin light-chain phosphorylation
Expressed sequence tag analysis shows that a light
chain kinase is present in amphioxus notochord (1124). A
scallop cAMP-dependent kinase that phosphorylates only
one regulatory light chain isoform and a protein that
modulates the kinase have been identified (1098 –1100,
1102). A Limulus light chain specific kinase has been
purified (1071). The Drosophila light chain kinase gene
has been sequenced. Tissue- and stage-specific isoforms
are present; only some forms are Ca2⫹/calmodulin dependent (586, 1179). Reducing Drosophila regulatory lightchain phosphorylation reduces myosin ATPase activity
(1154) and, by reducing muscle stretch activation (a special property of the asynchronous muscles present in this
organism, see second review) (1110, 1180), indirect flight
muscle power output (252, 1110, 1180). Regulatory lightchain phosphorylation increases tarantula striated muscle
contraction; phosphorylation levels vary with Ca2⫹ concentration (425). Ascaris cuticle muscle regulatory light
chain is 48% phosphorylated in controls; GABA application reduces this phosphorylation to 18% (747). Phosphorylation of Limulus regulatory light chains in purified thick
filaments changes filament structure (674). Twitchin also
phosphorylates regulatory light chains (415). Light chain
phosphatases have been characterized in clam (1194) and
scallop; the scallop enzyme is Ca2⫹ dependent (474).
5. Paramyosin/miniparamyosin
We cover here only paramyosin genes and mutants
(see second review for structure of paramyosin containing thick filaments). Paramyosin/miniparamyosin have
been studied in cephalochordata (167, 437), echinodermata (876, 1195, 1256, 1284), annelids (1013, 1018, 1270,
1284), mollusks [Polyplacophora (786), bivalves (2, 56 –
60, 197, 198, 205–208, 270, 392, 397, 529, 547–549, 636, 641,
642, 677, 704, 738, 740, 766, 775, 786, 789, 795–797, 944,
991, 992, 1027, 1113, 1137, 1255–1257, 1270, 1284, 1290,
1291, 1311) (called tropomyosin A in Refs. 57, 60, 642, 795,
1027 and tropomyosin in Refs. 56, 58, 397, 547, 549, 641,
797), gastropods (401, 529, 786, 958, 1018, 1216, 1256,
85 • JULY 2005 •
www.prv.org
1018
SCOTT L. HOOPER AND JEFFREY B. THUMA
1291), cephalopods (56, 561, 619, 762, 1197) (called tropomyosin A in Ref. 762 and tropomyosin in Ref. 619)],
brachiopods (1284), Chaetognatha (1011), platyhelminths
(91, 156, 298, 359, 387, 444, 478, 521, 635– 637, 830, 1056,
1087, 1207, 1284, 1305), nematodes [C. elegans (24, 277,
346, 402, 514, 710, 939, 1059, 1133, 1256, 1262, 1263),
Ascaris lumbricoides (1284), Brugia malayi (648, 1274),
O. volvulus (231, 688), Dirofilaria immitis (376, 688)],
Limulus (248, 270, 468, 675, 676) (called tropomyosin A in
Ref. 468), mites/ticks (297, 967, 1021), Crustacea (202,
270, 677, 834, 836, 852, 1016, 1040, 1256, 1284), Diptera (D.
melanogaster) (27, 90, 696, 743, 1013, 1018, 1231, 1232),
Coleoptera (Heliocopris japetus, Pachnoda ephippiata)
(145), Heteroptera (Lethocerus) (141, 145, 270, 676, 677),
and Orthoptera (Locusta migratoria, Schistocerca gregaria) (93, 590, 1284). References 270, 528, 677, 1013,
1018, 1256, and 1284 are reviews. Paramyosin is observed
in amphioxus notochord (311, 437), catch (56, 60, 270,
397, 448, 766, 1027, 1137), smooth (24, 401, 876, 1013,
1018, 1027, 1195, 1284), obliquely striated (24, 396, 561,
1018, 1262, 1284), and cross-striated (93, 145, 166, 202,
248, 270, 468, 561, 675, 676, 852, 944, 1013, 1016, 1018,
1021, 1040, 1056, 1231, 1284) muscle and thus is not a
marker of catch ability or muscle type (except for existing
only in invertebrates). Paramyosin expression varies in
mollusk cross, obliquely, and nonstriated (smooth) muscles, although which types express more is inconsistent
across species (561, 1018).
Paramyosins from two bivalves, a gastropod, and a
polyplacophore all have similar molecular weights, electrophoretic properties, solubility, and guanidine-HCl denaturation and pepsin and trypsin susceptibilities (786).
Paramyosins from two smooth and one striated molluscan muscles, and four arthropod (Limulus, Homarus,
barnacle, insect) striated muscles, have similar molecular
weights and are immunologically similar (270). Heteroptera and Coleoptera indirect flight muscle paramyosins
also have similar intrinsic sedimentation rates and circular dichromism values (145). However, this does not mean
that all paramyosins are identical; paramyosin from different species (248, 528, 529, 766, 876, 1256) and multiple
isoforms within single species (2, 202, 693, 834 – 836, 852,
1040, 1056, 1106, 1231, 1243) show differences in antigenic
structure, molecular weight, isoelectric focusing, phosphorylation ability, or amino acid composition. The two
to three paramyosin isoforms present in decapod crustacea show muscle specific expression (834 – 836, 852,
1040).
The sources of this variation are not completely
known [for very early work, some may arise from proteolytic degradation during extraction (1311)]. Paramyosin is
phosphorylatable (2, 197, 249, 1056, 1059, 1101, 1256,
1257), which could explain differences in antigenic structure, isoelectric focusing, and phosphorylation ability
(since paramyosins with different phosphorylation states
Physiol Rev • VOL
in vivo would vary in their ability to be further phosphorylated in vitro). The amount of phosphorylation
observed depends on extraction procedure (197). In
mollusks Ca2⫹-, cAMP-, and calmodulin-dependent kinases phosphorylate paramyosin (1036), neuromodulators alter phosphorylation ability (2), and phosphorylation alters myofibril ATPase activity (1257), suggesting
that paramyosin phosphorylation could be functionally
relevant.
Paramyosin isoforms could also arise from multiple
genes. Mytilus has two paramyosin genes, but whether
one is a separate miniparamyosin gene is unknown
(1255). Two genes may be present in the nematode Anisakis (939), and the platyhelminth Schistosoma has more
than one (444). Paramyosin cDNAs or genes have been
cloned in the cestode Taenia solium (1207) (from the T.
solium gene a mini-paramyosin cannot expressed); nematodes Brugia (648), Onchocerca (231, 688), and Dirofilaria (376, 688); tick Boophilus (297); and mite Blomia,
(967) but whether these are single copy is unclear.
C. elegans [unc-15 (514, 993, 1263)] has a single
paramyosin gene but two paramyosin isoforms, which are
differentially located in the thick filaments, one being
associated with the filament core and the other with the
myosin heavy chains (240, 693). Reference 514 asserts
that paramyosin expression differs in body wall and pharyngeal muscles, but we have been unable to find sources
showing this difference. Regardless, paramyosin is expressed in all C. elegans muscle (24). C. elegans paramyosin mutants show varying degrees of locomotion impairment, and in the most extreme phenotype are paralyzed
as adults (1263). The body wall muscles of these mutants
show aberrant myofilament lattice and thick filament
structure (277, 1263). Changing a single charge on the
paramyosin can disrupt thick filament assembly (346),
and paramyosin appears to be required to determine thick
filament length (710) [as is also the case in Limulus (468)
and locust (590)]. Many C. elegans paramyosin mutants
that disrupt body wall muscle function do not alter pharyngeal muscle function, although the pharyngeal thick
filaments are missing their normal paramyosin core
(1263). This difference is believed to be because pharyngeal muscles have less paramyosin, use different myosin
heavy chains than body wall muscles, and have only one
sarcomere (514, 1263). Paramyosin degradation is increased in C. elegans mutants lacking myosin heavy chain
B, although synthesis is unaffected (1279).
Drosophila (1232) has a single paramyosin gene that
is required for proper myoblast fusion, myofibril assembly, and muscle contraction (696). Paramyosin is much
more highly expressed in tubular than in flight muscle in
Drosophila (1231). Comparing paramyosin content of
solid- versus open-cored thick filaments (see second and
third reviews) across several insects shows that paramyo-
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
sin is a larger percentage of filament mass in solid-cored
filaments (93).
Paramyosin is an important component in allergic
and immune reactions and may be useful in platyhelminth
and nematode vaccine development (3, 8, 13, 25, 83, 105–
108, 159, 171, 193, 201, 241, 245, 284, 307, 315, 344, 357,
358, 422, 495, 509, 521–523, 527, 583, 585, 635, 636, 644,
645, 648, 680, 681, 683, 685, 700, 726, 727, 777–780, 830,
846 – 848, 934 –936, 939, 965–969, 988 –990, 1078 –1080,
1117, 1165, 1172, 1173, 1187–1189, 1208, 1209, 1242, 1303–
1306, 1341). An important point for working on these
animals is that aldehyde fixation blocks muscle paramyosin staining in them (637).
Miniparamyosin is present in Drosophila (90, 744,
1013), shrimp (1016), and mite striated (1021) muscle and
Eisenia obliquely striated and smooth (1013) muscle.
Whole animal extracts show it to be present in annelid,
snail, mussel, and multiple arthropod species, but not C.
elegans, a result supported by analysis of the C. elegans
paramyosin gene (744). Drosophila mini-paramyosin is
encoded by the same gene as paramyosin, but its expression is transcriptionally regulated from a separate promoter (27, 90, 744). Drosophila miniparamyosin is expressed only in adults and has several isoforms (with
much greater molecular weight variation than those
known for paramyosin) that are expressed in a muscle
specific pattern (90, 743). The sequences that regulate
miniparamyosin/paramyosin expression are beginning to
be described (27). Miniparamyosin overexpression results
in nearly normal thick filaments and sarcomere length.
However, some defects are present in indirect flight muscle with some flight impairment, and these defects increase with age (28).
6. Filagenins
C. elegans thick filaments are associated with three
30-, 28-, and 20-kDa proteins (␣-, ␤-, and ␥-filagenin, respectively) (240), whose cDNAs have been sequenced
(694, 695). The proteins are believed to form the thick
filament core by cross-linking paramyosin subfilaments
(693). The ␣-form is located with myosin heavy chain A in
the middle of the thick filaments while the ␥-form is
located more laterally with myosin heavy chain B (695).
The ␣- and ␥-forms appear at different developmental
stages (695). The ␤-form is not present in pharyngeal
muscles (694).
7. C. elegans unc-45
Temperature-sensitive unc-45 mutants have reduced
body wall muscle thick filament number and are paralyzed as adults (279); homozygous, non-temperature-sensitive mutations result in arrested muscle development
and are lethal (1213). The gene has been sequenced and is
expressed in all C. elegans muscles, and only in muscle
Physiol Rev • VOL
1019
(1212). The protein associates with myosin heavy chain B
in the thick filament. In myosin heavy chain B null mutants unc-45 muscle localization is not seen (and thick
filament structure is near normal) (22). unc-45 binds
myosin heads (both scallop and C. elegans, although unfortunately myosin heavy chain A and B were not separately tested) (72). Temperature-sensitive mutations in a
region of unc-45 homologous to two fungal proteins involved in protein segregation result in (at the restrictive
temperature) myosin heavy chain A and B being scrambled in the thick filaments (as opposed to being located in
distinct thick filament regions as in wild type) (71). This
work suggests that unc-45 is a chaperone that helps
correctly integrate myosin heavy chain B into the thick
filament (693, 1318).
8. Myonin
A 230-kDa thick filament associated protein named
myonin that appears to regulate myosin heavy chain
ATPase activity has been identified in clam (1310).
9. Flightin
Flightin is a 20-kDa protein that has been found only
in Drosophila (1225, 1229), but efforts to identify it in
other species have not been reported. It is found only in
indirect flight muscles and is localized in the A band
except for the H-band region (Fig. 8) (1229). The protein
has multiple phosphorylation sites, and its phosphorylation state increases with age, particularly after eclosion
(1227). The single amino acid myosin gene mutations
noted above that cause age-dependent myofibril degeneration also fail to accumulate phosphorylated forms of
flightin (618). Mutations that block thick filament assembly block all flightin phosphorylation, whereas mutations
that block thin filament assembly alter flightin phosphorylation temporal sequence (1224). Flightin null mutants
are viable but flightless (976). Mutant sarcomeres appear
relatively normal in pupa and early adult but are ⬃25%
longer than normal. When the animals attempt to fly, the
indirect flight muscles become disrupted, and eventually
site-specific thick filament cleavage occurs. Flightin binds
to the myosin rod, and the flightin:myosin molar ratio is
between 1:1 and 1:2 (44).
10. Zeelin
Zeelin 1 and 2 have been isolated from Lethocerus
muscles (294, 1037). Zeelin 1 is present in flight and leg
muscles, and zeelin 2 is only in flight muscles. In flight
muscles the zeelins are present in discrete A-band regions, whereas in leg muscles the entire A band is labeled
(the identification in Ref. 1037 that these proteins were in
the Z disk was an artifact of fiber glycerination) (294).
Both proteins are associated with the thick filament, with
85 • JULY 2005 •
www.prv.org
1020
SCOTT L. HOOPER AND JEFFREY B. THUMA
zeelin 1 closer to the filament shaft than zeelin 2, and ⬃0.5
mol zeelin/ mol myosin (294). Zeelin function is unknown.
The small size and location of the zeelins are similar to
Drosophila flightins, but antibodies to zeelin do not recognize flightin (294).
protein) 43-kDa Drosophila Z-disk associated protein has
been identified, but its structure and function are unknown (998).
C. ␣-Actinin and Other Z Line Proteins
A wide variety of Ca2⫹ binding proteins (calmodulin,
Ca -dependent Ca2⫹ binding protein, troponin C, the B
subunit of phosphatase 2B, myosin light chains, parvalbumins) have evolved from a common ancestor with four
Ca2⫹ binding domains (see Refs. 574, 1144 for vertebrate
references, Refs. 610, 611, 818 for extensive summaries of
Ca2⫹ binding proteins and their evolutionary relationships, and Ref. 746 for a comparison of Ca2⫹ binding site
amino acid complements). These proteins are divided into
two types (209, 210). Proteins of the first (e.g., vertebrate
parvalbumin) are present in high quantity, have high Ca2⫹
affinity, have Ca2⫹ binding regions with a stable configuration, do not interact with known targets, are believed to
function as Ca2⫹ buffers and to help resequester Ca2⫹
after muscle contraction or to protect against deleterious
effects of high Ca2⫹ concentrations during prolonged contractions, and show relatively rapid evolutionary divergence (reviewed in Ref. 421). Proteins of the second are
present in lower quantity, have lower Ca2⫹ affinity, have
Ca2⫹ binding regions whose conformation changes upon
Ca2⫹ binding, are believed to be more directly involved in
excitation-contraction coupling (examples include troponin C and calmodulin), and are relatively well conserved.
We cover here only muscle specific Ca2⫹ binding proteins
(in particular, calmodulin, which is present in a wide
variety of tissues, is not included). Reference 348 reviews
Ca2⫹ binding proteins in general.
Amphioxus (Branchiostoma lanceolatum) has seven
sarcoplasmic Ca2⫹-binding proteins of the first type, all
arising by alternative splicing from one gene. One isoform
has been crystallized and its structure determined at 2.4-Å
resolution (194, 196). The two isoforms for which it has
been examined bind three Ca2⫹ (582, 1141, 1146, 1150).
Immunohistochemistry shows the proteins are localized
at the Z line (1202). One protein of the second type, Ca2⫹
vector protein, has been characterized in amphioxus and
binds two Ca2⫹ (62, 63, 209, 211, 574, 1124). The molecule’s three-dimensional solution structure, backbone dynamics, and Ca2⫹ binding properties have been studied in
great detail (62, 63, 1169 –1171). Its cDNA has been sequenced (1320), and although clearly a Ca2⫹ binding protein, no homolog has been found in invertebrates or vertebrates (1202). The protein is present in a wide variety of
tissues, but in highest concentration (50 –100 ␮M) in muscle (209, 412). Immunohistochemistry shows it to be
strongly localized at the Z lines and weakly localized at
the M lines, and it is more difficult to extract from muscle
than the sarcoplasmic Ca2⫹-binding proteins (1202). Ca2⫹
D. Ca2ⴙ Binding Proteins and Their Targets
2⫹
␣-Actinin has been isolated or identified by immunohistochemistry in annelid (Eisenia) obliquely striated and
smooth muscle (1010); bivalve (scallop) cross-striated,
obliquely striated, and smooth muscle (559, 1074); gastropod (Helix) cross-striated (heart) and smooth muscle
(1010); cephalopod (squid) obliquely striated muscle
(1196); chaetognath (Sagitta friderici) body wall muscle;
platyhelminth (Schistosoma mansoni) muscle of unidentified type (706); C. elegans obliquely striated muscle
(316), mite cross-striated muscle (1021); and Drosophila
(1010, 1035, 1228), honeybee (Apis mellifera) (1034), and
giant water bug (Lethocerus cordofanus) cross-striated
(146, 393) muscle. In insects, ␣-actinin staining was seen
throughout the Z line (1010). In the obliquely striated
muscles of earthworm, snail, and C. elegans, in which thin
filaments attach to peglike dense bodies analogous to
cross-striated Z lines (see third review), the staining was
limited to these formations (316, 1010). In earthworm and
snail smooth muscle, the staining was limited to discontinuous patches, suggesting that in these muscles the thin
filaments also have discontinuously distributed anchorage sites (1010).
Drosophila has a single ␣-actinin gene that gives rise
to three isoforms (338, 1008). Two are muscle specific.
One is present in larval and adult head and abdominal
muscles and the other in adult indirect flight muscles and
tubular jump and leg muscles (1008, 1035, 1228). Mutations can be fatal or cause severe muscle defects (327,
338, 441– 443). Rescue experiments show that 1) the indirect/jump/leg muscle isoform can replace the cytoplasmic and larval/head/abdominal forms and 2) increasing
␣-actinin’s length 15% does not affect its function (264).
␣-Actinin and kettin (see below) bind actin simultaneously and well (1205). A C. elegans ␣-actinin cDNA
clone has been isolated (73).
In C. elegans a small heat shock protein, HSP25, that
binds ␣-actinin and vinculin but not actin, is found in
muscle-dense bodies and M lines (253). Early immunohistochemistry in Drosophila and honeybee identified small
(150 –200 kDa) (some present in only very restricted sets
of muscles) and larger (400, 600 kDa) (of which the
600-kDa band showed some muscle specificity) non-actinin, non-projectin Z-disk proteins (1034, 1035, 1228). An
apparently novel (the 10 amino acid sequence against
which the monoclonal antibody used to identify it was
raised showed no close match to any known Drosophila
Physiol Rev • VOL
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
vector protein forms a high-affinity, Ca2⫹-strengthened,
complex with Ca2⫹ vector target protein (209, 942). Ca2⫹
vector target protein is also present in both muscle and
nonmuscle tissues, but equally stains the Z and M lines
(1202). Ca2⫹ vector target protein does not appear to have
any enzymatic role, and its function is unknown. However, it has two immunoglobulin domains and thus could,
as do other immunoglobulin containing muscle proteins
(see sect. IIIE), contribute to sarcomere structure (1142).
cAMP-dependent protein kinase phosphorylates four
Ca2⫹ vector target protein sites, and phosphorylation decreases Ca2⫹ vector protein and Ca2⫹ vector target protein affinity (943).
Low-molecular-weight muscle-associated Ca2⫹ binding
proteins are present in several other invertebrates. Nereis diversicolor and Perinereis vancaurica tetradentata (annelid) sarcoplasmic Ca2⫹ binding proteins bind
three Ca2⫹, but that of Nereis virens only binds two.
The N. diversicolor and Perinereis proteins have been
sequenced (189, 213, 242, 577). The Nereis protein’s
three-dimensional structure (54, 195, 1230) and dynamics of Ca2⫹ and Mg2⫹ binding (274) have been determined, and the conformational changes associated with
Ca2⫹ binding are being intensively studied (178, 216,
266, 705, 953, 1094). Earthworm has three Ca2⫹ binding
protein isoforms that bind either two or three Ca2⫹, and
which are present in higher concentrations in fast relaxing muscles (consistent with their being involved in
Ca2⫹ buffering) (459, 460). The 20-kDa Ca2⫹ binding
proteins have been isolated from two scallops. The
Patinopecten protein has been sequenced and likely
binds two Ca2⫹ (191, 1143). Clams and oysters have
three Ca2⫹ binding protein isoforms (1141). Low (17
kDa)- and high (450 kDa)-molecular-weight Ca2⫹ binding proteins and a calsequestrin (the major Ca2⫹ binding protein in vertebrate skeletal and cardiac muscle)
homologs are present in the membrane fraction of
Mytilus anterior byssal retractor muscle (1301). A Ca2⫹
binding protein uniquely expressed in muscle that is
closely associated with the contractile machinery, and
which cross-reacts to antibodies to amphioxus sarcoplasmic Ca2⫹ binding protein, has been identified in
Aplysia, as has another present in both muscle and
neurons (932). A 20-kDa, muscle- and tegument-associated, Schistosoma Ca2⫹ binding protein has been isolated and sequenced (408, 409, 1118), and a gene identified that codes an 8-kDa Ca2⫹ binding protein of unknown location, but which is only expressed during the
animal’s free swimming stage (964). C. elegans calsequestrin is expressed only in body wall muscle. Its
function is unknown, but the sequences regulating its
expression are beginning to be identified (177).
Unlike other known invertebrate Ca2⫹ binding proteins, crayfish (102, 214, 498, 1287–1289), lobster (1289),
and shrimp (1145, 1147, 1289) sarcoplasmic Ca2⫹ binding
Physiol Rev • VOL
1021
proteins are dimers of two nonidentical proteins, each of
which binds three (for a total of 6 per dimer) Ca2⫹; ␣␣,
␣␤, and ␤␤ are all present. Ca2⫹ binding induces large
conformational changes. The amino acid sequences of the
crayfish (498) and shrimp proteins are known (1145,
1147), and the crayfish protein has been crystallized (612).
Additional Ca2⫹ binding proteins have been isolated from
crayfish and lobster muscle (420, 759), one of which has
sequence similarity to calcyphosine, a vertebrate nonmuscle Ca2⫹ binding protein (1051). Although early work
was unable to find sarcoplasmic Ca2⫹ binding proteins in
locust (212), 16- to 20-kDa Ca2⫹ binding proteins with
muscle specific (tubular vs. indirect) expression were
identified in Drosophila (46, 1160). The 23- to 24-kDa Ca2⫹
binding proteins were subsequently isolated from Drosophila and Calliphora (551, 560). The cDNA for the
Drosophila protein has been cloned and is expressed in
tubular but not indirect flight muscle (551). The gene is
not single copy (551), and multiple isoforms are observed
(560). A Drosophila calbindin homolog that is primarily
expressed in neurons but is present in a few muscles has
also been identified (986).
E. Giant Sarcomere-Associated Proteins
Like vertebrates, invertebrates also have large sarcomere proteins involved in centering the thick filaments in
the sarcomere and generating muscle stiffness. Invertebrate muscles, however, show a number of variabilities
not present in vertebrate sarcomeres (144, 323, 621). First,
invertebrate sarcomeres cover a very wide length range
(crayfish giant sarcomeres are ⬃8 ␮m at rest whereas
those of Drosophila indirect flight muscles are 3.3 ␮m).
Second, invertebrate muscles have very different abilities
to be stretched without damage (crayfish giant sarcomeres can stretch to 13 ␮m, Drosophila indirect flight
muscle sarcomeres to only ⬃3.5 ␮m). Third, invertebrate
muscles show a very wide passive stiffness range (a tension of 8 mN/mm2 stretches crayfish sarcomeres 4 ␮m,
but Drosophila sarcomeres only 0.2 ␮m).
Given these differences, invertebrates might be expected to have a wide variety of giant muscle proteins,
and indeed, a large number were early identified on the
basis of molecular weight, sarcomere position, and antibody reactivity (316, 455, 456, 544, 545, 639, 698, 730 –732,
757, 845, 850, 851, 884, 1014, 1016, 1021, 1033–1035, 1037,
1218, 1220, 1221, 1228, 1343). It is now clear that these
proteins belong to multiple isoform families of wide size
range, that gel electrophoresis molecular weight estimates for very large proteins can be highly inaccurate
(323), and that many of these proteins, even those from
different families, are sufficiently similar to immunologically cross-react. Gene sequencing in C. elegans and Drosophila has begun to provide order to this field. However,
85 • JULY 2005 •
www.prv.org
1022
SCOTT L. HOOPER AND JEFFREY B. THUMA
in older literature terms such as “mini-titin,” “titin/
twitchin,” “titinlike,” etc., are widespread, and what protein is being described is not always clear. Even worse,
vertebrate titin is also called connectin, and some authors
refer to all large (⬎2 MDa) invertebrate sarcomere proteins as connectin (756), even though sequence data show
that some of these large proteins are very different. Moreover, in an extremely unfortunate happenstance, a Drosophila cell adhesion molecule present on embryonic
muscles is called connectin (870 – 872). Readers must
therefore exercise caution in the older literature. Reviews
covering invertebrate sarcomere giant proteins (at least in
part) include References 95, 98, 142, 144, 544, 756, 758,
1184, 1185, 1190, 1343, 1345.
To provide order to these proteins, it is helpful to first
examine human skeletal muscle titin (Figs. 7A and 8) (182).
The protein’s first 80 kDa, with its multiple immunoglobulin
domain repeats, binds to the Z disk and possibly actin.
Additional immunoglobulin repeats (which can provide elasticity by unfolding under tension, Ref. 649) and an elastic
PEVK region occur next and span the adjacent half I band in
which the molecule moves toward the thick filaments. Next
is a large region of a repeated immunoglobulin/fibronectin
domain motif, which is associated with the thick filaments in the A band. The molecule ends with a 250-kDa
M-line region that contains a serine/threonine kinase
activity and multiple immunoglobulin repeats. Thus, in
addition to being huge (one basis for calling various
invertebrate giant proteins titin-like), titin also has regions in which only or primarily immunoglobulin repeats exist, areas expected to act as springs, a region
with a repeated immunoglobulin/fibronectin motif, and
a kinase (and hence a wide range of epitopes against
which antibodies can be raised). A large variety of
proteins could thus resemble one or another portion of
titin but still be distinctly different from one another.
Titin is believed to fulfill four functions (379, 755, 1186).
First, its springlike nature helps keep the thick filaments centered in the sarcomere. Second, this same
nature provides much of muscle passive resistance to
stretch and elasticity. Third, it may help determine
sarcomere length in development. Fourth, although its
FIG. 7. The motif structure of human titin and the major invertebrate giant sarcomere-associated proteins. Red rectangles, immunoglobulin
domains; green rectangles, fibronectin domains; blue rectangles, SEK domains; yellow rectangles, serine/threonine kinase activity; spring, PEVK
domain. SEK and PEVK domains are so named (single letter amino acid code) because these are the most prevalent amino acids in these regions.
In all cases the NH2 terminus is to the left. The sequences in B are aligned at a conserved site (vertical double headed arrow) where a linker region
between two immunoglobulin domains is missing. Data are from References 97, 99, 144, 306, 323, 587, 708, 906, 1097, 1107.
Physiol Rev • VOL
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
1023
FIG. 8. Sarcomere location of the major
invertebrate giant muscle proteins. Top: a repeat of the vertebrate sarcomere shown in Fig.
5. Bottom: position of vertebrate titin and major
invertebrate giant proteins. Triangles represent
serine/threonine kinase position when known.
Line lengths are not proportional to protein molecular weight.
target(s) is unknown, titin’s kinase could regulate the
activity of other sarcomere proteins.
Except for amphioxus (455) and ascidia (845), in
which the proteins are presumably vertebrate titin, no
known invertebrate giant protein is long enough to
span an entire half sarcomere (566). Proteins in the 3
MDa range (similar to titin’s molecular weight) have
been identified by gel electrophoresis in some (but are
clearly absent from other) crayfish muscles (732) and in
the 5 MDa range in barnacle (730), but these lengths are
insufficient to span half the large (10 ␮m) sarcomeres
present in them. Very large (4-MDa) proteins that connect the Z line to the thick filaments are present in the
annelid Neanthes, but how far along them the proteins
extend is unknown (545). All other invertebrate giant
proteins are also too small to span a half sarcomere.
Present best knowledge is thus that invertebrate elasticity and sarcomere structure arise from proteins that
connect Z line and thick filament proteins, but which do
not extend all the way to the M line.
Physiol Rev • VOL
1. Crayfish connectin/kettin, C. elegans kettin,
and Drosophila titin/kettin
Genes capable of producing 1) ⬃2-MDa proteins
(called connectin in crayfish and titin in Drosophila) with
an immunoglobulin repeat-rich region and springlike
PEVK regions but no protein kinase domain, and 2) ⬃550kDa proteins (called kettin in all species) consisting of
only immunoglobulin domains, have been completely or
partially sequenced in C. elegans (144, 391), Drosophila
(144, 587, 640, 708, 1337), Bombyx mori (584, 1123), and
crayfish (323). The reported Drosophila sequences show
some variation. The sequence predicted from the Drosophila genome project (144) lacks eight NH2-terminal
immunoglobulin repeats (left rectangle, below sequence)
present in the sequence of Reference 708, and has four
COOH-terminal fibronectin repeats (right rectangle) not
present in Reference 708. Drosophila and crayfish kettin
are produced by alternative splicing from the titin/connectin gene. C. elegans kettin is instead produced by its
own gene (see below for a large C. elegans sarcomere
85 • JULY 2005 •
www.prv.org
1024
SCOTT L. HOOPER AND JEFFREY B. THUMA
protein). In Bombyx, kettin is also produced by its own,
single-copy gene, and there are two titinlike genes. Given
the presence in C. elegans (see below) of another giant
sarcomere protein that is distantly, if at all, related to the
crayfish connectin/insect titin/C. elegans-Drosophilacrayfish kettin protein family, but which is nonetheless
named titin, the nomenclature in this field clearly needs
revision. Reference 144 suggests that the crayfish connectin/insect titin/C. elegans-Drosophila-crayfish kettin proteins be all renamed SLS proteins (from sallimus, the
gene that codes for Drosophila titin/kettin), with kettin
becoming just an SLS isoform.
Crayfish connectin runs from the Z line to the tips of
the thick filaments and stretches with the sarcomere to
lengths up to 3.5 ␮m (323). Crayfish connectin contains
both a PEVK region and a SEK region, which is also
elastic (322), and connectin proteolysis in skinned crayfish fibers abolishes resting tension (732). Antibodies to
an epitope on crayfish connectin that is not present in
Drosophila titin recognize a large protein in barnacle
ventral muscle (323), Calliphora larval muscle (323), and
beetle (Allomyrina), bumblebee (Bombus), and waterbug
(Lethocerus) leg muscle (323, 884). The beetle, bumblebee, and waterbug protein is recognized by antibodies to
vertebrate titin and is localized in the I band in beetle
(884). The crayfish connectin antibody also recognized a
3-MDa protein in waterbug, but not beetle or bumblebee,
flight muscle (884). A 540-kDa protein that localized to Z
lines, elongated upon muscle stretch, bound actin, and
was recognized by antibodies against vertebrate titin, and
a 500-kDa Lethocerus muscle protein was early isolated
from crayfish claw muscle (731). Given the cloning evidence for kettin in crayfish (323), this protein is presumably kettin.
C. elegans kettin has an immunological staining pattern consistent with dense body (equivalent to Z line)
expression, all muscles were stained (144, 587). Drosophila kettin is present in all larval body wall and visceral
muscles and in adult leg and flight muscles. Mutants of
what was believed to be the kettin gene have abnormal
muscles; homozygotes cannot hatch and heterozygotes
cannot fly (391). However, in light of titin and kettin
arising from one gene, it is now unclear if these effects are
due to a lack of kettin or titin. Kettin is located at the Z
line in flight muscle, with the NH2 terminus in the line and
the COOH terminus outside (140). Kettin binds to actin
filaments (1205) and in vitro promotes anti-parallel association of thin filaments (the same arrangement as that
present in vivo, see Fig. 5) (1205). Kettin’s immunoglobulin domains bind ␣-actinin and actin (640) and are consistent with a model in which each repeat and its flanking
region bind to one thin filament actin monomer, each
immunoglobulin/linker binding ⬃3 nm of the thin filament
(587, 1205). With this model, each sarcomere’s kettin
molecules would overlap by a small amount and extend
Physiol Rev • VOL
some 60 nm outside the 80- to 120-nm Z-line width (587).
Kettin competes with tropomyosin for actin binding,
which could exclude tropomyosin from the Z line (1205).
Kettin’s COOH terminus can bind myosin; kettin could
thus link the Z line to the thick filaments (621) and is
therefore believed to play a major role in muscle stiffness
(140, 621). However, it also has large numbers of immunoglobulin domains, and single kettin molecules respond
to tension with steplike relaxations presumably stemming
from unfolding of individual immunoglobulin domains
(649). Kettin may therefore also play a role in muscle
elasticity.
Drosophila titin is required for myoblast fusion and
the formation of striated muscle sarcomeres (708, 1337).
Titin also plays an important role in muscle elasticity.
Although immunoglobulin/fibronectin unfolding presumably can play a role in this process, it is not its only
source, since muscle length changes in 2.3 nm quanta
when force is applied to indirect flight muscles under
conditions in which titin is the sole length-absorbing element. This distance is much less than that which would
occur from immunoglobulin or fibronectin domain unfolding and may correspond to single beta-sheet unfolding
(123). Multiple titin isoforms (in addition to kettin) are
expressed with some muscle specificity (144). Drosophila
titin is also present on condensed chromosomes and required for proper chromosome structure and function
(707, 708). A titin-like protein is present in Anopheles
(144).
2. C. elegans titin
The C. elegans titin gene produces a 2,200-kDa protein with multiple immunoglobulin and fibronectin repeats, an elastic region, and a kinase; a 1,200-kDa protein
with the immunoglobulin and fibronectin repeats and
elastic region, but no kinase; and a 301-kDa protein consisting of only the kinase and the COOH-terminal (primarily) immunoglobulin repeats (306). Immunohistochemistry shows that the 2,200- and 1,200-kDa isoforms are
expressed in all muscles except in the pharynx. The 2,200and 1,200-kDa proteins bind at their NH2 terminal to the
dense bodies (Z-line equivalents). The 2,200-kDa protein
may be long enough to reach slightly into the A band, and
thus its kinase could act on thick filament proteins. The
position of the 301-kDa protein is unknown. The gene is
not expressed until after C. elegans muscles are formed.
No C. elegans titin mutants are available, and thus the
functions of these proteins are unknown, as are the kinase’s targets. A survey of the C. elegans genome searching for immunoglobulin containing proteins identified
twitchin, unc-89, unc-52, and several titinlike sequences
(1167).
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
3. Unidentified very large sarcomere proteins
Barnacle has 5,000- and 1,200-kDa proteins that
monoclonal antibodies to vertebrate titin recognize; the
5,000-kDa protein links the thick filament to the Z line
(730). This could be a crustacean equivalent to C. elegans
titin, but sequence data will be required to confirm this
relationship. Antibodies to vertebrate titin recognize a ⬎2
MDa scallop protein and label scallop thick filaments
along their entire length and most strongly at their tips
(1221). A similarly large titinlike (by antibody binding)
protein is present in squid (542). These data indicate that
mollusks contain a very large sarcomere protein, but its
relationship to other invertebrate muscle proteins is unknown.
4. Projectin/twitchin
A family of 600- to 1,000-kDa proteins with repeating
fibronectin/immunoglobulin repeats and a serine/threonine kinase domain exists in mollusks, C. elegans, crustaceans, and insects. The protein is called twitchin in C.
elegans, twitchin or twitchinlike in mollusks, and projectin in crustaceans and insects.
Twitchin (unc-22) was identified from C. elegans
mutants with a constant twitch and abnormal body wall
muscle sarcomeres (131, 678, 679, 809, 811, 813, 1003).
Twitchin consists of several immunoglobulin repeats,
multiple repeats of an immunoglobulin/fibronectin motif,
a serine/threonine kinase, and several immunoglobulin
repeats (96, 97, 314, 810). Up to six fibronectin/immunoglobulin repeats can be deleted without strong phenotypic effects (562). Bacterially expressed twitchin kinase
autophosphorylates (664). The 60 residues COOH terminal to the kinase core inhibit kinase activity (664) by
binding to the active site (intrasteric regulation) (457, 458,
578, 579). The Ca2⫹ binding protein S100A1 relieves the
inhibition and increases kinase activity 10-fold (413, 507).
Although several other Ca2⫹ binding proteins, including
calmodulin, bind to the autoinhibitory sequence (417),
they do not relieve the inhibition (664). Twitchin staining
is observed at both the A band and the dense body,
suggesting it links the thick filaments to the dense body
(767, 810). There are likely at least two twitchin genes, as
unc-22 mutants remove staining from body wall, anal, and
vulval muscles, but not pharynx (810). Mutations in unc54, which codes for myosin heavy chain B, can suppress
the unc-22 phenotype; the unc-54 mutants in isolation
have normal muscle structure, but the animals are stiff
and slow (812).
Striated and smooth scallop muscle and mussel
smooth catch muscle contains a 600 – 800 kDa (0.2 ␮m)
rod-shaped protein with a kinase domain (1220). Antibodies against the protein stain Limulus, Lethocerus (indirect flight), crayfish, and chicken muscle and bind to the
A/I junction in all except Lethocerus, in which they bind
Physiol Rev • VOL
1025
to the I band. Antibodies raised to the immunoglobulin
repeat or kinase domains of C. elegans twitchin recognize
the protein. Although both C. elegans twitchin and titin
have immunoglobulin repeats and a kinase, the protein’s
molecular weight suggests it is a projectin/twitchin.
Immunohistochemistry, immunoelectron microscopy,
and partial amino acid and cDNA sequencing indicate that
Aplysia californica has a twitchin-related protein that colocalizes with contractile filaments (416, 822, 954). Like C.
elegans twitchin, Aplysia twitchin autophosphorylates (416,
822), has an autoinhibitory sequence that binds Ca2⫹/calmodulin without activating the kinase (137, 413, 414, 416,
417), and is activated (in this case, ⬎1,000-fold) by Ca2⫹/
S100A1 binding (413, 414, 417). However, Aplysia twitchin
phosphorylates regulatory myosin light chains well (415),
whereas C. elegans twitchin only poorly phosphorylates
peptides containing the expected nematode regulatory myosin light-chain phosphorylation site (98, 664). C. elegans
and Aplysia twitchin also differ with respect to Ca2⫹/calmodulin affinity, sensitivity to naphthalene sulfonamide inhibitors, and the effect of Zn2⫹ on Ca2⫹/S100 activation
(required in Aplysia, inhibitory in C. elegans) (98, 414, 417).
Studies on neuropeptide modulation of Aplysia muscle relaxation indicate that cAMP-dependent protein kinase phosphorylates Aplysia twitchin, suggesting that twitchin may
regulate muscle relaxation (954), consistent with unc-22’s
phenotypic effects (809). Aplysia twitchin’s sarcomere location is unknown. Phosphorylation of a Mytilus twitchin-like
protein is involved in catch (see second review) (151, 152,
324, 326, 1088, 1089, 1298).
Early work with monoclonal antibodies against
Lethocerus muscle proteins revealed an 800-kDa protein
(639). A Drosophila protein, projectin, that is believed to
be the homolog of this Lethocerus protein, has been sequenced (47– 49, 232, 329, 1107). Projectin is very similar
to C. elegans twitchin (49), and projectin antibodies
cross-react with C. elegans twitchin (1228). Several isoforms, all of which retain the kinase and which show
muscle specific expression (1082, 1228), arise by alternative splicing (48, 232, 1107). The bent locus codes Drosophila projectin. Homozygous embryos do not hatch,
and in the most severe allele no muscle contraction ever
occurs (48, 329). Heterozygote muscles show reduced
stretch activation (821, 1226). Projectin, paramyosin, and
myosin aggregates form different diameter fibers depending on projectin’s phosphorylation state (287). In locust,
projectin (along with paramyosin) may help determine
thick filament length (590). Drosophila projectin binds to
thin filaments (1271), and locust projectin blocks the
formation in vitro of paracrystals from solutions of actin
monomers, instead promoting thin filament formation
(949). Locust projectin autophosphorylates (287, 1271)
and phosphorylates 30-, 100-, 165-, and 200-kDa proteins
(but not myosin light chain) in vitro (287, 1271). The
30-kDa protein may be troponin I, and troponin phosphor-
85 • JULY 2005 •
www.prv.org
1026
SCOTT L. HOOPER AND JEFFREY B. THUMA
ylation by projectin increases troponin’s Ca2⫹ sensitivity
(1271). Actomyosin, actin filaments, and myosin inhibit
projectin autophosphorylation, and calmodulin stimulates
it (287). Projectin phosphorylation of other proteins, however, is unaffected by actomyosin, actin filaments, myosin, calmodulin, Ca2⫹, Ca2⫹ and calmodulin, or cAMP
(287, 1271). Calmodulin-dependent protein kinase II
(CaMKII) binds to projectin, which increases CaMKII activity twofold (286).
In synchronous muscle, projectin is located in the A
band similar to twitchin (47, 639, 1035, 1228). In locust,
the molecule extends into the I band, and the molecule’s
COOH terminus, with the kinase domain, is located in the
A band (1082). In asynchronous muscle, it is instead
located in the I band and forms filaments connecting the
Z line and thick filament (47, 639, 1033, 1035, 1228). Neither myosin nor actin is required for projectin’s initial
localization to the Z line during sarcomere assembly, but
in the absence of actin projectin’s Z line localization is
later lost (45). Proteins believed to be projectin have been
isolated from locust, bee, beetle, and Lethocerus (142,
884, 1034), and staining with an antibody likely to have
been raised against Drosophila projectin stains (as expected) Drosophila cross-striated muscle and annelid
(Eisenia) obliquely striated and smooth muscle (1014).
A large sarcomere protein with immunological similarity to vertebrate titin and honeybee, beetle, and Drosophila projectin that contains a serine/threonine kinase
and autophosphorylates was early isolated from crayfish
(456, 745). This protein, crayfish projectin, has been
cloned, and flexor and closer muscles express different
isoforms (906).
5. C. elegans unc-89
The unc-89 mutants have disorganized muscle structure and lack M lines (95, 1266). The unc-89 gene has
been cloned and is believed to be a homolog of vertebrate
obscurin (98, 99, 1097, 1122). The gene produces four
isoforms that are differentially expressed in different
muscles. Two of the isoforms contain extensive immunoglobulin repeats. One of these isoforms also contains two
protein kinase domains, although only one is believed to
be functional. The other two isoforms are short versions
containing only the tandem protein kinase domains. All
the isoforms are localized in the center of the A band. The
structure of a region of the molecule that may be involved
in protein-protein interactions has been determined in
detail (120, 121). unc-98, a protein containing four Zn
finger motifs, is also located at the M line in C. elegans
(787). A 400-kDa M-line protein has been identified in
Lethocerus (142, 639), but without sequence data the
relationship of this protein and the unc-89 product is
unknown.
Physiol Rev • VOL
F. Miscellaneous Other Proteins
1. Nebulin
Nebulin-like proteins are present in obliquely striated
Eisenia muscle (1015) and cross-striated muscle in Antarctic mussel shrimp (1016), Antarctic mite Halacarellus
(1021), lancelet Branchiostoma (cephalochordate) (313),
and hagfish skeletal and heart muscle (312). Nebulin-sized
proteins are present in echinoderms, an annelid, mollusks, crustaceans, and insects (698). In Eisenia and Halacarellus, anti-nebulin staining occurs along the entire
sarcomere, whereas in Branchiostoma staining is only in
the I band.
2. Nesprin
Vertebrate nesprins are giant (1 MDa) actin binding
(and nuclear) proteins with extensive spectrinlike repeats
(1335). Spectrin-related Drosophila proteins of 200 –300
kDa molecular mass, located at muscle adhesion sites in
early development and later at the Z line and which were
originally believed to be possible dystrophin orthologs,
were early identified (1007, 1233). It now appears these
proteins arise from a nesprin gene ortholog that could
produce isoforms as large as 11,720 amino acids (1335). A
similar gene has been identified in C. elegans (1335).
3. Amphiphysin
Vertebrates have two amphiphysins, one involved in
endocytosis and the other in striated muscle t tubules.
Drosophila has a single amphiphysin gene whose product
is involved only in muscle, as mutants have normal synaptic transmission but severely disorganized t tubule/sarcoplasmic reticulum systems (973, 1332) and defective
membrane protein integration in the postsynaptic muscle
membrane (761, 1327, 1332).
4. Various muscle cell attachment proteins
To transmit force to effectors, the thin filaments of
the terminal sarcomeres must attach to the plasma membrane, and thence to tendons or their equivalent. This
attachment is accomplished by a protein complex including integrin, talin, vinculin, and filamen. Several of these
proteins, prominently integrin, also control the complex’s
development. This complex is not fully understood in
vertebrates and is less so in invertebrates. We here therefore simply list a number of proteins that attach thin
filaments to the muscle cell membrane, or muscle cells to
tendons or body wall structures, including primarily references showing that they serve a structural role. These
proteins include connectin (Drosophila) (963), DIM-1 (C.
elegans) (1001), integrin and integrin-associated proteins
[jellyfish (975), C. elegans (349, 432, 450, 691, 712, 787,
1004 –1006, 1054), Drosophila (122, 124, 147, 183, 263, 666,
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
713, 854, 955, 1292, 1297, 1331), kakapo (Drosophila)
(380, 1120), myotactin (C. elegans) (454), spondin (Drosophila) (1201), talin (C. elegans) (219, 825), jellyfish
(975), Drosophila (136)], and vinculin (C. elegans) (74,
75). References 117, 125, 132, 134, 135, 149, 154, 389, 390,
465, 633, 947, 1002, 1329, 1330 are reviews or papers
describing muscle adhesion-deficient mutants or antibodies against cell adhesion molecules. Analysis of talin homologs from yeast, slime mold, and nematode reveals a
conserved protein domain believed to be specialized for
thin filament binding (772). Early monoclonal antibody
work identified three to five C. elegans muscle attachment
proteins (317); which, if any, of the above proteins correspond to these antibodies is unknown.
5. Dystrophin-related proteins
Mutations in the human dystrophin gene cause Duchenne muscular dystrophy. The dystrophin complex may
also help transmit sarcomere force to the plasma membrane, or help maintain membrane integrity under the
stress of sarcomere contraction (1031). The human dystrophin glycoprotein complex consists of dystrophin, dystrobrevin, distroglycan, sarcoglycan, sarcospan, syntrophin, and nitric oxide synthase (168). Given this close
association, all these genes will be considered together
here. The sea urchin dystrophin gene has been identified
and shares with vertebrates a complex structure with
multiple promoters and gene products (853, 1250). Immunohistochemistry and Western blot analysis show that
leech (Pondtobdella) has dystrophin, dystroglycan, sarcoglycan, syntrophin, and sarcospan homologs (1020, 1022).
C. elegans (113, 352, 382, 384) and Drosophila (243, 377,
853) have dystrophin, dystrobrevin, dystroglycan, sarcoglycan, and syntrophin (but not sarcospan) homologs.
The C. elegans and Drosophila dystrophin genes again
have multiple promoters and gene products (113, 853). C.
elegans muscle has a nitric oxide synthase binding protein
that can partially suppress dystrophin mutation (354).
The C. elegans proteins are also likely a complex, as
dystrophin (113), dystrobrevin (352, 355), dystroglycan
(384), sarcoglycan (384), and syntrophin (382, 384) mutants all have the same phenotype: hyperactivity, head
bending during forward locomotion, a tendency to hypercontract, and ACh hypersensitivity. Further genetic evidence of interaction is provided by data that increased
dystrobrevin expression delays muscle degeneration and
locomotor defects in dystrophin mutants (353). In vitro
binding assays give direct evidence of dystrophin, dystrobrevin, and syntrophin interaction and show that dystrophin/distrobrevin binding requires the second coiled-coil
dystrobrevin domain (351, 383). The basis of the mutant
phenotype is not well understood, but acetylcholinesterase activity is decreased (by an unknown mechanism) in
dystrophin mutants, which may explain the ACh hyperPhysiol Rev • VOL
1027
sensitivity (356). Ca2⫹ channel activity is important for
phenotype expression, as increased Ca2⫹ channel activity
increases, and reduced Ca2⫹ channel activity decreases,
muscle degeneration in dystrophin null mutants (742).
References 168, 186, 226, 999, 1066, and 1067 are reviews
of C. elegans and Drosophila as model systems for studying Duchenne muscular dystrophy.
6. Intermediate filaments
In vertebrates another protein that may help transmit
force from the sarcomere to the cell membrane, or help
maintain muscle cell integrity in the face of the stress
involved, is the intermediate filament protein desmin. In
mammals desmin is one of a large number of intermediate
filament types (others include keratin, neurofilament protein, and nuclear lamins) that can be ordered into five or
six classes on the basis of sequence homology (9, 319 are
primarily vertebrate reviews). Invertebrate intermediate
filaments are much less well understood, and their relationship to the mammalian classes remains unclear (256,
345, 530, 994, 996, 997, 1249, 1267, 1268). We cover here
only work directly involving invertebrate muscle intermediate filaments.
Early work showed surprising variation with respect
to the presence of intermediate filaments in invertebrate
muscles, even between closely related species. For instance, large amounts of intermediate filaments were
present in all muscle types (smooth, irregularly, and obliquely striated) in Ascaris, located throughout the sarcomere and around dense bodies, but not in C. elegans
muscle (80 – 82). A wide survey of major phylogenetic
groups showed that intermediate filaments were also
present in Acanthocephala, Echiura, and Chaetognatha
muscle, but absent from the other examined invertebrate
groups, including Urochordata and Cephalochordata (78,
81). It was subsequently shown that some invertebrate
intermediate filament proteins lack an epitope present in
all vertebrate intermediate filament proteins, and thus
that lack of muscle staining in some groups may be due to
the lack of this epitope rather than a lack of intermediate
filaments (995). Particularly in light of later work showing
muscle intermediate filaments are present in some of the
above “negative” groups (see below), the absence assignments made in this early work should be viewed with
caution.
Molecular genetic work has shown that Styela (a
urochordate) has two intermediate filament type genes,
one of which is expressed in smooth body wall muscle
and shows sequence similarity to desmin (997). Similarly,
one of the five intermediate filament protein genes in
Ciona (another urochordate) is expressed in muscle (535,
1249). Despite some confusion (994), it is now also clear
that in Branchiostoma (Cephalochordata) 1 of the 10
known (530) intermediate filament proteins is present in
85 • JULY 2005 •
www.prv.org
1028
SCOTT L. HOOPER AND JEFFREY B. THUMA
muscle (997). With respect to nematodes, however, the
early data have held, with muscles from species from
multiple genera (Acanthocheilonema, Anisakis, Ascaris,
Dirifilaria, Onchocerca, Toxocara, Trichinella) containing intermediate filament proteins or showing staining
(116, 1046, 1065, 1333, 1334), but none of the four intermediate filament proteins (of a total of 11) in C. elegans
for which it has been investigated being present in muscle
(although the absence of some does cause displaced muscles and paralysis) (364, 398, 531–534). Staining is also
present in the muscles of several platyhelminths (Echinococcus, Schistosoma, Taenia) (1046).
Intermediate filaments composed of a protein immunologically similar to desmin are also present in several,
but not all, annelids (79, 81, 239). An intermediate filament
protein gene identified using a tissue sample containing
body muscle and attached epidermis has been sequenced
in Lumbricus terristris, but it is unclear if this gene
codes the above protein (126). Proteins that form 2- to
4-nm filaments, but which are not similar to desmin or
vimentin by two-dimensional gel electrophoresis or immunological cross-reactivity, are present in sea urchin
muscle (957). Arthropods appear to lack all intermediate
filaments (except nuclear lamins) (81), an absence confirmed by the lack of intermediate filament genes in the
Drosophila genome (365, 1026).
7. The 29-kDa ascidian protein
A body wall protein from Halocynthia roretzi of
unknown function that copurifies with thin filaments is
localized near the plasma membrane, has some similarity
to heat shock proteins and ␣-cystallin, and has been isolated and sequenced (1084, 1151).
G. Summary
The most striking impression from the above is the
large number and variety of different proteins, and different isoforms of individual proteins, present across invertebrates. This variety provides two great opportunities.
First, because these proteins interact and thus must have
evolved together, comparison of concerted changes in
their sequences should provide powerful evidence with
respect to evolutionary relationships. In addition to the
studies on actin mentioned earlier, this work has been
begun with sequence comparisons of myosin regulatory
and essential light chain and troponin C (187); troponin C,
myosin heavy chain, myosin regulatory and essential light
chain, and actin (903); vertebrate, C. elegans, and Drosophila titin, twitchin, projectin, and unc-89 (427, 554);
and vertebrate, ascidian, and Drosophila troponin I (407).
The first work is in agreement with the phyla and lower
assignments in Figures 1–3 but does not resolve the question of invertebrate monophyly. The second indicates that
Physiol Rev • VOL
skeletal and cardiac muscle type tissues evolved before
the vertebrate/arthropod split, but vertebrate smooth
muscle appears to have arisen independently of other
muscle types, and arthropod striated, urochordate
smooth, and vertebrate muscle (except for smooth) share
a common ancestry. The third suggests that titin/kettin,
twitchin, projectin, and myosin light chain kinases form a
related group of which unc-89 is not a member, but it was
performed with too few invertebrate species to provide
insight into invertebrate phylogenetic relationships. The
fourth shows that a single troponin I gene (from which
multiple isoforms are produced by alternative splicing),
such as is present in ascidia and Drosophila, is the ancestral condition, and the gene duplications leading to the
three troponin I genes present in vertebrates occurred
after the ascidian/vertebrate split. Although these early
observations are intriguing, much further work is necessary for the phylogenetic opportunities present here to be
fully realized. Notable lacks are investigation of muscle
proteins in Cnidaria, in which so far only actin, myosin,
and tropomyosin have been shown to exist, and detailed
study of giant muscle proteins in Lophotrochozoa.
The second opportunity this variety provides is in
understanding protein ensembles. Despite their variety,
the different sets of proteins present in the various invertebrates must successfully perform relatively similar
functions (form thin and thick filaments and sarcomeres,
contract and relax on demand). Investigating how the
multiplicity of forms noted above continue to do so, and
the relationship between these forms and differences in
muscle structure and function, will presumably provide
great insight into fundamental questions of protein interaction and function that can be successfully applied to
other systems. The utility of this approach is well demonstrated by a recent analysis of myosin heavy and light
chains, which showed that head, neck, and tail domains of
the heavy chains, and the essential and regulatory light
chains, show very similar phylogenetic trees (600). This
work is beginning in Drosophila, in which large numbers
of myofibril mutants are available (332, 815, 869, 1223).
Comparison of phenotypes of single gene heterozygotes
to mutants pairwise heterozygous in various muscle component genes has been used to investigate gene product
interactions and how critical correct regulation of gene
product amount is in myofibril development (89, 331, 441,
1223), and null mutants of actin and myosin have been
used to examine the contributions of thin and thick filaments to sarcomere assembly (89). Similar genetic opportunities are available in C. elegans, as is also the ability to
use antisense RNA to inhibit expressions of specific genes
(301). The increasing availability of cloned genes in ascidia should allow similar work to proceed in this “higher”
invertebrate (174). The development of robotic approaches for simultaneously identifying the components
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
of multiprotein mixtures (proteomes) should dramatically
enhance these efforts (36).
It is critical to point out, however, that to date most
work has been performed in a small number of “model”
systems (Drosophila, C. elegans, Bivalvia, decapod crustacea), organisms heavily studied due to their disease or
economic importance (parasitic platyhelminths and nematodes), or organisms relatively closely related to vertebrates (Cephalochordata, Urochordata, Echinodermata).
Although this bias was understandable in the early days in
which the challenge was defining the sarcomeric muscle
protein complement, and protein and gene analysis and
sequencing tools were relatively primitive, neither justification exists today. These concerns about relatively small
numbers of studied species are deepened by the observation that Drosophila and C. elegans have apparently undergone extensive gene loss and rapid sequence divergence relative to the common ancestral metazoan (601).
Full achievement of the insights about invertebrate phylogeny and protein structure and function that the multiplicity of invertebrate muscle proteins allows can only be
achieved by overcoming this species parochialism and
examining a much larger range of the enormous number
of invertebrate species that exist.
This work would not have been possible without, and we express our
deep gratitude to, C. A. Barcroft, R. P. Harrison, R. R. Heck III, K. H.
Hobbs, S. McKean, B. Revill, J. Thuma, and R. Thuma for help in creating
and maintaining the reference database and to the Ohio University
library system and OHIOLINK electronic journal center for providing
several thousand journal article photocopies and PDFs.
This work was supported by grants from the National Institutes of
Health and the University of Cologne, Cologne, Germany (to S. L.
Hooper).
Address for reprint requests and other correspondence: S. L. Hooper,
Neuroscience Program, Dept. of Biological Sciences, Irvine Hall, Ohio
University, Athens, OH 45701 (E-mail: [email protected]).
REFERENCES
1. Abbas MK and Cain GD. Analysis of isoforms of actin from
Schistosoma mansoni by two-dimensional gel electrophoresis.
Parasitol Res 76: 178 –180, 1989.
2. Achazi RK. Phosphorylation of molluscan paramyosin. Pflügers
Arch 379: 197–201, 1979.
3. Acosta LP, Waine G, Aligui GDL, Tiu WU, Olveda RM, and
McManus DP. Immune correlate study on human Schistosoma
japonicum in a well-defined population in Leyte, Philippines. II.
Cellular immune responses to S. japonicum recombinant and
native antigens. Acta Tropica 84: 137–149, 2002.
4. Adoutte A, Balavoine G, Lartillot N, Lespinet O,
Prud’homme B, and de Rosa R. The new animal phylogeny:
reliability and implications. Proc Natl Acad Sci USA 97: 4453–
4456, 2000.
5. Aerne B, Groger H, Schuchert P, Spring J, and Schmid V. The
polyp and its medusa: a molecular approach. Sci Mar 60: 7–16,
1996.
6. Aidley DJ. The Physiology of Excitable Cells. Cambridge, UK:
Cambridge Univ. Press, 1998.
7. Aki T, Kodama T, Fujikawa A, Miura K, Shigeta S, Wada T,
Jyo T, Murooka Y, Oka S, and Ono K. Immunochemical characterization of recombinant and native tropomyosins as a new
allergen from the house dust mite, Dermatophagoides farinae. J
Allergy Clin Immunol 96: 74 – 83, 1995.
Physiol Rev • VOL
1029
8. Akpek EK, Liu SH, Thompson R, and Gottsch JD. Identification of paramyosin as a binding protein for calgranulin C in
experimental helminthic keratitis. Invest Ophthalmol Vis Sci 43:
2677–2684, 2002.
9. Albers K and Fuchs E. The molecular biology of intermediate
filament proteins. Int Rev Cytol 134: 243–279, 1992.
10. Albertson DG. Mapping muscle protein genes by in situ hybridization using biotin-labeled probes. EMBO J 4: 2493–2498, 1985.
11. Allen ML and Christensen BM. Flight muscle-specific expression of act88F: GFP in transgenic Culex quinquefasciatus Say
(Diptera: Culicidae). Parasitol Int 53: 307–314, 2004.
12. Allhouse LD, Potter JD, and Ashley CC. A novel method of
extraction of TnC from skeletal muscle myofibrils. Pflügers Arch
437: 695–701, 1999.
13. Al-Sherbiny M, Osman A, Barakat R, El Morshedy H,
Bergquist R, and Olds R. In vitro cellular and humoral responses
to Schistosoma mansoni vaccine candidate antigens. Acta
Tropica 88: 117–130, 2003.
14. Ambrosio J, Cruz-Rivera M, Allan J, Morán E, Ersfeld K, and
Flisser A. Identification and partial characterization of a myosinlike protein from cysticerci and adults of Taenia solium using a
monoclonal antibody. Parasitology 114: 545–553, 1997.
15. Ambrosio JR, Reynoso-Ducoing O, Hernandez-Sanchez H,
Correa-Pina D, Gonzalez-Malerva L, Cruz-Rivera M, and
Flisser A. Actin expression in Taenia solium cysticerci (Cestoda): tisular distribution and detection of isoforms. Cell Biol Int
27: 727–733, 2003.
16. An HS and Mogami K. Isolation of 88F actin mutants of Drosophila melanogaster and possible alterations in the mutant actin
structures. J Mol Biol 260: 492–505, 1996.
17. Anderson P. Molecular genetics of nematode muscle. Annu Rev
Genet 23: 507–525, 1989.
18. Ando M, Ando M, Tsukamasa Y, Makiodan Y, and Miyoshi M.
Muscle firmness and structure of raw and cooked arrow squid
mantle as affected by freshness. J Food Sci 64: 659 – 662, 1999.
19. Ando M, Nakamura H, Harada R, and Yamane A. Effect of
super chilling storage on maintenance of freshness of kuruma
prawn. Food Sci Technol Res 10: 25–31, 2004.
20. Anson M, Drummond DR, Geeves MA, Hennessey ES, Ritchie
MD, and Sparrow JC. Actomyosin kinetics and in vitro motility
of wild-type Drosophila actin and the effects of two mutations in
the Act88F gene. Biophys J 68: 1991–2003, 1995.
21. Anyanful A, Sakube Y, Takuwa K, and Kagawa H. The third
and fourth tropomyosin isoforms of Caenorabditis elegans are
expressed in the pharynx and intestines and are essential for
development and morphology. J Mol Biol 313: 525–537, 2001.
22. Ao WY and Pilgrim D. Caenorhabditis elegans unc-45 is a
component of muscle thick filaments and colocalizes with myosin
heavy chain B, but not myosin heavy chain A. J Cell Biol 148:
375–384, 2000.
23. Araki I and Satoh N. Cis-regulatory elements conserved in the
proximal promoter region of an ascidian embryonic muscle myosin heavy chain gene. Dev Genes Evol 206: 54 – 63, 1996.
24. Ardizzi JP and Epstein HF. Immunochemical localization of
myosin heavy chain isoforms and paramyosin in developmentally
and structurally diverse muscle cell types of the nematode Caenorhabditis elegans. J Cell Biol 105: 2763–2770, 1987.
25. Arlian LG. Arthropod allergens and human health. Annu Rev
Entomol 47: 395– 433, 2002.
26. Arner A, Lofgren M, and Morano I. Smooth, slow and smart
muscle motors. J Muscle Res Cell Motil 24: 165–173, 2003.
27. Arredondo JJ, Ferreres RM, Maroto M, Cripps RM, Marco R,
Bernstein SI, and Cervera M. Control of Drosophila paramyosin/miniparamyosin gene expression. Differential regulatory
mechanisms for muscle-specific transcription. J Biol Chem 276:
8278 – 8287, 2001.
28. Arredondo JJ, Mardahl-Dumesnil M, Cripps RM, Cervera M,
and Bernstein SI. Overexpression of miniparamyosin causes
muscle dysfunction and age-dependent myofibril degeneration in
the indirect flight muscles of Drosophila melanogaster. J Muscle
Res Cell Motil 22: 287–299, 2001.
85 • JULY 2005 •
www.prv.org
1030
SCOTT L. HOOPER AND JEFFREY B. THUMA
29. Arruda LK, Vailes LD, Ferriani VPL, Santos BR, Pomes A,
and Chapman MD. Cockroach allergens and asthma. J Allergy
Clin Immunol 107: 419 – 428, 2001.
30. Asakawa T and Azuma N. Proteolytic digestion of myosin from
abalone Haliotis discus smooth muscle [in Japanese, English
summary]. Nippon Suisan Gakk 56: 297–305, 1990.
31. Asakawa T and Azuma N. Inhibitory effect of MgATP on the
release of regulatory light chain from scallop myosin and light
chain composition of scallop myosin hybridized with abalone light
chain 2 at 30°C. J Biochem 94: 395– 401, 1983.
32. Asakawa T, Yazawa Y, and Azuma N. Light chains of abalone
myosin. UV absorption difference spectrum and resensitization of
desensitized scallop myosin. J Biochem 1981: 1805–1814, 1981.
33. Asero R, Mistrello G, Roncarolo D, and Amato S. A case of
allergy to airborne, heat-labile shrimp allergens. J Allergy Clin
Immunol 109: 371–372, 2002.
34. Ashiba G and Szent-Györgyi AG. Essential light chain exchange
in scallop myosin. Biochemistry 24: 6618 – 6623, 1985.
35. Ashley CC, Lea TJ, Hoar PE, Kerrick WGL, Strang PF, and
Potter JD. Functional characterization of the two isoforms of
troponin C from the arthropod Balanus nubilus. J Muscle Res Cell
Motil 12: 532–542, 1991.
36. Ashman K, Houthaeve T, Clayton J, Wilm M, Podtelejnikov
A, Jensen ON, and Mann M. The application of robotics and
mass spectrometry to the characterisation of the Drosophila
melanogaster indirect flight muscle proteome. Lett Pept Sci 4:
57– 65, 1997.
37. Asturias JA, Arilla MC, Gomez-Bayon N, Martinez A, Martinez J, and Palacios R. Sequencing and high level expression in
Escherichia coli of the tropomyosin allergen (Der p 10) from
Dermatophagoides pteronyssinus. Biochim Biophys Acta 1397:
27–30, 1998.
38. Asturias JA, Eraso E, Arilla MC, Gomez-Bayon N, Inacio F,
and Martinez A. Cloning, isolation, and IgE-binding properties of
Helix aspersa (brown garden snail) tropomyosin. Int Arch Allergy
Immunol 128: 90 –96, 2002.
39. Asturias JA, Eraso E, and Martinez A. Cloning and high level
expression in Escherichia coli of an Anisakis simplex tropomyosin isoform. Mol Biochem Parasitol 108: 263–266, 2000.
40. Asturias JA, Eraso E, Moneo I, and Martinez A. Is tropomyosin an allergen in Anisakis? Allergy 55: 898 – 899, 2000.
41. Asturias JA, Gomez-Bayon N, Arilla MC, Martinez A, Palacios R, Sanchez-Gascon F, and Martinez J. Molecular characterization of American cockroach tropomyosin (Periplaneta
americana Allergen 7), a cross-reactive allergen. J Immunol 162:
4342– 4348, 1999.
42. Audemard E, Bertrand R, Bonet A, Chaussepied P, and Mornet D. Pathway for the communication between the ATPase and
actin sites in myosin. J Muscle Res Cell Motil 9: 197–218, 1988.
43. Avery L. The genetics of feeding in Caenorhabditis elegans.
Genetics 133: 897–917, 1993.
44. Ayer G and Vigoreaux JO. Flightin is a myosin rod binding
protein. Cell Biochem Biophys 38: 41–54, 2003.
45. Ayme-Southgate A, Bounaix C, Riebe TE, and Southgate R.
Assembly of the giant protein projectin during myofibrillogenesis
in Drosophila indirect flight muscles. BMC Cell Biol 5: 17, 2004.
46. Ayme-Southgate A, Lasko P, French C, and Pardue ML. Characterization of the gene for mp20: a Drosophila muscle protein
that is not found in asynchronous oscillatory flight muscle. J Cell
Biol 108: 521–531, 1989.
47. Ayme-Southgate A, Southgate R, and McEliece MK. Drosophila projectin: a look at protein structure and sarcomeric assembly.
Adv Exp Med Biol 481: 251–264, 2000.
48. Ayme-Southgate A, Southgate R, Saide J, Benian GM, and
Pardue ML. Both synchronous and asynchronous muscle isoforms of projectin (the Drosophila bent locus product) contain
functional kinase domains. J Cell Biol 128: 393– 403, 1995.
49. Ayme-Southgate A, Vigoreaux J, Benian G, and Pardue ML.
Drosophila has a twitchin/titin related gene that appears to encode projectin. Proc Natl Acad Sci USA 88: 7973–7977, 1991.
50. Ayuso R, Lehrer SB, and Reese G. Identification of continuous,
allergenic regions of the major shrimp allergen Pen a 1 (tropomyosin). Int Arch Allergy Immunol 127: 27–37, 2002.
Physiol Rev • VOL
51. Ayuso R, Reese G, Leong-Kee S, Plante M, and Lehrer SB.
Molecular basis of arthropod cross-reactivity: IgE-binding crossreactive epitopes of shrimp, house dust mite, and cockroach
tropomyosins. Int Arch Allergy Immunol 129: 38 – 48, 2002.
52. Azuma N, Asakura A, and Koichi Y. Myosin from molluscan
abalone, Haliotis discus. J Biochem 77: 973–981, 1975.
53. Baader CD, Schmid V, and Schuchert P. Characterization of a
tropomyosin cDNA from the hydrozoan Podocoryne carnea.
FEBS Lett 328: 63– 66, 1993.
54. Babu YS, Cox JA, and Cook WJ. Crystallization and preliminary
X-ray investigation of sarcoplasmic calcium-binding protein from
Nereis diversicolor. J Biol Chem 262: 11884 –11885, 1987.
55. Bailey JT. Composition of the myosins and myogen of skeletal
muscle. Biochem J 31: 1406 –1413, 1937.
56. Bailey K. Invertebrate tropomyosin. Biochim Biophys Acta 24:
612– 619, 1957.
57. Bailey K. The proteins of adductor muscles. Pubbl Staz Zool
Napoli 29: 96 –108, 1956.
58. Bailey K, De Milstein CP, Kay CM, and Smillie LB. Characterization of a tryptic fragment isolated from the insoluble tropomyosin of Pinna nobilis. Biochim Biophys Acta 90: 503–520,
1964.
59. Bailey K and Milstein CP. Tryptic hydrolysis of paramyosin
from invertebrates: rate and extent of proteolysis. Biochim Biophys Acta 90: 492–502, 1964.
60. Bailey K and Rüegg JC. Further chemical studies on the tropomyosins of lamellibranch muscle with special reference to Pecten
maximus. Biochim Biophys Acta 38: 239 –245, 1960.
61. Baker JP and Titus MA. A family of unconventional myosins
from the nematode Caenorhabditis elegans. J Mol Biol 272: 523–
535, 1997.
62. Baladi S, Tsvetkov PO, Petrova TV, Takagi T, Sakamoto H,
Lobachov VM, Makarov AA, and Cox JA. Folding units in
calcium vector protein of amphioxus: structural and functional
properties of its amino- and carboxy- terminal halves. Protein Sci
10: 771–778, 2001.
63. Baladi S, Tsvetkov PO, Petrova TV, Takagi T, Sakamoto H,
Lobachov VM, Makarov AA, and Cox JA. Folding units in
calcium vector protein of amphioxus: structural and funtional
properties of its amino- and carboxy-terminal halves (erratum to
Protein Sci 10: 771–778, 2001). Protein Sci 10: 1279, 2001.
64. Ball E, Karlik CC, Beall CJ, Saville DL, Sparrow JC, Bullard
B, and Fyrberg EA. Arthrin, a myofibrillar protein of insect flight
muscle, is an actin-ubiquitin conjugate. Cell 51: 221–228, 1987.
65. Ban M, Ushio H, and Yamanaka H. Post-mortem biochemical
changes in muscle of Buccinum striatissium during storage.
Fish Sci 64: 482– 486, 1998.
66. Bandman E. Myosin isoenzyme transitions in muscle development, maturation, and disease. Int Rev Cytol 97: 97–131, 1985.
67. Bárány M and Bárány K. Myosin from the striated adductor
muscle of scallop (Pecten arradians). Biochemistry 345: 37–56,
1966.
68. Barbas JA, Galceran J, Krahjentgens I, Delapompa JL, Canal I, Pongs O, and Ferrus A. Troponin I is encoded in the
haplolethal region of the shaker gene complex of Drosophila.
Genes Dev 5: 132–140, 1991.
69. Barbas JA, Galceran J, Torroja L, Prado A, and Ferrus A.
Abnormal muscle development in the heldup3 mutant of Drosophila melanogaster is caused by a splicing defect affecting selected
troponin I isoforms. Mol Cell Biol 13: 1433–1439, 1993.
70. Barouch WW, Breese KE, Davidoff SA, Leszyk J, SzentGyörgyi AG, Theibert JL, and Collins JH. Amino acid sequences of myosin essential and regulatory light chains from two
clam species: comparison with other molluscan myosin light
chains. J Muscle Res Cell Motil 12: 321–332, 1991.
71. Barral JM, Bauer CC, Ortiz I, and Epstein HF. unc-45 mutations in Caenorhabditis elegans implicate a CRO1/She4p-like domain in myosin assembly. J Cell Biol 143: 1215–1225, 1998.
72. Barral JM, Hutagalung AH, Brinker A, Hartl FU, and Epstein
HF. Role of the myosin assembly protein unc-45 as a molecular
chaperone for myosin. Science 295: 669 – 671, 2002.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
73. Barstead RJ, Kleiman L, and Waterston RH. Cloning, sequencing, and mapping of an alpha-actinin gene from the nematode
Caenorhabditis elegans. Cell Motil Cytoskeleton 20: 69 –78, 1991.
74. Barstead RJ and Waterston RH. The basal component of the
nematode dense body is vinculin. J Biol Chem 264: 10177–10185,
1989.
75. Barstead RJ and Waterston RH. Vinculin is essential for muscle
function in the nematode. J Cell Biol 114: 715–724, 1991.
76. Bartegi A, Fattoum A, Dagorn C, Gabrion J, and Kassab R.
Isolation, characterization, and immunocytochemical localization
of caldesmon-like protein from molluscan striated muscle. Eur
J Biochem 185: 589 –595, 1989.
77. Barthmaier P and Fyrberg E. Monitoring development and
pathology of Drosophila indirect flight muscles using green fluorescent protein. Dev Biol 169: 770 –774, 1995.
78. Bartnik E, Kossmagk-Stephan K, Osborn M, and Weber K. In
gastropods intermediate filaments occur in epithelia, glia, connective tissue cells, and neurones but not in muscle. Eur J Cell Biol
43: 329 –338, 1987.
79. Bartnik E, Kossmagk-Stephan K, and Weber K. Evidence for
two intermediate filament prototypes in the invertebrate Myxicola; neurofilaments and non-neuronal intermediate filaments differ in subunit size and immunological properties. Eur J Cell Biol
44: 219 –228, 1987.
80. Bartnik E, Osborn M, and Weber K. Intermediate filaments in
muscle and epithelial cells of nematodes. J Cell Biol 102: 2033–
2041, 1986.
81. Bartnik E and Weber K. Widespread occurrence of intermediate
filaments in invertebrates: common principles and aspects of
diversion. Eur J Cell Biol 50: 17–33, 1989.
82. Bartnik E and Weber K. Intermediate filaments in the giant
muscle cells of the nematode Ascaris lumbricoides: abundance
and three-dimensional complexity of arrangements. Eur J Cell
Biol 45: 291–301, 1987.
83. Bashir M, Bickle Q, Bushara H, Cook L, Shi FH, He DA,
Huggins M, Lin JJ, Malik K, Moloney A, Mukhtar M, Ping Y,
Xu ST, Taylor M, and Shi YC. Evaluation of defined antigen
vaccines against Schistosoma bovis and Schistosoma japonicum
in bovines. Tropical Geographical Med 46: 255–258, 1994.
84. Basi GS, Boardman M, and Storti RV. Alternative splicing of a
Drosophila tropomyosin gene generates muscle tropomyosin isoforms with different carboxy-terminal ends. Mol Cell Biol 4: 2828 –
2836, 1984.
85. Basi GS and Storti RV. Structure and DNA sequence of the
tropomyosin I gene from Drosophila melanogaster. J Biol Chem
261: 817– 827, 1986.
86. Bautch VL, Storti RV, Mischke D, and Pardue ML. Organization and expression of Drosophila tropomyosin genes. J Mol Biol
162: 231–250, 1982.
87. Beach RL and Jeffery WR. Multiple actin genes encoding the
same ␣-muscle isoform are expressed during ascidian development. Dev Biol 151: 55– 66, 1992.
88. Beall CJ and Fyrberg E. Muscle abnormalities in Drosophila
melanogaster heldup mutants are caused by missing or aberrant
troponin-I isoforms. J Cell Biol 114: 941–951, 1991.
89. Beall CJ, Sepanski MA, and Fyrberg EA. Genetic dissection of
Drosophila myofibril formation: effects of actin and myosin heavy
chain null alleles. Genes Dev 3: 131–140, 1989.
90. Becker KD, O’Donnell PT, Heitz JM, Vito M, and Bernstein
SI. Analysis of Drosophila paramyosin: identification of a novel
isoform which is restricted to a subset of adult muscles. J Cell
Biol 116: 669 – 681, 1992.
91. Becker MM. Gene cloning and complete nucleotide sequence of
Philippine Schistosoma japonicum paramyosin. Acta Tropica 59:
143–147, 1995.
92. Beifuss MJ and Durica DS. Sequence analysis of the indirect
flight muscle actin encoding gene of Drosophila simulans. Gene
118: 163–170, 1992.
93. Beinbrech G, Meller U, and Sasse W. Paramyosin content and
thick filament structure in insect muscles. Cell Tissue Res 241:
607– 614, 1985.
Physiol Rev • VOL
1031
94. Bejsovec A and Anderson P. Myosin heavy-chain mutations that
disrupt Caenorhabditis elegans thick filament assembly. Genes
Dev 2: 1307–1317, 1988.
95. Benian G, Ayme-Southgate A, and Tinley TL. The genetics and
molecular biology of the titin/connectin-like proteins of invertebrates. Rev Physiol Biochem Pharmacol 138: 235–268, 1999.
96. Benian GM, Kiff JE, Neckelmann N, Moerman DG, and Waterston RH. Sequence of an unusually large protein implicated in
regulation of myosin activity in C. elegans. Nature 342: 45–50,
1989.
97. Benian GM, Lhernault SW, and Morris ME. Additional sequence complexity in the muscle gene, unc-22, and its encoded
protein, twitchin, of Caenorhabditis elegans. Genetics 134: 1097–
1104, 1993.
98. Benian GM, Tang XX, and Tinley TL. Twitchin and related
giant IG superfamily members of C. elegans and other invertebrates. Adv Biophys 33: 183–198, 1996.
99. Benian GM, Tinley TL, Tang XX, and Borodovsky M. The
Caenorhabditis elegans gene unc-89, required for muscle M line
assembly, encodes a giant modular protein composed of Ig and
signal transduction domains. J Cell Biol 132: 835– 848, 1996.
100. Bennett PM and Marston SB. Calcium regulated thin filaments
from molluscan catch muscles contain a caldesmon-like regulatory protein. J Muscle Res Cell Motil 11: 302–312, 1990.
101. Benoist P, Mas JA, Marco R, and Cervera M. Differential
muscle type expression of the Drosophila troponin T gene: a
3-base pair microexon is involved in visceral and hypodermic
muscle specification. J Biol Chem 273: 7538 –7546, 1998.
102. Benzonana G, Cox JA, Kohler L, and Stein EA. Caractérisation d’une nouvelle métallo-protéine calcique du myogène de certains crustacés [in French]. C R Acad Sci Paris 279: 1491–1493,
1974.
103. Benzonana G, Kohler L, and Stein EA. Regulatory proteins of
crayfish tail muscle. Biochim Biophys Acta 368: 247–258, 1974.
104. Berg JS, Powell BC, and Cheney RE. A millennial myosin
census. Mol Biol Cell 12: 780 –794, 2001.
105. Bergquist NR. Controlling schistosomiasis by vaccination: a realistic option? Parasitol Today 11: 191–194, 1995.
106. Bergquist NR. Schistosomiasis vaccine development: progress
and prospects. Mem Inst Oswaldo Cruz 93: 95–101, 1998.
107. Bergquist NR. Schistosomiasis: from risk assessment to control.
Trends Parasitol 18: 309 –314, 2002.
108. Bergquist R, Al-Sherbiny M, Barakat R, and Olds R. Blueprint
for schistosomiasis vaccine development. Acta Tropica 82: 183–
192, 2002.
109. Bernstein SI, Hansen CJ, Becker KD, Wassenberg DR, Roche
ES, Donady JJ, and Emerson CP Jr. Alternative RNA splicing
generates transcripts encoding a thorax-specific isoform of Drosophila melanogaster myosin heavy chain. Mol Cell Biol 6: 2511–
2519, 1986.
110. Bernstein SI and Milligan RA. Fine tuning a molecular motor:
the location of alternative domains in the Drosophila myosin
head. J Mol Biol 271: 1– 6, 1997.
111. Bernstein SI, Mogami K, Donady JJ, and Emerson CP Jr.
Drosophila muscle myosin heavy chain encoded by a single gene
in a cluster of muscle mutations. Nature 302: 393–397, 1983.
112. Bernstein SI, O’Donnell PT, and Cripps RM. Molecular genetic
analysis of muscle development, structure, and function in Drosophila. Int Rev Cytol 143: 63–152, 1993.
113. Bessou C, Giugia JB, Franks CJ, Holden-Dye L, and Segalat
L. Mutations in the Caenorhabditis elegans dystrophin-like gene
dys-1 lead to hyperactivity and suggest a link with cholinergic
transmission. Neurogenetics 2: 61–72, 1998.
114. Bickle QD, Bøgh HO, Johansen MV, and Zhang YB. Comparison of the vaccine efficacy of ␥-irradiated Schistosoma japonicum cercariae with the defined antigen Sj562(IrV-5) in pigs. Vet
Parasitol 100: 51– 62, 2001.
115. Bing W, Razzaq A, Sparrow J, and Marston S. Tropomyosin
and troponin regulation of wild type and E93K mutant actin
filaments from Drosophila flight muscle. J Biol Chem 273: 15016 –
15021, 1998.
116. Bisoffi M and Betschart B. Ascaris suum: molecular cloning of
an intermediate filament. Trop Med Int Health 1: 640 – 645, 1996.
85 • JULY 2005 •
www.prv.org
1032
SCOTT L. HOOPER AND JEFFREY B. THUMA
117. Bitsch C and Bitsch J. The endoskeletal structures in arthropods: cytology, morphology, and evolution. Arthropod Struct Dev
30: 159 –177, 2002.
118. Blaxter M, Daub J, Guilano D, Parkinson J, Whitton C, and
Filarial Genome Project. Pathogen genomes and human health.
The Brugia malayi genome project: expressed sequence tags and
gene discovery. Trans R Soc Trop Med Hyg 96: 7–17, 2002.
119. Blaxter ML, Raghavan N, Ghosh I, Guiliano D, Lu W, Williams SA, Slatko B, and Scott AL. Genes expressed in Brugia
malayi infective third stage larvae. Mol Biochem Parasitol 77:
77–93, 1996.
120. Blomberg N, Baraldi E, Sattler M, Saraste M, and Nilges M.
Structure of a PH domain from the C. elegans muscle protein
UNC-89 suggests a novel function. Structure 8: 1079 –1087, 2000.
121. Blomberg N, Sattler M, and Nilges M. Letter to the Editor: 1H,
15
N, and 13C resonance assignment of the PH domain from C.
elegans UNC-89. J Biomol NMR 15: 269 –270, 1999.
122. Bloor JW and Brown NH. Genetic analysis of the Drosophila
alpha(PS2) integrin subunit reveals discrete adhesive, morphogenetic, and sarcomeric functions. Genetics 148: 1127–1142, 1998.
123. Blyakhman F, Tourovskaya A, and Pollack AJ. Intact connecting filaments change length in 2.3 nm quanta. Adv Exp Med Biol
481: 305–318, 2000.
124. Bogaert T, Brown N, and Wilcox M. The Drosophila PS2 antigen is an invertebrate integrin that, like the fibronectin receptor,
becomes localized to muscle attachments. Cell 51: 929 –940, 1987.
125. Bökel C and Brown NH. Integrins in development: moving on,
responding to, and sticking to the extracellular matrix. Dev Cell 3:
311–321, 2002.
126. Bovenschulte M, Riemer D, and Weber K. The sequence of a
cytoplasmic intermediate filament (IF) protein from the annelid
Lumbricus terristris emphasizes a distinctive feature of protostomic IF proteins. FEBS Lett 360: 223–226, 1995.
127. Bovenschulte M and Weber K. Deuterostomic actin genes and
the definition of the chordates: cDNA cloning and gene organization for cephalochordates and hemichordates. J Mol Evol 45:
653– 660, 1997.
128. Brading A. The Autonomic Nervous System and its Effectors.
Oxford, UK: Blackwell Science, 1999.
129. Brault V, Reedy MC, Sauder U, Kammerer RA, Aebi U, and
Schoenenberger C. Substitution of flight muscle specific actin by
human (beta) cytoplasmic actin in the indirect flight muscle of
Drosophila. J Cell Sci 112: 3627–3639, 1999.
130. Brault V, Sauder U, Reedy MC, Aebi U, and Schoenenberger
CA. Differential epitope tagging of actin in transformed Drosophila produces distinct effects on myofibril assembly and function of
the indirect flight muscle. Mol Biol Cell 10: 135–149, 1999.
131. Brenner S. The genetics of Caenorhabditis elegans. Genetics 77:
71–94, 1974.
132. Brower DL. Platelets with wings: the maturation of Drosophila
integrin biology. Curr Opin Cell Biol 15: 607– 613, 2003.
133. Brown LD and Cantino ME. Nonuniform distribution of myosin
light chains within the thick filaments of lobster slow muscle:
immunocytochemical study. J Exp Zool 290: 6 –17, 2001.
134. Brown NH. Integrins hold Drosophila together. Bioessays 15:
383–390, 1993.
135. Brown NH. Cell-cell adhesion via the ECM: integrin genetics in fly
and worm. Matrix Biol 19: 191–201, 2000.
136. Brown NH, Gregory SL, Rickoll WL, Fessler LI, Prout M,
White RAH, and Fristrom JW. Talin is essential for integrin
function in Drosophila. Dev Cell 3: 569 –579, 2002.
137. Buku A, Probst WC, Weiss KR, and Heierhorst J. Studies of
the calmodulin binding site of twitchin with synthetic peptides
using fluorescence and CD spectroscopy. Biochem Biophys Res
Commun 218: 854 – 859, 1996.
138. Bullard B, Bell J, Craig R, and Leonard K. Arthrin: a new
actin-like protein in insect flight muscle. J Mol Biol 182: 443– 454,
1985.
139. Bullard B, Dabrowska R, and Winkelman L. The contractile
and regulatory proteins of insect flight muscle. Biochem J 135:
277–286, 1973.
Physiol Rev • VOL
140. Bullard B, Goulding D, Ferguson C, and Leonard K. Links in
the chain: the contribution of kettin to the elasticity of insect
muscles. Adv Exp Med Biol 481: 207–220, 2000.
141. Bullard B, Hammond KS, and Luke BM. The site of paramyosin
in insect flight muscle and the presence of an unidentified protein
between myosin filaments and Z line. J Mol Biol 115: 417– 440,
1977.
142. Bullard B and Leonard K. Modular proteins of insect muscle.
Adv Biophys 33: 211–221, 1996.
143. Bullard B, Leonard K, Larkins A, Butcher G, Karlik CC, and
Fyrberg E. Troponin of asynchronous flight muscle. J Mol Biol
204: 621– 637, 1988.
144. Bullard B, Linke WA, and Leonard K. Varieties of elastic protein in invertebrate muscles. J Muscle Res Cell Motil 23: 435– 447,
2002.
145. Bullard B, Luke B, and Winkelman L. The paramyosin of insect
flight muscle. J Mol Biol 75: 359 –367, 1973.
146. Bullard B and Sainsbury GM. The proteins in the Z line of insect
flight muscle. Biochem J 161: 399 – 403, 1977.
147. Bunch TA, Graner MW, Fessler LI, Fessler JH, Schneider
KD, Kerschen A, Choy LP, Burgess BW, and Brower DL. The
PS2 integrin ligand tiggrin is required for proper muscle function
in Drosophila. Development 125: 1679 –1689, 1998.
148. Burgess S, Walker M, Knight PJ, Sparrow J, Schmitz S, Offer
G, Bullard B, Leonard K, Holt J, and Trinick J. Structural
studies of arthrin: monoubiquitinated actin. J Mol Biol 341: 1161–
1173, 2004.
149. Burke RD. Invertebrate integrins: structure, function, and evolution. Int Rev Cytol 191: 257–284, 1999.
150. Bush RK and Hefle SL. Food allergens. Crit Rev Food Sci Nutr
36: S119 –S163, 1996.
151. Butler TM, Mooers SU, Li CQ, Narayan S, and Siegman MJ.
Regulation of catch muscle by twitchin phosphorylation: effects
on force, ATPase, and shortening. Biophys J 75: 1904 –1914, 1998.
152. Butler TM, Narayan SR, Mooers SU, Hartshorne DJ, and
Siegman MJ. The myosin cross-bridge cycle and its control by
twitchin phosphorylation in catch muscle. Biophys J 80: 415– 426,
2001.
153. Cadoret JP, Debon R, Cornudella L, Lardans V, Morvan A,
Roch P, and Boulo V. Transient expression assays with the
proximal promoter of a newly characterized actin gene from the
oyster Crassostrea gigas. FEBS Lett 460: 81– 85, 1999.
154. Calderwood DA, Shattil SJ, and Ginsberg MH. Integrins and
actin filaments: reciprocal regulation of cell adhesion and signaling. J Biol Chem 275: 22607–22610, 2000.
155. Campos A, Bernard PH, Faucconier A, Landa A, Gomez E,
Hernandez R, Willms K, and Laclette PJ. Cloning and sequencing of two actin genes from Taenia solium. Mol Biochem Parasitol 40: 87–94, 1990.
156. Cancela M, Carmona C, Rossi S, Frangione B, Goni F, and
Berasain P. Purification, characterization, and immunolocalization of paramyosin from the adult stage of Fasciola hepatica.
Parasitol Res 92: 441– 448, 2004.
157. Cao J and Liu S. Immunization of mice with native tropomyosin
from Schistosoma japonicum and Oncomelania hupensis. Chin
J Parasitol Parasitic Dis 16: 406 – 410, 1998.
158. Cao J, Liu S, Song G, and Xu Y. Cloning of cDNA encoding
Schistosoma japonicum tropomyosin and its expression in Escherischia coli. Chin Med J 115: 1465–1469, 2002.
159. Capron A, Riveau GJ, Bartley PB, and McManus DP. Prospects for a schistosome vaccine. Curr Drug Targets Immune
Endocr Metab Disord 2: 281–290, 2002.
160. Carlhoff D and D’Haese J. Slow type muscle cells in the earthworm gizzard with a distinct, Ca2⫹-regulated myosin isoform.
J Comp Physiol B Biochem Syst Environ Physiol 157: 589 –597,
1987.
161. Carlini DB, Reece KS, and Graves JE. Actin gene family evolution and the phylogeny of coleoid cephalopods (Mollusca:
Cephalopoda). Mol Biol Evol 17: 1353–1370, 2000.
162. Castagnone-Sereno P, Leroy F, and Abad P. cDNA cloning and
expression analysis of a calponin gene from the plant-parasitic
nematode Meloidogyne incognita. Mol Biochem Parasitol 112:
149 –152, 2001.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
163. Castellani L and Cohen C. Myosin rod phosphorylation and the
catch state of molluscan muscles. Science 235: 334 –337, 1987.
164. Castellani L and Cohen C. Rod phosphorylation favors folding
in a catch muscle myosin. Proc Natl Acad Sci USA 84: 4058 – 4062,
1987.
165. Castellani L, Elliott BW Jr, and Cohen C. Phosphorylatable
serine residues are located in non-helical tailpiece of a catch
muscle. J Muscle Res Cell Motil 9: 533–540, 1988.
166. Castellani L and Vibert P. Location of paramyosin in relation to
the subfilaments within the thick filaments of scallop striated
muscle. J Muscle Res Cell Motil 13: 174 –182, 1992.
167. Castellani-Ceresa L and Lanzavecchia G. Isolation and identification of paramyosin from amphioxus notochord. J Muscle Res
Cell Motil 3: 75– 85, 1982.
168. Chamberlain JS and Benian GM. Muscular dystrophy: the
worm turns to genetic disease. Curr Biol 10: R795–R797, 2000.
169. Chantler PD and Szent-Györgyi AG. Regulatory light-chains
and scallop myosin: full dissociation, reversibility and co-operative effects. J Mol Biol 138: 473– 492, 1980.
170. Chen CL and Chung CY. Characterization of actomyosin gel
formation of squids. J Chin Agric Chem Soc 27: 322–330, 1989.
171. Chen H, Nara T, Zeng X, Satoh M, Wu G, Jiang W, Yi F,
Kojima S, Zhang S, and Hirayama K. Vaccination of domestic
pig with recombinant paramyosin against Schistosoma japonicum in China. Vaccine 18: 2142–2146, 2000.
172. Cheney RE, Riley MA, and Mooseker MS. Phylogenetic analysis of the myosin superfamily. Cell Motil Cytoskeleton 24: 215–
223, 1993.
173. Chiba S, Awazu S, Itoh M, Chin-Bow ST, Satoh N, Satou Y,
and Hastings KEM. A genomewide survey of developmentally
relevant genes in Ciona intestinalis. IX. Genes for muscle structural proteins. Dev Genes Evol 213: 291–302, 2003.
174. Chiba S and Nishikata T. Genes of the ascidian: an annotated
list as of 1997. Zool Sci 15: 625– 643, 1998.
175. Chiba S, Ojima T, and Nishita K. Absence of troponin in foot
muscle of surf clam Pseudocardium sachalinensis. Nippon Suisan Gakk 58: 1919 –1923, 1992.
176. Chiba S, Satou Y, Nishikata T, and Satoh N. Isolation and
characterization of cDNA clones for epidermis specific and muscle specific genes in Ciona savignyi embryos. Zool Sci 15: 239 –
246, 1998.
177. Cho JH, Oh YS, Park KW, Yu JR, Choi KY, Shin JY, Kim DH,
Park WJ, Hamada T, Kagawa H, Maryon EB, Bandyopadhyay
J, and Ahnn J. Calsequestrin, a calcium sequestering protein
localized at the sarcoplasmic reticulum, is not essential for body
wall muscle function in Caenorhabditis elegans. J Cell Sci 113:
3947–3958, 2000.
178. Christova P, Cox JA, and Craescu CT. Ion induced conformational and stability changes in Nereis sarcoplasmic calcium binding protein: evidence that the apo state is a molten globule.
Proteins 40: 177–184, 2000.
179. Chu KH, Wong SH, and Leung PSC. Tropomyosin is the major
mollusk allergen: reverse transcriptase polymerase chain reaction, expression, and IgE reactivity. Mar Biotechnol 2: 499 –509,
2000.
180. Chun M and Falkenthal S. Ifm(3)(2)2 is a myosin heavy chain
allele that disrupts myofibrillar assembly only in the indirect flight
muscle of Drosophila melanogaster. J Cell Biol 107: 2613–2621,
1988.
181. Clark AG, Leicht BG, and Muse SV. Length variation and
secondary structure of introns in the Mlc1 gene in six species of
Drosophila. Mol Biol Evol 13: 471– 482, 1996.
182. Clark KA, McElhinny AS, Beckerle MC, and Gregorio CC.
Striated muscle cytoarchitecture: an intricate web of form and
function. Annu Rev Cell Dev Biol 18: 637–706, 2002.
183. Clark KA, McGrail M, and Beckerle MC. Analysis of PINCH
function in Drosophila demonstrates its requirement in integrindependent cellular processes. Development 130: 2611–2621, 2003.
184. Cleto CL, Vandenberghe AE, MacLean DW, Pannunzio P,
Tortorelli C, Meedel TH, Satou Y, Satoh N, and Hastings
KEM. Ascidian larva reveals ancient origin of vertebrate skeletal
muscle troponin I characteristics in chordate locomotory muscle.
Mol Biol Evol 20: 2113–2122, 2003.
Physiol Rev • VOL
1033
185. Collier VL, Kronert WA, O’Donnell PT, Edwards KA, and
Bernstein SI. Alternative myosin hinge regions are utilized in a
tissue specific fashion that correlates with muscle contraction
speed. Genes Dev 4: 885– 895, 1990.
186. Collins CA and Morgan JE. Duchenne’s muscular dystrophy:
animal models used to investigate pathogenesis and develop therapeutic strategies. Int J Exp Pathol 84: 165–172, 2003.
187. Collins JH. Myosin light chains and troponin C: structural and
evolutionary relationships revealed by amino acid sequence comparisons. J Muscle Res Cell Motil 12: 3–25, 1991.
188. Collins JH. Structure and evolution of troponin-C and related
proteins. Soc Exp Biol Symp 30: 303–307, 1976.
189. Collins JH, Cox JA, and Theibert JL. Amino acid sequence of
a sarcoplasmic calcium-binding protein from the sandworm Nereis diversicolor. J Biol Chem 263: 15378 –15385, 1988.
190. Collins JH, Jakes R, Kendrick-Jones J, Leszyk J, Barouch W,
Theibert JL, Spiegel J, and Szent-Györgyi AG. Amino acid
sequence of myosin essential light chain from the scallop
Aquipecten irradians. Biochemistry 25: 7651–7656, 1986.
191. Collins JH, Johnson JD, and Szent-Györgyi AG. Purification
and characterization of a scallop sarcoplasmic calcium-binding
protein. Biochemistry 22: 341–345, 1983.
192. Collins JH, Theibert JL, Francois JM, Ashley CC, and Potter
JD. Amino acid sequences and Ca2⫹-binding properties of two
isoforms of barnacle troponin C. Biochemistry 30: 702–707, 1991.
193. Conraths FJ, Harnett HW, Worms MJ, and Parkhouse RME.
Immunological cross-reaction between an Onchocerca paramyosin-like molecule and a microfilaria surface antigen. Trop Med
Parasitol 43: 135–138, 1992.
194. Cook WJ, Babu YS, and Cox JA. Crystallization and preliminary
X-ray investigation of a sarcoplasmic calcium-binding protein
from amphioxus. J Mol Biol 221: 1071–1073, 1991.
195. Cook WJ, Ealick SE, Babu YS, Cox JA, and Vijaykumar S.
Three-dimensional structure of a sarcoplasmic calcium-binding
protein from Nereis diversicolor. J Biol Chem 266: 652– 656, 1991.
196. Cook WJ, Jeffrey LC, Cox JA, and Vijaykumar S. Structure of
a sarcoplasmic calcium-binding protein from amphioxus refined
at 2.4 Å resolution. J Mol Biol 229: 461– 471, 1993.
197. Cooley LB, Johnson WH, and Krause S. Phosphorylation of
paramyosin and its possible role in the catch mechanism. J Biol
Chem 254: 2195–2198, 1979.
198. Cooley LB and Krause S. pH titrations of molluscan paramyosin
at two different ionic strengths. Biophys J 32: 755–766, 1980.
199. Cooper AD and Crain WR Jr. Complete nucleotide sequence of
a sea urchin actin gene. Nucleic Acids Res 10: 4081– 4092, 1982.
200. Cope M, Jamie TV, Whisstock J, Rayment I, and KendrickJones J. Conservation within the myosin motor domain: implications for structure and function. Structure 4: 969 –987, 1996.
201. Correa-Oliveira R, Pearce EJ, Oliveira GC, Golgher DB, Katz
N, Bahia LG, Carvalho OS, Gazzinelli G, and Sher A. The
human immune response to defined immunogens of Schistosoma
mansoni: elevated antibody levels to paramyosin in stool negative
individuals from two endemic areas in Brazil. Trans R Soc Trop
Med Hyg 83: 798 – 804, 1989.
202. Costello WJ and Govind CK. Contractile proteins of fast and
slow fibers during differentiation of lobster claw muscle. Dev Biol
104: 434 – 440, 1984.
203. Cotton JLS and Mykles DL. Cloning of a crustacean myosin
heavy chain isoform-exclusive expression in fast muscle. J Exp
Zool 267: 578 –586, 1993.
204. Courchesne-Smith CL and Tobin SL. Tissue specific expression of the 79B actin gene during Drosophila development. Dev
Biol 133: 313–321, 1989.
205. Cowgill RW. Proteolysis of paramyosin from Mercenaria mercenaria and properties of its most stable segment. Biochemistry 14:
503–509, 1975.
206. Cowgill RW. Segments of paramyosin formed by cleavage at sites
of cysteine residues. Biochemistry 14: 4277– 4279, 1975.
207. Cowgill RW. Location and properties of sulfhydryl groups on the
muscle protein paramyosin from Mercenaria mercenaria. Biochemistry 13: 2467–2474, 1974.
85 • JULY 2005 •
www.prv.org
1034
SCOTT L. HOOPER AND JEFFREY B. THUMA
208. Cowgill RW. Susceptibility of paramyosin to proteolysis and its
relationship to regions of different stability. Biochemistry 11:
4532– 4539, 1972.
209. Cox JA. Isolation and characterization of a new Mr 18000 protein
with calcium vector properties in amphioxus muscle and identification of its endogenous target protein. J Biol Chem 261: 13173–
13178, 1986.
210. Cox JA. Intracellular calcium-binding proteins in signal transduction. Chimia 46: 159 –161, 1992.
211. Cox JA, Alard P, and Schaad O. Comparative molecular modeling of amphioxus calcium vector protein with calmodulin and
troponin C. Protein Eng 4: 23–32, 1990.
212. Cox JA, Kretsinger RH, and Stein EA. Sarcoplasmic calciumbinding proteins in insect muscle isolation and properties of locust calmodulin. Biochim Biophys Acta 670: 441– 444, 1981.
213. Cox JA and Stein EA. Characterization of a new sarcoplasmic
calcium binding protein with magnesium induced cooperativity in
the binding of calcium. Biochemistry 20: 5430 –5436, 1981.
214. Cox JA, Wnuk W, and Stein EA. Isolation and properties of a
sarcoplasmic calcium-binding protein from crayfish. Biochemistry 15: 2613–2618, 1976.
215. Cox KH, Angerer LM, Lee JJ, Davidson EH, and Angerer RC.
Cell lineage-specific programs of expression of multiple actin
genes during sea urchin embryogenesis. J Mol Biol 188: 159 –172,
1985.
216. Craescu CT, Prêcheur B, van Riel A, Sakamoto H, Cox JA,
and Engelborghs Y. 1H and 15N resonance assignment of the
calcium-bound form of the Nereis diversicolor sarcoplasmic
Ca2⫹-binding protein. J Biomol NMR 12: 565–566, 1998.
217. Craig R, Szent-Györgyi AG, Beese L, Flicker P, Vibert P, and
Cohen C. Electron microscopy of thin filaments decorated with a
Ca2⫹-regulated myosin. J Mol Biol 140: 35–55, 1980.
218. Crain WR Jr, Boshar MF, Cooper AD, Durica DS, Nagy A, and
Steffen D. The sequence of a sea urchin muscle actin gene
suggests a gene conversion with a cytoskeletal actin gene. J Mol
Evol 25: 37– 45, 1987.
219. Cram EJ, Clark SG, and Schwarzbauer JE. Talin loss-of-function uncovers roles in cell contractility and migration in C. elegans. J Cell Sci 116: 3871–3878, 2003.
220. Crampton AL, Miller C, Baxter GD, and Barker SC. Expressed sequenced tags and new genes from the cattle tick
Boophilus microplus. Exp Appl Acarol 22: 177–186, 1998.
221. Crimmins DL and Thoma RS. Chromatographic analysis of tropomyosins from rabbit skeletal, chicken gizzard, and earthworm
muscle. J Chromatogr 599: 51– 63, 1992.
222. Cripps RM, Ball E, Stark M, Lawn A, and Sparrow JC. Recovery of dominant, autosomal flightless mutants of Drosophila
melanogaster and identification of a new gene required for normal
muscle structure and function. Genetics 137: 151–164, 1994.
223. Cripps RM, Becker KD, Mardahl M, Kronert WA, Hodges D,
and Bernstein SI. Transformation of Drosophila melanogaster
with the wild-type myosin heavy-chain gene: rescue of mutant
phenotypes and analysis of defects caused by overexpression.
J Cell Biol 126: 689 – 699, 1994.
224. Cripps RM and Sparrow JC. Polymorphism in a Drosophila
indirect flight muscle-specific tropomyosin isozyme does not affect flight ability. Biochem Genet 30: 159 –168, 1992.
225. Csizmadia AM, Bonet-Kerrache A, Nyitray L, and Mornet D.
Purification and properties of caldesmon-like protein from molluscan smooth muscle. Comp Biochem Physiol B Biochem 108:
59 – 63, 1994.
226. Culetto E and Sattelle DB. A role for Caenorhabditis elegans in
understanding the function and interactions of human disease
genes. Human Mol Genet 9: 869 – 877, 2000.
227. Cummins C and Anderson P. Regulatory myosin light-chain
genes of Caenorhabditis elegans. Mol Cell Biol 8: 5339 –5349,
1988.
228. D’Haese J and Carlhoff D. Localization and histochemical characterization of myosin isoforms in earthworm body wall muscle. J
Comp Physiol B Biochem Syst Environ Physiol 157: 171–179,
1987.
229. Da Silva CMD, Ferreira HB, Picón M, Gorfinkiel N, Ehrlich
R, and Zaha A. Molecular cloning and characterization of actin
Physiol Rev • VOL
230.
231.
232.
233.
234.
235.
236.
237.
238.
239.
240.
241.
242.
243.
244.
245.
246.
247.
248.
249.
250.
genes from Echinococcus granulosus. Mol Biochem Parasitol 60:
209 –220, 1993.
Daguin C, Bonhomme F, and Borsa P. The zone of sympatry
and hybridization of Mytilus edulis and M. galloprovincialis, as
described by intron length polymorphism at locus mac-1. Heredity 86: 342–354, 2001.
Dahmen A, Gallin M, Schumacher M, and Erttmann KD.
Molecular cloning and pre-mRNA maturation of Onchocerca volvulus paramyosin. Mol Biochem Parasitol 57: 335–338, 1993.
Daley J, Southgate R, and Ayme-Southgate A. Structure of the
Drosophila projectin protein: isoforms and implication for projectin filament assembly. J Mol Biol 279: 201–210, 1998.
Daul CB, Slattery M, Reese G, and Lehrer SB. Identification of
the major brown shrimp (Penaeus aztecus) allergen as the muscle
protein tropomyosin. Int Arch Allergy Immunol 105: 49 –55, 1994.
Davis AH, Blanton R, and Klich P. Stage and sex specific
differences in actin gene expression in Schistosoma mansoni.
Mol Biochem Parasitol 17: 289 –298, 1985.
Davis MB, Dietz J, Standiford DM, and Emerson CP. Transposable element insertions respecify alternative exon splicing in
three Drosophila myosin heavy chain mutants. Genetics 150:
1105–1114, 1998.
Deak II. Mutations of Drosophila melanogaster that affect muscles. J Embryol Exp Morphol 40: 35– 63, 1977.
Deak II, Bellamy PR, Bienz M, Dubuis Y, Fenner E, Gollin M,
Rahmi A, Ramp T, Reinhardt CA, and Cotton B. Mutations
affecting the indirect flight muscles of Drosophila melanogaster.
J Embryol Exp Morphol 69: 61– 81, 1982.
De Couet HG, Mazander KD, and Gröschel-Stewart U. A
study of invertebrate actins by isoelectric focusing and immunodiffusion. Experientia 36: 404 – 405, 1980.
De Eguileor M, Cotelli F, Valvassori R, Brivio M, and Di
Lernia L. Functional significance of intermediate filament meshwork in annelid helical muscles. J Ultrastruct Mol Struct Res 100:
183–193, 1988.
Deitiker PR and Epstein HF. Thick filament substructures in
Caenorhabditis elegans: evidence for two populations of
paramyosin. J Cell Biol 123: 303–311, 1993.
De Jesus AR, Araujo I, Bacellar O, Magalhaes A, Pearce E,
Harn D, Strand M, and Carvalho EM. Human immune responses to Schistosoma mansoni vaccine candidate antigens.
Infect Immun 68: 2797–2803, 2000.
Dekeyzer N, Engelborghs Y, and Volckaert G. Cloning, expression, and purification of a sarcoplasmic calcium-binding protein
from the sandworm Nereis diversicolor via a fusion product with
chloramphenicol acetyltransferase. Protein Eng 7: 125–130, 1994.
Dekkers LC, van der Plas MK, van Loenen PB, de dunnen JT,
van Ommen GJB, Fradkin LG, and Noordermeer JN. Embryonic expression patterns of the Drosophila dystropin-associated
glycoprotein complex orthologs. Gene Exp Patterns 4: 153–159,
2004.
De la Pompa JL. Functional relationships between genes of the
shaker gene complex of Drosophila. Mol Gen Genet 244: 197–204,
1994.
Deng JS, Gold D, LoVerde PT, and Fishelson Z. Inhibition of
the complement membrane attack complex by Schistosoma mansoni paramyosin. Infect Immun 71: 6402– 6410, 2003.
DesGroseillers L, Auclair D, and Wickham L. Nucleotide sequence of an actin cDNA gene from Aplysia californica. Nucleic
Acids Res 18: 3654 –3654, 1990.
De Villafranca GW. Some physico-chemical properties of myosin
B from the horseshoe crab, Limulus polyphemus. Comp Biochem
Physiol 26: 443– 454, 1968.
De Villafranca GW and Ena Haines V. Paramyosin from arthropod cross-striated muscle. Comp Biochem Physiol 47: 9 –26, 1974.
Dey CS, Deitiker PR, and Epstein HF. Assembly dependent
phosphorylation of myosin and paramyosin of native thick filaments in Caenorhabditis elegans. Biochem Biophys Res Commun 186: 1528 –1532, 1992.
Dibb NJ, Brown DM, Karn J, Moerman DG, Bolten SL, and
Waterston RH. Sequence analysis of mutations that affect the
synthesis, assembly, and enzymatic activity of the unc-54 myosin
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
251.
252.
253.
254.
255.
256.
257.
258.
259.
260.
261.
262.
263.
264.
265.
266.
267.
268.
269.
270.
heavy chain of Caenorhabditis elegans. J Mol Biol 183: 543–551,
1985.
Dibb NJ, Maruyama IN, Krause M, and Karn J. Sequence
analysis of the complete Caenorhabditis elegans myosin heavy
chain gene family. J Mol Biol 205: 603– 613, 1989.
Dickinson MH, Hyatt CJ, Lehmann FO, Moore JR, Reedy
MC, Simcox A, Tohtong R, Vigoreaux JO, Yamashita H, and
Maughan DW. Phosphorylation dependent power output of
transgenic flies: an integrated study. Biophys J 73: 3122–3134,
1997.
Ding L and Candido EPM. HSP25, a small heat shock protein
associated with dense bodies and M lines of body wall muscle in
Caenorhabditis elegans. J Biol Chem 275: 9510 –9517, 2000.
Dissanayake S, Xu M, and Piessens WF. Myosin heavy chain is
a dominant parasite antigen recognized by antibodies in sera from
donors with filarial infections. Mol Biochem Parsitol 56: 349 –351,
1992.
Dissous C, Torpier G, Duvaux-Miret O, and Capron A. Structural homology of tropomyosins from the human trematode
Schistosoma mansoni and its intermediate host Biomphalaria
glabrata. Mol Biochem Parasitol 43: 245–255, 1990.
Dodement H, Riemer D, and Weber K. Structure of an invertebrate gene encoding cytoplasmic intermediate filament (IF) proteins: implications for the origin and the diversification of IF
proteins. EMBO J 9: 4083– 4094, 1990.
Domingo A, Gonzalez-Jurado J, Maroto M, Diaz C, Vinós J,
Carrasco C, Cervera M, and Marco R. Troponin T is a calciumbinding protein in insect muscle: in vivo phosphorylation, musclespecific isoforms, and developmental profile in Drosophila melanogaster. J Muscle Res Cell Motil 19: 393– 403, 1998.
Drouin G. Testing claims of gene conversion between multigene
family members: examples from echinoderm actin genes. J Mol
Evol 54: 138 –139, 2002.
Drummond D, Hennessey E, and Sparrow J. The binding of
mutant actins to profilin, ATP, and DNase I. Eur J Biochem 209:
171–179, 1992.
Drummond DR, Hennessey ES, and Sparrow JC. Characterization of missense mutations in the Act88F gene of Drosophila
melanogaster. Mol Gen Genet 226: 70 – 80, 1991.
Drummond DR, Hennessey ES, and Sparrow JC. Stability of
mutant actins. Biochem J 274: 301–303, 1991.
Drummond DR, Peckham M, Sparrow JC, and White DC.
Alteration in crossbridge kinetics caused by mutations in actin.
Nature 348: 440 – 442, 1990.
Drysdale R, Rushton E, and Bate M. Genes required for embryonic muscle development in Drosophila melanogaster. A survey of the X-chromosome. Roux’s Arch Dev Biol 202: 276 –295,
1993.
Dubreuil RR and Wang P. Genetic analysis of the requirements
for alpha-actinin function. J Muscle Res Cell Motil 21: 705–713,
2000.
Durica DS, Schloss JA, and Crain WR Jr. Organization of actin
gene sequences in the sea urchin: molecular cloning of an introncontaining DNA sequence coding for a cytoplasmic actin. Proc
Natl Acad Sci USA 77: 5683–5687, 1980.
Durussel I, Luan-Rilliet Y, Petrova T, Takagi T, and Cox JA.
Cation binding and conformation of tryptic fragments of Nereis
sarcoplasmic calcium-binding protein: calcium-induced homoand heterodimerization. Biochemistry 32: 2394 –2400, 1993.
Dyer WJ and Hiltz DE. Comparative quality of fresh and of
frozen and thawed scallop meats and post-thaw keeping quality.
Nippon Suisan Gakk 40: 235–243, 1974.
Eberl M, Mountford AP, Jankovic D, and Beck E. Isolation of
T cell antigens by using a recombinant protein library and its
application to the identification of novel vaccine candidates
against schistosomiasis. Infect Immun 67: 3383–3389, 1999.
Ehara T, Nakagawa K, Tamiya T, Noguchi SF, and Tsuchiya
T. Effect of paramyosin on invertebrate natural actomyosin gel
formation. Fish Sci 70: 306 –313, 2004.
Elfvin M, Levine RJC, and Dewey MM. Paramyosin in invertebrate muscles. I. Identification and localization. J Cell Biol 71:
261–272, 1976.
Physiol Rev • VOL
1035
271. Endo T, Matsumoto K, Hama T, Ohtsuka Y, Katsura G, and
Obinata T. Temporal and spatial expression of distinct troponin
T genes in embryonic/larval tail striated muscle and adult body
wall smooth muscle of ascidian. Cell Struct Funct 22: 197–203,
1997.
272. Endo T, Matsumoto K, Hama T, Ohtsuka Y, Katsura G, and
Obinata T. Distinct troponin T genes are expressed in embryonic/
larval tail striated muscle and adult body wall smooth muscle of
ascidian. J Biol Chem 271: 27855–27862, 1996.
273. Endo T and Obinata T. Troponin and its components from
ascidian smooth muscle. J Biochem 89: 1599 –1608, 1981.
274. Engelborghs Y, Mertens K, Willaert K, Luanrilliet Y, and Cox
JA. Kinetics of conformational changes in Nereis sarcoplasmic
calcium-binding protein upon binding of divalent ions. J Biol
Chem 265: 18809 –18815, 1990.
275. Epstein HF. Genetic analysis of myosin assembly in Caenorhabditis elegans. Mol Neurobiol 4: 1–25, 1990.
276. Epstein HF and Bernstein SI. Genetic approaches to understanding muscle development. Dev Biol 154: 231–244, 1992.
277. Epstein HF, Ortiz CL, and Berliner GC. Assemblages of multiple thick filaments in nematode mutants. J Muscle Res Cell Motil
8: 527–536, 1987.
278. Epstein HF, Ortiz I, and Mackinnon LA. The alteration of
myosin isoform compartmentation in specific mutants of Caenorhabditis elegans. J Cell Biol 103: 985–993, 1986.
279. Epstein HF and Thomson JN. Temperature-sensitive mutation
affecting myofilament assembly in Caenorhabditis elegans. Nature 250: 579 –580, 1974.
280. Epstein HF, Waterston RH, and Brenner S. A mutant affecting
the heavy chain of myosin in Caenorhabditis elegans. J Mol Biol
90: 291–300, 1974.
281. Ericsson C, Petho Z, and Mehlin H. An on-line two-dimensional
polyacrylamide gel electrophoresis protein database of adult Drosophila melanogaster. Electrophoresis 18: 484 – 490, 1997.
282. Erondu NE and Donelsen JE. Characterization of a mysoin like
antigen from Onchocerca volvulus. Mol Biochem Parasitol 40:
213–224, 1990.
283. Esteves A, Senorale M, and Ehrlich R. A tropomyosin gene is
differentially expressed in the larval stage of Echinococcus granulosus. Parasitol Res 89: 501–502, 2003.
284. Estuningsih SE, Smooker PM, Wiedosari E, Widjajanti S,
Vaiano S, Partoutomo S, and Spithill TW. Evaluation of antigens of Fasciola gigantica as vaccines against tropical fasciolosis
in cattle. Int J Parasitol 27: 1419 –1428, 1997.
285. Fagotti A, Di Rosa I, Simoncelli F, Chaponnier C, Gabbiani
G, and Pascolini R. Actin isoforms in amphioxus Branchiostoma
lanceolatum. Cell Tissue Res 292: 173–176, 1998.
286. Fährmann M, Erfmann M, and Beinbrech G. Binding of
CaMKII to the giant muscle protein projectin: stimulation of
CaMKII activity by projectin. Biochim Biophys Acta 1569: 127–
134, 2002.
287. Fährmann M, Fonk I, and Beinbrech G. The kinase activity of
the giant protein projectin of the flight muscle of Locusta migratoria. Insect Biochem Mol Biol 32: 1401–1407, 2002.
288. Falkenthal S, Graham M, and Wilkinson J. The indirect flight
muscle of Drosophila accumulates a unique myosin alkali light
chain isoform. Dev Biol 121: 263–272, 1987.
289. Falkenthal S, Parker VP, and Davidson N. Developmental
variations in the splicing pattern of transcripts from the Drosophila gene encoding myosin alkali light chain result in different
carboxyl-terminal amino acid sequences. Proc Natl Acad Sci USA
82: 449 – 453, 1985.
290. Falkenthal S, Parker VP, Mattox WW, and Davidson N. Drosophila melanogaster has only one myosin alkali light-chain gene
which encodes a protein with considerable amino acid sequence
homology to chicken myosin alkali light chains. Mol Cell Biol 4:
956 –965, 1984.
291. Fan JJ, Minchella DJ, Day SR, McManus DP, Tiu WU, and
Brindley PJ. Generation, identification, and evaluation of expressed sequence tags from different developmental stages of the
Asian blood fluke Schistosoma japonicum. Biochem Biophys Res
Commun 252: 348 –356, 1998.
85 • JULY 2005 •
www.prv.org
1036
SCOTT L. HOOPER AND JEFFREY B. THUMA
292. Fang H and Brandhorst BP. Evolution of actin gene families of
sea urchins. J Mol Evol 39: 347–356, 1994.
293. Fang H and Brandhorst BP. Expression of the actin gene family
in embryos of the sea urchin Lytechinus pictus. Dev Biol 173:
306 –317, 1996.
294. Ferguson C, Lakey A, Hutchings A, Butcher GW, Leonard
KR, and Bullard B. Cytoskeletal proteins of insect muscle: location of zeelins in Lethocerus flight and leg muscle. J Cell Sci 107:
1115–1129, 1994.
295. Fernandes J, Reshef A, Patton L, Ayuso R, Reese G, and
Lehrer SB. Immunoglobulin E antibody reactivity to the major
shrimp allergen, tropomyosin, in unexposed Orthodox Jews. Clin
Exp Allergy 33: 956 –961, 2003.
296. Fernandez A, Garcia T, Asensio L, Rodriguez MA, Gonzalez
I, Cespedes A, Hernandez PE, and Martin R. Identification of
the clam species Ruditapes decussatus (grooved carpet shell),
Venerupis pullastra (pullet carpet shell), and Ruditapes philippinarum (Japanese carpet shell) by PCR-RFLP. J Agric Food
Chem 48: 3336 –3341, 2000.
297. Ferreira CAS, Barbosa MC, Silveira TCL, Valenzuela JG, Vaz
ID, and Masuda A. cDNA cloning, expression, and characterization of a Boophilus microplus paramyosin. Parasitology 125:
265–274, 2002.
298. Ferrer E, Moyano E, Benitez L, Gonzalez LM, Bryce D, Foster-Cuevas M, Davila I, Cortez MM, Harrison LJS, Parkhouse RME, and Garate T. Cloning and characterization of
Taenia saginata paramyosin cDNA. Parasitol Res 91: 60 – 67,
2003.
299. Ferrús A, Acebes A, Marı́n MC, and Hernández-Hernández
A. A genetic approach to detect muscle protein interactions in
vivo. Trends Cardiovasc Med 10: 293–298, 2000.
300. Files JG, Carr S, and Hirsh D. Actin gene family of Caenorhabditis elegans. J Mol Biol 164: 355–375, 1983.
301. Fire A, Albertson D, Harrison SW, and Moerman DG. Production of antisense RNA leads to effective and specific inhibition
of gene expression in C. elegans muscle. Development 113: 503–
514, 1991.
302. Fire A and Waterston RH. Proper expression of myosin genes in
transgenic nematodes. EMBO J 8: 3419 –3428, 1989.
303. Firtel RA. Multigene families encoding actin and tubulin. Cell 24:
6 –7, 1981.
304. Fisher DA and Bode HR. Nucleotide sequence of an actin encoding gene from Hydra attenuata: structural characteristics and
evolutionary implications. Gene 84: 55– 64, 1989.
305. Fitzhugh GH and Marden JH. Maturational changes in troponin
T expression Ca2⫹-sensitivity and twitch contraction kinetics in
dragonfly flight muscle. J Exp Biol 200: 1473–1482, 1997.
306. Flaherty DB, Gernert KM, Shmeleva N, Tang X, Mercer KB,
Borodovsky M, and Benian GM. Titins in C. elegans with unusual features: coiled-coil domains, novel regulation of kinase
activity, and two new possible elastic regions. J Mol Biol 323:
533–549, 2002.
307. Flanigan TP, King CH, Lett RR, Nanduri J, and Mahmoud
AAF. Induction of resistance to Schistosoma mansoni infection
in mice by purified parasite paramyosin. J Clin Invest 83: 1010 –
1014, 1989.
308. Flick GJ and Lovell RT. Post-mortem biochemical changes in
the muscle of the gulf shrimp, Penaeus aztecus. J Food Sci 37:
609 – 611, 1972.
309. Flicker PF, Wallimann T, and Vibert P. Electron microscopy of
scallop myosin—location of regulatory light chains. J Mol Biol
169: 723–741, 1983.
310. Flood PR. Fine structure of the notochord of amphioxus. Symp
Zool Soc Lond 36: 81–104, 1975.
311. Flood PR, Guthrie DM, and Banks JR. Paramyosin muscle in
the notochord of amphioxus. Nature 222: 87– 88, 1969.
312. Fock U and Hinssen H. Nebulin is a thin filament protein of the
cardiac muscle of the agnathans. J Muscle Res Cell Motil 23:
205–213, 2002.
313. Fock U and Hinssen H. Identification and localisation of nebulin
as a thin filament component of invertebrate chordate muscles.
J Comp Physiol B Biochem Syst Environ Physiol 169: 555–560,
1999.
Physiol Rev • VOL
314. Fong S, Hamill SJ, Proctor M, Freund SMV, Benian GM,
Chothia C, Bycroft M, and Clarke J. Structure and stability of
an immunoglobulin superfamily domain from twitchin, a muscle
protein of the nematode Caenorhabditis elegans. J Mol Biol 264:
624 – 639, 1996.
315. Fonseca CT, Cunha-Neto E, Kalil J, de Jesus AR, CorreaOliveira R, Carvalho E, and Oliveira S. Identification of immunodominant epitopes of Schistosoma mansoni vaccine candidate
antigens using human T cells. Mem Inst Oswaldo Cruz 99: 63– 66,
2004.
316. Francis R and Waterston RH. Muscle organization in Caenohabditis elegans: localization implicated in thin filament attachment and I band organization. J Cell Biol 101: 1532–1549, 1985.
317. Francis R and Waterston RH. Muscle cell attachment in Caenorhabditis elegans. J Cell Biol 114: 465– 479, 1991.
318. Frenkel MJ, Savin KW, Bakker RE, and Ward CW. Characterization of cDNA clones coding for muscle tropomyosin of the
nematode Trichostrongylus colubriformis. Mol Biochem Parasitol 37: 191–199, 1989.
319. Fuchs E and Weber K. Intermediate filaments: structure, dynamics, function, and disease. Annu Rev Biochem 63: 345–382, 1994.
320. Fujinoki M, Tomiyama T, and Ishimoda-Takagi T. Tropomyosin isoforms present in the sea anemone, Anthopleura japonica
(Anthozoa, Cnidaria). J Exp Zool 293: 649 – 663, 2002.
321. Fukuda I, Imagawa S, Iwao K, Horiguchi T, and Watanabe T.
Isolation of actin-encoding cDNAs from symbiotic corals. DNA
Res 9: 217–223, 2002.
322. Fukuzawa A, Hiroshima M, Maruyama K, Yonezawa N, Tokunaga M, and Kimura S. Single-molecule measurement of elasticity 449 of serine, glutamate, and lysine-rich repeats of invertebrate connectin reveals that its elasticity is caused entropicially
by random coil structure. J Muscle Res Cell Motil 23: 449 – 453,
2002.
323. Fukuzawa A, Shimamura J, Takemori S, Kanzawa N,
Yamaguchi M, Sun P, Maruyama K, and Kimura S. Invertebrate connectin spans as much as 3.5 micrometers in the giant
sarcomeres of crayfish claw muscle. EMBO J 20: 4826 – 4835, 2001.
324. Funabara D, Kinoshita S, Watabe S, Siegman MJ, Butler TM,
and Hartshorne DJ. Phosphorylation of molluscan twitchin by
the cAMP-dependent protein kinase. Biochemistry 40: 2087–2095,
2001.
325. Funabara D, Nakaya M, and Watabe S. Isolation and characterization of a novel 45 kDa calponin-like protein from anterior
byssus retractor muscle of the mussel Mytilus galloprovincialis.
Fish Sci 67: 511–517, 2001.
326. Funabara D, Watabe S, Mooers SU, Narayan S, Dudas C,
Hartshorne DJ, Siegman MJ, and Butler TM. Twitchin from
molluscan catch muscle-primary structure and relationship between site-specific phosphorylation and mechanical function.
J Biol Chem 278: 29308 –29316, 2003.
327. Fyrberg C, Ketchum A, Ball E, and Fyrberg E. Characterization of lethal Drosophila melanogaster ␣-actinin mutants. Biochem Genet 36: 299 –310, 1998.
328. Fyrberg C, Parker H, Hutchison B, and Fyrberg E. Drosophila
melanogaster genes encoding three troponin C isoforms and a
calmodulin-related protein. Biochem Genet 32: 119 –135, 1994.
329. Fyrberg CC, Labeit S, Bullard B, Leonard K, and Fyrberg E.
Drosophila projectin: relatedness to titin and twitchin and correlation with lethal(4) 102 Cda and bent-dominant mutants. Proc R
Soc Lond B Biol Sci 249: 33– 40, 1992.
330. Fyrberg EA. Study of contractile and cytoskeletal proteins using
Drosophila genetics. Cell Motil Cytoskeleton 14: 118 –127, 1989.
331. Fyrberg EA and Beall C. Genetic approaches to myofibril form
and function in Drosophila. Trends Genet 6: 126 –131, 1990.
332. Fyrberg EA, Beall C, and Fyrberg CC. From genes to tensile
forces— genetic dissection of contractile protein assembly and
function in Drosophila melanogaster. J Cell Sci Suppl 14: 27–29,
1991.
333. Fyrberg EA, Bernstein SI, and Raghavan KV. Basic methods
for Drosophila muscle biology. Methods Cell Biol 44: 237–258,
1994.
334. Fyrberg EA, Bond BJ, Hershey ND, Mixter KS, and Davidson
N. The actin genes of Drosophila: protein coding regions are
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
335.
336.
337.
338.
339.
340.
341.
342.
343.
344.
345.
346.
347.
348.
349.
350.
351.
352.
353.
354.
355.
highly conserved but intron positions are not. Cell 24: 107–116,
1981.
Fyrberg EA, Fyrberg CC, Beall C, and Saville DL. Drosophila
melanogaster troponin T mutations engender three distinct syndromes of myofibrillar abnormalities. J Mol Biol 216: 657– 675,
1990.
Fyrberg EA, Fyrberg CC, Biggs JR, Saville D, Beall CJ, and
Ketchum A. Functional nonequivalence of Drosophila actin isoforms. Biochem Genet 36: 271–287, 1998.
Fyrberg EA and Karlik CC. Genetic rescue of muscle defects
associated with a mutant Drosophila melanogaster tropomyosin
allele. Mol Cell Biol 7: 2977–2980, 1987.
Fyrberg EA, Kelly M, Ball E, Fyrberg C, and Reedy MC.
Molecular genetics of Drosophila alpha-actinin: mutant alleles
disrupt Z disc integrity and muscle insertions. J Cell Biol 110:
1999 –2011, 1990.
Fyrberg EA, Kindle KL, and Davidson N. The actin genes of
Drosophila: a dispersed multigene family. Cell 19: 365–378, 1980.
Fyrberg EA, Mahaffey JW, Bond BJ, and Davidson N. Transcripts of the six Drosophila actin genes accumulate in a stageand tissue-specific manner. Cell 33: 115–123, 1983.
Galkin VE, Orlova A, Lukoyanova N, VanLoock MS, Haag P,
Bullard B, and Egelman EH. The location of ubiquitin in Lethocerus arthrin. J Mol Biol 325: 623– 628, 2003.
Garcia R, Pazaliaga B, Ernst SG, and Crain WR. Three sea
urchin actin genes show different patterns of expression-muscle
specific, embryo specific, and constitutive. Mol Cell Biol 4: 840 –
845, 1984.
Garone L, Theibert JL, Miegel A, Maeda Y, Murphy C, and
Collins JH. Lobster troponin C: amino acid sequences of three
isoforms. Arch Biochem Biophys 291: 89 –91, 1991.
Gazarian KG, Gazarian TG, Solı́s CF, Hernández R, Shoemaker CB, and Laclette JP. Epitope mapping on N-terminal
region of Taenia solium paramyosin. Immunol Lett 72: 191–195,
2000.
Geisler N, Schünemann J, Weber K, Häner M, and Aebi U.
Assembly and architecture of invertebrate cytoplasmic intermediate filaments reconcile features of vertebrate cytoplasmic and
nuclear lamin-type intermediate filaments. J Mol Biol 282: 601–
617, 1998.
Gengyo-Ando K and Kagawa H. Single charge change on the
helical surface of the paramyosin rod dramatically disrupts thick
filament assembly in Caenorhabditis elegans. J Mol Biol 219:
429 – 441, 1991.
George EL, Ober MB, and Emerson CP Jr. Functional domains
of the Drosophila melanogaster muscle myosin heavy chain gene
are encoded by alternatively spliced exons. Mol Cell Biol 9: 2957–
2974, 1989.
Gerday C. Soluble calcium-binding proteins from fish and invertebrate muscle. Mol Physiol 2: 63– 87, 1982.
Gettner SN, Kenyon C, and Reichardt LF. Characterization of
␤pat-3 heterodimers, a family of essential integrin receptors in C.
elegans. J Cell Biol 129: 1127–1141, 1995.
Geyer PK and Fyrberg EA. 5⬘-Flanking sequence required for
regulated expression of a muscle-specific Drosophila melanogaster actin gene. Mol Cell Biol 6: 3388 –3396, 1986.
Gieseler K, Abdel-Dayem M, and Segalat L. In vitro interactions of Caenorhabditis elegans dystrophin with dystrobrevin and
syntrophin. FEBS Lett 461: 59 – 62, 1999.
Gieseler K, Bessou C, and Segalat L. Dystrobrevin and dystrophin-like mutants display similar phenotypes in the nematode
Caenorhabditis elegans. Neurogenetics 2: 87–90, 1999.
Gieseler K, Grisoni K, Mariol MC, and Segalat L. Overexpression of dystrobrevin delays locomotion defects and muscle degeneration in a dystrophin-deficient Caenorhabditis elegans. Neuromuscular Disorders 12: 371–377, 2002.
Gieseler K, Grisoni K, and Segalat L. Genetic suppression of
phenotypes arising from mutations in dystrophin-related genes in
Caenorhabditis elegans. Curr Biol 10: 1092–1097, 2000.
Gieseler K, Mariol MC, Bessou C, Migaud M, Franks CJ,
Holden-Dye L, and Segalat L. Molecular, genetic, and physiological characterisation of dystrobrevin-like (dyb-1) mutants of
Caenorhabditis elegans. J Mol Biol 307: 107–117, 2001.
Physiol Rev • VOL
1037
356. Giugia JB, Gieseler K, Arpagaus M, and Segalat L. Mutations
in the dystrophin-like dys-1 gene of Caenorhabditis elegans result
in reduced acetylcholinesterase activity. FEBS Lett 463: 270 –272,
1999.
357. Gobert GN. Immunolocalization of schistosome proteins. Microsc Res Tech 42: 176 –185, 1998.
358. Gobert GN. The role of microscopy in the investigation of
paramyosin as a vaccine candidate against Schistosoma japonicum. Parasitol Today 14: 115–118, 1998.
359. Gobert GN, Stenzel DJ, Jones MK, Allen DE, and McManus
DP. Schistosoma japonicum: immunolocalization of paramyosin
during development. Parasitology 114: 45–52, 1997.
360. Goetinck S and Waterston RH. The Caenorhabditis elegans
unc-87 protein is essential for maintenance, but not assembly, of
bodywall muscle. J Cell Biol 127: 71–78, 1994.
361. Goetinck S and Waterston RH. The Caenorhabditis elegans
muscle affecting gene unc-87 encodes a novel thin filament associated protein. J Cell Biol 127: 79 –93, 1994.
362. Goetz DW and Whisman BA. Occupational asthma in a seafood
restaurant worker: cross-reactivity of shrimp and scallops. Ann
Allergy Asthma Immunol 85: 461– 466, 2000.
363. Goldberg A and Lehman W. Troponin-like proteins from muscles of the scallop, Aequipecten irradians. Biochem J 171: 413–
418, 1978.
364. Goldman RD. Worms reveal essential functions for intermediate
filaments. Proc Natl Acad Sci USA 98: 7659 –7661, 2001.
365. Goldstein LSB and Gunawardena S. Flying through the Drosophila cytoskeletal genome. J Cell Biol 150: F63–F68, 2000.
366. Gomez-Chiarri M, Kirby VL, and Powers DA. Isolation and
characterization of an actin promoter from the red abalone (Haliotis rufescens). Mar Biotechnol 1: 269 –278, 1999.
367. Gómez-Guillén MC, Borderı́as AJ, and Montero P. Salt, nonmuscle proteins, and hydrocolloids affecting rigidity changes during gelation of giant squid (Dosidicus gigas). J Agric Food Chem
45: 616 – 621, 1997.
368. Gómez-Guillén MC, Hurtado JL, and Montero P. Autolysis
and protease inhibition effects on dynamic viscoelastic properties
during thermal gelation of squid muscle. J Food Sci 67: 2491–2496,
2002.
369. Gómez-Guillén MC and Montero P. Improvement of giant squid
(Dosidicus gigas) muscle gelation by using gelling ingredients. Z
Lebensm Unters Forsch 204: 379 –384, 1997.
370. Gómez-Guillén MC, Montero P, Solas MT, and Borderı́as AJ.
Thermally induced aggregation of giant squid (Dosidicus gigas)
mantle proteins. Physicochemical contribution of added ingredients. J Agric Food Chem 46: 3440 –3446, 1998.
371. Gómez-Guillén MC, Solas MT, Borderı́as AJ, and Montero P.
Ultrastructural and rheological changes during gelation of giant
squid (Dosidicus gigas) muscle. Z Lebensm Unters Forsch 202:
215–220, 1996.
372. Gómez-Guillén MC, Solas MT, Borderias J, and Montero P.
Effect of heating temperature and NaCl concentration on ultrastructure and texture of gels made from giant squid (Dosidicus
gigas) with addition of starch, ␫-carrageenan and egg white. Z
Lebensm Unters Forsch 202: 221–227, 1996.
373. Goodsir J. On the anatomy of Amphioxus lanceolatus. Trans R
Soc Edinburgh 15: 247–263, 1844.
374. Goodson HV and Spudich JA. Molecular evolution of the myosin family: relationships derived from comparisons of amino acid
sequences. Proc Natl Acad Sci USA 90: 659 – 663, 1993.
375. Goodwin EB, Szent-Györgyi AG, and Leinwand LA. Cloning
and characterization of the scallop essential and regulatory myosin light chain cDNAs. J Biol Chem 262: 11052–11056, 1987.
376. Grandea AG III, Tuyen LK, Asikin N, Davis TB, Phillipp M,
Cohen C, and McReynolds LA. A gt11 cDNA recombinant that
encodes Dirofilaria immitis paramyosin. Mol Biochem Parasitol
35: 31– 42, 1989.
377. Greener MJ and Roberts RG. Conservation of components of
the dystrophin complex in Drosophila. FEBS Lett 482: 13–18,
2000.
378. Greenwald IS and Horvitz HR. unc-93(e1500): a behavioral
mutant of Caenorhabditis elegans that defines a gene with wildtype null phenotype. Genetics 96: 147–164, 1980.
85 • JULY 2005 •
www.prv.org
1038
SCOTT L. HOOPER AND JEFFREY B. THUMA
379. Gregorio CC, Granzier H, Sorimachi H, and Labeit S. Muscle
assembly: a titanic achievement? Curr Opin Cell Biol 11: 18 –25,
1999.
380. Gregory SL and Brown NH. kakapo, a gene required for adhesion between and within cell layers in Drosophila, encodes a large
cytoskeletal linker protein related to plectin and dystrophin. J Cell
Biol 143: 1271–1282, 1998.
381. Gremke L, Lord PCW, Sabacan L, Lin SC, Wohlwill A, and
Storti RV. Coordinate regulation of Drosophila tropomyosin
gene expression is controlled by multiple muscle type specific
positive and negative enhancer elements. Dev Biol 159: 513–527,
1993.
382. Grisoni K, Gieseler K, Mariol MC, Martin E, Carre-Pierrat
M, Moulder G, Barstead R, and Segalat L. The stn-1 syntrophin
gene of C. elegans is functionally related to dystrophin and dystrobrevin. J Mol Biol 332: 1037–1046, 2003.
383. Grisoni K, Gieseler K, and Segalat L. Dystrobrevin requires a
dystrophin-binding domain to function in Caenorhabditis elegans.
Eur J Biochem 269: 1607–1612, 2002.
384. Grisoni K, Martin E, Gieseler K, Mariol MC, and Segalat L.
Genetic evidence for a dystrophin-glycoprotein complex (DGC) in
Caenorhabditis elegans. Gene 294: 77– 86, 2002.
385. Gröger H, Callaerts P, Gehring WJ, and Schmid V. Gene
duplication and recruitment of a specific tropomyosin into striated muscle cells in the jellyfish Podocoryne carnea. J Exp Zool
285: 378 –386, 1999.
386. Gröger H, Callaerts P, Jurgen GW, and Volker S. Gene duplication and recruitment of a specific tropomyosin into striated
muscle cells in the jellyfish Podocoryne carnea. J Exp Zool 288:
93–93, 2000.
387. Grossman Z, Ram D, Markovics A, Tarrab-Hazbi R, Lantner
F, Ziv E, and Schechter I. Schistosoma mansoni: stage-specific
expression of muscle specific genes. Exp Parasitol 70: 62–71,
1990.
388. Guimarães PM, Leal-Bertioli SCM, Curtis RH, Davis EL, and
Bertioli DJ. Isolation of two cDNAS encoding a tropomyosin and
an intermediate filament protein from the soybean cyst nematode
Heterodera glycines. Nematropica 33: 87–95, 2003.
389. Gullberg D, Velling T, Lohikangas L, and Tiger CF. Integrins
during muscle development and in muscular dystrophies. Pediatr
Pathol Mol Med 18: 303–327, 1999.
390. Hahn BS and Labouesse M. Tissue integrity: hemidesmosomes
and resistance to stress. Curr Biol 11: R858 –R861, 2001.
391. Hakeda S, Endo S, and Saigo K. Requirements of kettin, a giant
muscle protein highly conserved in overall structure in evolution,
for normal muscle function, viability, and flight activity of Drosophila. J Cell Biol 148: 101–114, 2000.
392. Halsey JF and Harrington WF. Substructure of paramyosin.
Correlation of helix stability, trypsin digestion kinetics, and amino
acid composition. Biochem 12: 693–701, 1971.
393. Hammond KS and Goll DE. Purification of insect myosin and
␣-actinin. Biochem J 151: 189 –192, 1975.
394. Hanke PD and Storti RV. Nucleotide sequence of a cDNA clone
encoding a Drosophila muscle tropomyosin II isoform. Gene 45:
211–214, 1986.
395. Hanke PD and Storti RV. The Drosophila melanogaster tropomyosin II gene produces multiple proteins by use of alternative
tissue-specific promoters and alternative splicing. Mol Cell Biol 8:
3591–3602, 1988.
396. Hanson J and Lowy J. The structure of the muscle fibres in the
translucent part of the adductor of the oyster Crossostrea angulata. Proc R Soc Lond B Biol Sci 154: 173–196, 1961.
397. Hanson J, Lowy J, Huxley HE, Bailey K, Kay CM, and Rüegg
JC. Structure of molluscan tropomyosin. Nature 180: 1134 –1135,
1957.
398. Hapiak V, Hresko MC, Schriefer LA, Saiyasisongkhram K,
Bercher M, and Plenefisch J. mua-6, a gene required for tissue
integrity in Caenorhabditis elegans, encodes a cytoplasmic intermediate filament. Dev Biol 263: 330 –342, 2003.
399. Hardwicke PMD, Wallimann T, and Szent-Györgyi AG. Regulatory and essential light chain interactions in scallop myosin I.
Protection of essential light-chain thiol groups by regulatory lightchains. J Mol Biol 156: 141–152, 1982.
Physiol Rev • VOL
400. Harrington WF and Roger ME. Myosin. Annu Rev Biochem 53:
35–73, 1984.
401. Harris DA, Sherbany AA, and Schwartz JH. Purification and
characterization of muscle proteins from Aplysia californica.
Biol Bull 166: 482– 493, 1984.
402. Harris HE and Epstein HF. Myosin and paramyosin of Caenorhabditis elegans: biochemical and structural properties of wildtype and mutant proteins. Cell 10: 709 –719, 1977.
403. Hartmann S, Adam R, Marti T, Kirsten C, Seidinger S, and
Lucius R. A 41-kDa antigen of the rodent filaria Acanthocheilonema viteae with homologies to tropomyosin induces hostprotective immune responses. Parasitol Res 83: 390 –393, 1997.
404. Hasegawa Y. Complete nucleotide sequences of cDNA encoding
for tropomyosin isoforms from the catch muscle of scallop Patinopecten yessoensis. Fish Sci 67: 988 –990, 2001.
405. Hasegawa Y. Complete nucleotide sequence of a cDNA encoding
a myosin heavy chain from mantle tissue of scallop Patinopecten
yessoensis. Fish Sci 68: 403– 415, 2002.
406. Hastings GA and Emerson CP Jr. Myosin functional domains
encoded by alternative exons are expressed in specific thoracic
muscles of Drosophila. J Cell Biol 114: 263–276, 1991.
407. Hastings KEM. Molecular evolution of the vertebrate troponin I
gene family. Cell Tissue Res 22: 205–211, 1997.
408. Havercroft JC, Huggins MC, Dunne DW, and Taylor DW.
Characterization of Sm20, a 20 kilodalton calcium-binding protein
of Schistosoma mansoni. Mol Biochem Parasitol 38: 211–219,
1990.
409. Havercroft JC, Smith AL, and Williams RH. Schistosoma
mansoni: immuno-localization of the calcium binding protein
Sm20. Parasite Immunol 13: 593– 604, 1991.
410. He M and Haymer DS. Codon bias in actin multigene families
and effects on the reconstruction of phylogenetic relationships. J
Mol Evol 41: 141–149, 1995.
411. He M and Haymer DS. The actin gene family in the oriental fruit
fly Bactrocera dorsalis. Muscle specific actins. Insect Biochem
Mol Biol 24: 891–906, 1994.
412. Head JF and Perry SV. The interaction of the calcium-binding
protein (troponin C) with bivalent cations and the inhibitory
protein (troponin I). Biochem J 137: 145–154, 1974.
413. Heierhorst J, Kobe B, Feil SC, Parker MW, Benian GM,
Weiss KR, and Kemp BE. Ca2⫹/S100 regulation of giant protein
kinases. Nature 380: 636 – 639, 1996.
414. Heierhorst J, Mann RJ, and Kemp BE. Interaction of the
recombinant S100A1 protein with twitchin kinase, and comparison with other Ca2⫹-binding proteins. Eur J Biochem 249: 127–
133, 1997.
415. Heierhorst J, Probst WC, Kohanski RA, Buku A, and Weiss
KR. Phosphorylation of myosin regulatory light chains by the
molluscan twitchin kinase. Eur J Biochem 233: 426 – 431, 1995.
416. Heierhorst J, Probst WC, Vilim FS, Buku A, and Weiss KR.
Autophosphorylation of molluscan twitchin and interaction of its
kinase domain with calcium/calmodulin. J Biol Chem 269: 21086 –
21093, 1994.
417. Heierhorst J, Tang XX, Lei JY, Probst WC, Weiss KR, Kemp
BE, and Benian GM. Substrate specificity and inhibitor sensitivity of Ca2⫹/S100 dependent twitchin kinases. Eur J Biochem 242:
454 – 459, 1996.
418. Hennessey ES, Drummond DR, and Sparrow JC. Molecular
genetics of actin function. Biochem J 282: 657– 671, 1993.
419. Hennessey ES, Drummond DR, and Sparrow JC. Post-translational processing of the amino terminus affects actin function.
Eur J Biochem 197: 345–352, 1991.
420. Hergenhahn HG, Kegel G, and Sedlmeier D. Ca2⫹-binding
proteins in crayfish abdominal muscle. Evidence for a calmodulin
lacking trimethyllysine. Biochim Biophys Acta 787: 196 –203,
1984.
421. Hermann A and Cox JA. Sarcoplasmic calcium-binding protein.
Comp Biochem Physiol B Biochem 111: 337–345, 1995.
422. Hernandez MG, Hafalla JC, and Acosta LP. Paramyosin is a
major target of the human IgA response against Schistosoma
japonicum. Parasite Immunol 21: 641– 647, 1999.
423. Herranz R, Dı́iz-Castillo C, Nguyen TP, Lovato TL, Cripps
RM, and Marco R. Expression patterns of the whole troponin C
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
424.
425.
426.
427.
428.
429.
430.
431.
432.
433.
434.
435.
436.
437.
438.
439.
440.
441.
442.
443.
444.
gene repertoire during Drosophila development. Gene Exp Patterns 4: 183–190, 2004.
Hess NK and Bernstein SI. Developmentally regulated alternative splicing of Drosophila myosin heavy chain transcripts: in vivo
analysis of an unusual 3⬘ splice site. Dev Biol 146: 339 –344, 1991.
Hidalgo C, Craig R, Ikebe M, and Padron R. Mechanism of
phosphorylation of the regulatory light chain of myosin from
tarantula striated muscle. J Muscle Res Cell Motil 22: 51–59, 2001.
Hidalgo C, Padron R, Horowitz R, Zhao FQ, and Craig R.
Purification of native myosin filaments from muscle. Biophys J 81:
2817–2826, 2001.
Higgins DG, Labeit S, Gautel M, and Gibson TJ. The evolution
of titin and related giant muscle proteins. J Mol Evol 38: 395– 404,
1994.
Hikosaka A, Kusakabe T, and Satoh N. Short upstream sequences associated with the muscle specific expression of an
actin gene in ascidian embryos. Dev Biol 166: 763–769, 1994.
Hikosaka A, Satoh N, and Makabe KW. Regulated spatial expression of fusion gene constructs with the 5⬘ upstream region of
Halocynthia roretzi muscle actin gene in Ciona savignyi embryos. Roux’s Arch Dev Biol 203: 104 –112, 1993.
Hiromi Y and Hotta Y. Actin gene mutations in Drosophila-heat
shock activation in the indirect flight muscles. EMBO J 4: 1681–
1687, 1985.
Hiromi Y, Okamoto H, Gehring WJ, and Hotta Y. Germline
transformation with Drosophila mutant actin genes induces constitutive expression of heat shock genes. Cell 44: 293–301, 1986.
Hobert O, Moerman DG, Clark KA, Beckerle MC, and Ruvkun G. A conserved LIM protein that affects muscular adherens
junction integrity and mechanosensory function in Caenorhabditis elegans. J Cell Biol 144: 45–57, 1999.
Hodge T and Cope MJTV. A myosin family tree. J Cell Sci 113:
3353–3354, 2000.
Hodges D and Bernstein SI. Suboptimal 5⬘ splice and 3⬘ splice
sites regulate alternative splicing of Drosophila melanogaster
myosin heavy chain transcripts in vitro. Mech Dev 37: 127–140,
1992.
Hodges D, Cripps RM, O’Connor ME, and Bernstein SI. The
role of evolutionarily conserved sequences in alternative splicing
at the 3⬘ end of Drosophila melanogaster myosin heavy chain
RNA. Genetics 151: 263–276, 1999.
Hoekstra R, Visser A, Otsen M, Tibben J, Lenstra JA, and
Roos MH. EST sequencing of the parasitic nematode Haemonchus contortus suggests a shift in gene expression during transition to the parasitic stages. Mol Biochem Parasitol 110: 53– 68,
2000.
Holland LZ. Muscle development in amphioxus: morphology,
biochemistry, and molecular biology. Isr J Zool 42: S235–S246,
1996.
Holland LZ, Pace DA, Blink ML, Kene M, and Holland ND.
Sequence and expression of amphioxus alkali myosin light-chain
(Amphimlc-Alk) throughout development—implications for vertebrate myogenesis. Dev Biol 171: 665– 676, 1995.
Holmes JM, Whiteley NM, Magnay JL, and El Haj AJ. Comparison of the variable loop regions of myosin heavy chain genes
from Antarctic and temperate isopods. Comp Biochem Physiol B
Biochem 131: 349 –359, 2002.
Homyk T Jr. Behavioral mutants of Drosophila melanogaster. II.
Behavioral analysis and focus mapping. Genetics 87: 105–128,
1977.
Homyk T Jr and Emerson CP Jr. Functional interactions between unlinked muscle genes within haploinsufficient regions of
the Drosophila genome. Genetics 119: 105–121, 1988.
Homyk T Jr and Sheppard DE. Behavioral mutants of Drosophila melanogaster. I. Isolation and mapping of mutations which
decrease flight ability. Genetics 87: 95–104, 1977.
Homyk T Jr, Szydonya J, and Suzuki DT. Behavioral mutants
of Drosophila melanogaster. III. Isolation and mapping of mutations by direct visual observations of behavioral phenotypes. Mol
Gen Genet 177: 553–565, 1980.
Hooker CW, Yang W, and Becker MM. Schistosoma japonicum: heterogeneity in paramyosin genes. Acta Tropica 59: 131–
141, 1995.
Physiol Rev • VOL
1039
445. Hoppe PE, Andrews RC, and Parikh PD. Differential requirement for the nonhelical tailpiece and the C terminus of the myosin
rod in Caenorhabditis elegans muscle. Mol Biol Cell 14: 1677–
1690, 2003.
446. Hoppe PE and Waterston RH. A region of the myosin rod
important for interaction with paramyosin in Caenorhabditis elegans striated muscle. Genetics 156: 631– 643, 2000.
447. Hoppe PE and Waterston RH. Hydrophobicity variations along
the surface of the coiled-coil rod may mediate striated muscle
myosin assembly in Caenorhabditis elegans. J Cell Biol 135:
371–382, 1996.
448. Horie N, Tsuchiya T, and Matsumoto JJ. Studies on ATPase
activity of actomysosin of squid mantle muscles. Nippon Suisan
Gakk 41: 1039 –1045, 1975.
449. Horovitch SJ, Storti RV, Rich A, and Pardue ML. Multiple
actins in Drosophila melanogaster. J Cell Biol 82: 86 –92, 1979.
450. Hosono R and Kuno S. Properties of the unc-52 gene and its
related mutations in the nematode Caenorhabditis elegans. Zool
Sci 2: 81– 88, 1985.
451. Hotta Y and Benzer S. Mapping behavior in Drosophila mosaics.
Nature 240: 527–536, 1972.
452. Houdusse A and Cohen C. Target sequence recognition by the
calmodulin superfamily: implications from light chain binding to
the regulatory domain of scallop myosin. Proc Natl Acad Sci USA
92: 10644 –10647, 1995.
453. Houdusse A, Szent-Györgyi AG, and Cohen C. Three conformational states of scallop myosin S1. Proc Natl Acad Sci USA 97:
11238 –11243, 2000.
454. Hresko MC, Schriefer LA, Shrimankar P, and Waterston RH.
Myotactin, a novel hypodermal protein involved in muscle cell
adhesion in Caenorhabditis elegans. J Cell Biol 146: 659 – 672,
1999.
455. Hu DH, Kimura S, and Maruyama K. Sodium dodecyl sulfate gel
electrophoresis studies of connectin-like high molecular weight
proteins of various types of vertebrate and invertebrate muscles.
J Biochem 99: 1485–1492, 1986.
456. Hu DH, Matsuno A, Terakado K, Matsuura T, Kimura S, and
Maruyama K. Projectin is an invertebrate connectin (titin): isolation from crayfish claw muscle and localization in crayfish claw
muscle and insect flight muscle. J Muscle Res Cell Motil 11:
497–511, 1990.
457. Hu SH, Lei JY, Wilce MCJ, Valenzuela MRL, Benian GM,
Parker MW, and Kemp BE. Crystallization and preliminary X-ray
analysis of the auto-inhibited twitchin kinase. J Mol Biol 236:
1259 –1261, 1994.
458. Hu SH, Parker MW, Lei JY, Wilce MCJ, Benian GM, and
Kemp BE. Insights into autoregulation from the crystal structure
of twitchin kinase. Nature 369: 581–584, 1994.
459. Huch R and D’Haese J. Quantification of the soluble calciumbinding protein (SCBP) in various muscle tissues of the terrestrial
oligochaete Lumbricus terrestris L. Soil Biol Biochem 24: 1231–
1235, 1992.
460. Huch R, D’Haese J, and Gerday C. A soluble calcium-binding
protein from the terrestrial annelid Lumbricus terrestris L.
J Comp Physiol B Biochem Syst Environ Physiol 158: 325–334,
1988.
461. Hunter SJ, Martin SAM, Thompson FJ, Tetley L, and Devaney E. The isolation of differentially expressed cDNA clones
from the filarial nematode Brugia pahangi. Parasitology 119:
189 –198, 1999.
462. Hurtado JL, Montero P, Borderias J, and Solas MT. Morphological and physical changes during heating of pressurized common octopus muscle up to cooking temperature. Food Sci Technol
Int 7: 329 –338, 2001.
463. Hurtado JL, Montero P, Borderı́as J, and Solas MT. Highpressure/temperature treatment effect on the characteristics of
octopus (Octopus vulgaris) arm muscle. Eur Food Res Technol
213: 22–29, 2001.
464. Hyatt CJ and Maughan DW. Fourier analysis of wing beat
signals: assessing the effects of genetic alterations of flight muscle
structure in Diptera. Biophys J 67: 1149 –1154, 1994.
465. Hynes RO and Zhao Q. The evolution of cell adhesion. J Cell Biol
150: F89 –F95, 2000.
85 • JULY 2005 •
www.prv.org
1040
SCOTT L. HOOPER AND JEFFREY B. THUMA
466. Ibrahim MS, Eisinger SW, and Scott AL. Muscle actin gene
from Aedes aegypti (Diptera: Culicidae). J Med Entomol 33: 955–
962, 1996.
467. Iguchi SMM, Tsuchiya T, and Matsumoto JJ. Studies on the
freeze denaturation of squid actomyosin. Nippon Suisan Gakk 47:
1499 –1506, 1981.
468. Ikemoto N and Kawaguti S. Elongating effect of tropomyosin A
on the thick myofilaments in the long-sarcomere muscle of the
horse-shoe crab. Proc Jpn Acad 43: 974 –979, 1967.
469. Inoue A, Ojima T, and Nishita K. Cloning and sequencing of
cDNA for Akazara scallop tropomyosin. Fish Sci 65: 772–776,
1999.
470. Inoue A, Ojima T, and Nishita K. Cloning and sequencing of a
cDNA for Akazara scallop troponin T. J Biochem 120: 834 – 837,
1996.
471. Inoue A, Ojima T, and Nishita K. Cloning and sequence of a
cDNA for troponin T of Ezo giant scallop striated muscle. Fish Sci
64: 459 – 463, 1998.
472. Inoue A, Ojima T, and Nishita K. Partial nucleotide sequence of
a cDNA encoding C-terminal polymerizable region of akazara
scallop tropomyosin. Fish Sci 64: 164 –165, 1998.
473. Inoue A, Ojima T, and Nishita K. N-terminal modification and
its effect on the biochemical characteristics of akazara scallop
tropomyosins expressed in Escherichia coli. J Biochem 136: 107–
114, 2004.
474. Inoue K, Sohma H, and Morita F. Ca2⫹-dependent protein
phosphatase which dephosphorylates regulatory light chain a in
scallop smooth muscle myosin. J Biochem 107: 872– 878, 1990.
475. Irvine M, Huima T, Prince AM, and Lustigman S. Identification and characterization of an Onchocerca volvulus cDNA clone
encoding a highly immunogenic calponin-like protein. Mol Biochem Parasitol 65: 135–136, 1994.
476. Irving T, Bhattacharya S, Tesic I, Moore J, Farman G, Simcox A, Vigoreaux J, and Maughan D. Changes in myofibrillar
structure and function produced by N-terminal deletion of the
regulatory light chain in Drosophila. J Muscle Res Cell Motil 22:
675– 683, 2001.
477. Ishida K and Konno K. Characteristic properties of native tropomyosin and of troponin from squid mantle. Nippon Suisan
Gakk 48: 843– 849, 1982.
478. Ishii AI and Sano M. Isolation and identification of paramyosin
from liver fluke muscle layer. Comp Biochem Physiol B Biochem
65: 537–541, 1980.
479. Ishikawa M, Ishida M, Shimakura K, Nagashima Y, and
Shiomi K. Tropomyosin, the major oyster Crassostrea gigas allergen and its IgE-binding epitopes. J Food Sci 63: 44 – 47, 1998.
480. Ishikawa M, Ishida M, Shimakura K, Nagashima Y, and
Shiomi K. Purification and IgE-binding epitopes of a major allergen in the gastropod Turbo cornutus. Biosci Biotechnol Biochem
62: 1337–1343, 1998.
481. Ishikawa M, Nagashima Y, and Shiomi K. Immunological comparison of shellfish allergens by competitive enzyme-linked immunosorbent assay. Fish Sci 65: 592–595, 1999.
482. Ishikawa M, Shimakura K, Nagashima Y, and Shiomi K. Isolation and properties of allergenic proteins in the oyster Crassostrea gigas. Fish Sci 63: 610 – 614, 1997.
483. Ishikawa M, Suzuki F, Ishida M, Nagashima Y, and Shiomi K.
Identification of tropomyosin as a major allergen in the octopus
Octopus vulgaris and elucidation of its IgE binding epitopes. Fish
Sci 67: 934 –942, 2001.
484. Ishikawa M, Nagashima Y, and Shiomi K. Identification of the
oyster allergen Cra g 2 as tropomyosin. Fish Sci 64: 854 – 855,
1998.
485. Ishimoda-Takagi T. Immunological purification of sea urchin
egg tropomyosin. J Biochem 83: 1757–1762, 1978.
486. Ishimoda-Takagi T, Chino I, and Sato H. Evidence for the
involvement of muscle tropomyosin in the contractile elements of
the coelom esophagus complex in sea urchin embryos. Dev Biol
105: 365–376, 1984.
487. Ishimoda-Takagi T, Itoh M, and Koyama H. Distribution of
tropomyosin isoforms in spiny lobster muscles. J Exp Zool 277:
87–98, 1997.
Physiol Rev • VOL
488. Ishimoda-Takagi T and Kobayashi M. Molecular heterogeneity
and tissue specificity of tropomyosin obtained from various bivalves. Comp Biochem Physiol B Biochem 88: 443– 452, 1987.
489. Ishimoda-Takagi T, Kobayashi M, and Yaguchi M. Polymorphism and tissue-specificity of scallop tropomyosin. Comp Biochem Physiol B Biochem 83: 515–521, 1986.
490. Ishimoda-Takagi T, Motohashi A, and Ishikita S. Characterization of two distinct isoforms of tropomyosins present in the
eggs of the sea urchin, Strongylocentrotus intermedius. Comp
Biochem Physiol B Biochem 95: 403– 413, 1990.
491. Ishimoda-Takagi T and Ozaki S. Effect of trichloroacetic acid
on the isolation of tropomyosin from sea urchin lantern muscle.
J Biochem 93: 801– 805, 1983.
492. Ishimoda-Takagi T, Yamazaki K, Ando T, Minami H, and
Nagata A. Sea urchin egg tropomyosin isoforms with muscle-type
and nonmuscle-type antigenicities. Comp Biochem Physiol B Biochem 112: 415– 427, 1995.
493. Iwamoto M, Yamanaka H, Watabe S, and Hashimoto K.
Changes in ATP and related breakdown compounds in the adductor muscle of Itayagai scallop Pecten albicans during storage at
various temperatures. Nippon Suisan Gakk 57: 153–156, 1991.
494. Iwasaki K, Kikuchi K, Funabara D, and Watabe S. cDNA
cloning of tropomyosin from the anterior byssus retractor muscle
of mussel and its structural integrity from the deduced amino acid
sequence. Fish Sci 63: 731–734, 1997.
495. James SL. Shistosoma spp: progress toward a defining vaccine.
Exp Parasitol 63: 247–252, 1987.
496. Janes DP, Patel H, and Chantler PD. Primary structure of
myosin from the striated adductor muscle of the Atlantic scallop,
Pecten maximus, and expression of the regulatory domain.
J Muscle Res Cell Motil 21: 415– 422, 2000.
497. Jantsch-Plunger V and Fire A. Combinatorial structure of a
body muscle specific transcriptional enhancer in Caenorhabditis
elegans. J Biol Chem 269: 27021–27028, 1994.
498. Jauregui-Adell J, Wnuk W, and Cox JA. Complete amino acid
sequence of the sarcoplasmic calcium-binding protein (SCP-I)
from crayfish (Astacus leptodactylus). FEBS Lett 243: 209 –212,
1989.
499. Jeffery WR. Actin as a tissue specific marker in studies of ascidian development and evolution. Adv Dev Biol 3: 137–183, 1994.
500. Jeffery WR, Swalla BJ, Ewing N, and Kusakabe T. Evolution
of the ascidian anural larva: evidence from embryos and molecules. Mol Biol Evol 16: 646 – 654, 1999.
501. Jenkins RE, Taylor MJ, Gilvary NJ, and Bianco AE. Tropomyosin implicated in host protective responses to microfilariae in
onchocerciasis. Proc Natl Acad Sci USA 95: 7550 –7555, 1998.
502. Jeong KY, Hwang H, Lee J, Lee IY, Kim DS, Hong CS, Re HI,
and Yong TS. Allergenic characterization of tropomyosin from
the dusky brown cockroach, Periplaneta fuliginosa. Clin Diagn
Lab Immunol 11: 680 – 685, 2004.
503. Jeong KY, Lee J, Lee IY, Hong CS, Ree HI, and Yong TS.
Expression of tropomyosin from Blattella germanica as a recombinant non-fusion protein in Pichia pastoris and comparison of
its IgE reactivity with its native counterpart. Protein Exp Purif 37:
273–278, 2004.
504. Jeong KY, Lee J, Lee IY, Ree HI, Hong CS, and Yong TS.
Allergenicity of recombinant Bla g 7, German cockroach tropomyosin. Allergy 58: 1059 –1063, 2003.
505. Jeong KY, Yum HY, Lee IY, Ree HI, Hong CS, Kim DS, and
Yong TS. Molecular cloning and characterization of tropomyosin,
a major allergen of Chironomus kiiensis, a dominant species of
nonbiting midges in Korea. Clin Diagn Lab Immunol 11: 320 –324,
2004.
506. Jeoung BJ, Reese G, Hauck P, Oliver JB, Daul CB, and
Lehrer SB. Quantification of the major brown shrimp allergen
Pen a 1 (tropomyosin) by a monoclonal antibody-based sandwich
ELISA. J Allergy Clin Immunol 100: 229 –234, 1997.
507. Johnson KA and Quiocho FA. Protein kinases: twitching worms
catch S100. Nature 380: 585–587, 1996.
508. Johnson PJ, Foran DR, and Moore GP. Organization and evolution of the actin gene family in sea urchins. Mol Cell Biol 3:
1824 –1833, 1983.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
509. Jones MK, Gobert GN, Zhang LH, Sunderland P, and McManus DP. The cytoskeleton and motor proteins of human schistosomes and their roles in surface maintenance and host-parasite
interactions. Bioessays 26: 752–765, 2004.
510. Jones MK, Yang W, and McManus DP. Immunolocalization of
the 38.3 kDa calponin-like protein in stratified muscles of the tail
of Schistosoma japonicum cercariae. Parasitol Int 50: 129 –133,
2001.
511. Joseph H. Einige anatomische und histologische Notizen über
Amphioxus. Arb Zool Inst Univ Wien 13: 125–154, 1902.
512. Just I, Hennessey ES, Drummond DR, Aktories K, and Sparrow JC. ADP-ribosylation of Drosophila indirect flight muscle
actin and arthrin by Clostridium botulinum C2 toxin and Clostridium perfringens Iota toxin. Biochem J 291: 409 – 412, 1993.
513. Kabat-Zinn J and Singer RH. Sea urchin tube feet: unique
structures that allow a cytological and molecular approach to the
study of actin and its gene expression. J Cell Biol 89: 109 –114,
1981.
514. Kagawa H, Gengyo K, McLachlan AD, Brenner S, and Karn J.
Paramyosin gene (unc15) of Caenorhabditis elegans. Molecular
cloning, nucleotide sequence, and models for thick filament structure. J Mol Biol 207: 311–333, 1989.
515. Kagawa H, Sugimoto K, Matsumoto H, Inoue T, Imadzu H,
Takuwa K, and Sakube Y. Genome structure, mapping, and
expression of the tropomyosin gene tmy-1 of Caenorhabditis
elegans. J Mol Biol 251: 603– 613, 1995.
516. Kagawa H, Takuwa K, and Sakube Y. Mutations and expressions of the tropomyosin gene and the troponin C gene of Caenorhabditis elegans. Cell Struct Funct 22: 213–218, 1997.
517. Kagawa H and Bando T. Tissue expression of the tropomyosin
gene of Caenorhabditis elegans-enhancer modification of the promoter control. Seikagaku 74: 1176 –1180, 2002.
518. Kagawa M, Matsumoto M, and Hatae K. Taste differences
among three kinds of squid and the effect of cold storage on the
taste. J Home Econ Jpn 50: 1245–1254, 1999.
519. Kagawa M, Matsumoto M, and Hatae K. Differences in texture
among three varieties of squid and the effect of cold storage on
the texture. J Home Econ Jpn 51: 699 –708, 2000.
520. Kagawa M, Matsumoto M, Yoneda C, Mitsuhashi T, and Hatae K. Changes in meat texture of three varieties of squid in the
early stage of cold storage. Fisheries Sci 68: 783–792, 2002.
521. Kalinna BH, Becker MM, and McManus DP. Engineering and
expression of a full length cDNA encoding Schistosoma japonicum paramyosin: purification of the recombinant protein and its
recognition by infected patient sera. Acta Trop 65: 111–115, 1997.
522. Kalinna BH and McManus DP. IgG (Fc␥) binding protein of
Taenia crassiceps (cestoda) exhibits sequence homology and
antigenic similarity with schistosome paramyosin. Parasitology
106: 289 –296, 1993.
523. Kalinna GH and McManus DP. A vaccine against the Asian
schistosome, Schistosoma japonicum: an update on paramysosin
as a target of protective immunity. Int J Parasitol 27: 11213–
11219, 1997.
524. Kamidochi M, Higashihara M, Yazawa Y, and Tanaka I. Scallop striated tropomyosin prepared by a new method. Proc Jpn
Acad 75: 307–310, 1999.
525. Kamiya S and Konno K. Calcium sensitivity of H-meromyosin
and subfragment-1 from squid mantle muscle. Nippon Suisan
Gakk 50: 1889 –1896, 1984.
526. Kang WK and Naya Y. Sequence of the cDNA encoding an actin
homolog in the crayfish Procumbarus clarkii. Gene 133: 303–304,
1993.
527. Kantha SS. Paramyosin as helminth vaccine candidate. Lancet
336: 320 –320, 1990.
528. Kantha SS, Watabe S, and Hashimoto K. Comparative biochemistry of paramyosin—a review. J Food Biochem 14: 61– 88,
1990.
529. Kantha SS, Watabe S, and Hashimoto K. Immunological comparison between the paramyosins from adductor muscle of bivalves and from foot muscle of a gastropod. Comp Biochem
Physiol B Biochem 96: 89 –92, 1990.
530. Karabinos A, Riemer D, Erber A, and Weber K. Homologues
of vertebrate type I, II, and III intermediate filament (IF) proteins
Physiol Rev • VOL
531.
532.
533.
534.
535.
536.
537.
538.
539.
540.
541.
542.
543.
544.
545.
546.
547.
548.
549.
550.
1041
in an invertebrate: the IF multigene family of the cephalochordate
Branchiostoma. FEBS Lett 437: 15–18, 1998.
Karabinos A, Schmidt H, Harborth J, Schnabel R, and Weber
K. Essential roles for four cytoplasmic intermediate filament proteins in Caenorhabditis elegans development. Proc Natl Acad Sci
USA 98: 7863–7868, 2001.
Karabinos A, Schulze E, Klisch T, Wang J, and Weber K.
Expression profiles of the essential intermediate filament (IF)
protein A2 and the IF protein C2 in the nematode Caenorhabditis
elegans. Mech Dev 117: 311–314, 2002.
Karabinos A, Schulze E, Schünemann J, Parry DAD, and
Weber K. In vivo and in vitro evidence that the four essential
intermediate filament (IF) proteins A1, A2, A3, and B1 of the
nematode Caenorhabditis elegans form an obligate heteropolymeric IF system. J Mol Biol 333: 307–319, 2003.
Karabinos A, Wang J, Wenzel D, Panopoulou G, Lehrach H,
and Weber K. Developmentally controlled expression patterns of
intermediate filament proteins in the chaphalochordate Branchiostoma. Mech Dev 101: 283–288, 2001.
Karabinos A, Zimek A, and Weber K. The genome of the early
chordate Ciona intestinalis encodes only five cytoplasmic intermediate filament proteins including a single type I and type II
keratin and a unique IF-annexin fusion protein. Gene 326: 123–129,
2004.
Karlik CC, Coutu MD, and Fyrberg EA. A nonsense mutation
within the act88F actin gene disrupts myofibril formation in Drosophila indirect flight muscles. Cell 38: 711–719, 1984.
Karlik CC and Fyrberg EA. Two Drosophila melanogaster tropomyosin genes: structural and functional aspects. Mol Cell Biol 6:
1965–1973, 1986.
Karlik CC and Fyrberg EA. An insertion within a variably
spliced Drosophila tropomyosin gene blocks accumulation of
only one encoded isoform. Cell 41: 57– 66, 1985.
Karlik CC, Mahaffey JW, Coutu MD, and Fyrberg EA. Organization of contractile protein genes within the 88F subdivision of
the D. melanogaster third chromosome. Cell 37: 469 – 481, 1984.
Karlik CC, Saville DL, and Fyrberg EA. Two missense alleles
of the Drosophila melanogaster act88F actin gene are strongly
antimorphic but only weakly induce synthesis of heat shock proteins. Mol Cell Biol 7: 3084 –3091, 1987.
Karn J, Brenner S, and Barnett L. Protein structural domains
in the Caenorhabditis elegans unc-54 myosin heavy chain gene
are not separated by introns. Proc Natl Acad Sci USA 80: 4253–
4257, 1983.
Kasamatsu C, Kimura S, Kagawa M, and Hatae K. Identification of high molecular weight proteins in squid muscle by Western
blotting analysis and postmortem rheological changes. Biosci
Biotechnol Biochem 68: 1119 –1124, 2004.
Katayama S, Shima J, and Saeki H. Solubility improvement of
shellfish muscle proteins by reaction with glucose and its soluble
state in low-ionic-strength medium. J Agric Food Chem 50: 4327–
4332, 2002.
Kawamura Y, Ohtsuka H, Murata H, Maki S, Ohtani Y,
Manabe T, Kimura S, and Maruyama K. Comparative aspects
of muscle elastic proteins. Adv Biophys 33: 175–181, 1996.
Kawamura Y, Suzuki J, Kimura S, and Maruyama K. Characterization of connectin-like proteins of obliquely striated muscle
of a polychaete (Annelida). J Muscle Res Cell Motil 15: 623– 632,
1994.
Kawashima K and Yamanaka H. Effects of storage temperatures on the post-mortem biochemical changes in scallop adductor muscle. Nippon Suisan Gakk 58: 2175–2180, 1992.
Kay CM. Some physico-chemical properties of Pinna nobilis
tropomyosin. Biochim Biophys Acta 27: 469 – 477, 1958.
Kay CM. The partial specific volume of muscle proteins. Biochim
Biophys Acta 38: 420 – 427, 1960.
Kay CM and Bailey K. Further physico-chemical studies on
Pinna nobilis tropomyosin. Biochim Biophys Acta 31: 20 –25,
1959.
Kazzaz JA and Rozek CE. Tissue specific expression of the
alternately processed Drosophila myosin heavy chain messenger
RNAs. Dev Biol 133: 550 –561, 1989.
85 • JULY 2005 •
www.prv.org
1042
SCOTT L. HOOPER AND JEFFREY B. THUMA
551. Kelly LE, Phillips AM, Delbridge M, and Stewart R. Identification of a gene family from Drosophila melanogaster encoding
proteins with homology to invertebrate sarcoplasmic calciumbinding proteins (SCPS). Insect Biochem Mol Biol 27: 783–792,
1997.
552. Kendrick-Jones J, Lehman W, and Szent-Györgyi AG. Regulation in molluscan muscles. J Mol Biol 54: 313–326, 1970.
553. Kendrick-Jones J, Szentkiralyi EM, and Szent-Györgyi AG.
Regulatory light chains in myosins. J Mol Biol 104: 747–775, 1976.
554. Kenny PA, Liston EM, and Higgins DG. Molecular evolution of
immunoglobulin and fibronectin domains in titin and related muscle proteins. Gene 232: 11–23, 1999.
555. Kenyon F, Welsh M, Parkinson J, Whitton C, Blaxter ML,
and Knox DP. Expressed sequence tag survey of gene expression
in the scab mite Psoroptes ovis-allergens, proteases and freeradical scavengers. Parasitology 126: 451– 460, 2003.
556. Khaitlina SY. Polymerization of beta-like actin from scallop adductor muscle. FEBS Lett 198: 221–224, 1986.
557. Khaitlina SY. Functional specificity of actin isoforms. Int Rev
Cytol 202: 35–98, 2001.
558. Khaitlina SY, Antropova O, Kuznetsova I, Turoverov K, and
Collins JH. Correlation between polymerizability and conformation in scallop beta-like actin and rabbit skeletal muscle alphaactin. Arch Biochem Biophys 368: 105–111, 1999.
559. Khaitlina SY, Tskhovrebova LA, and Sheludko NS. Unusual
extraction of alpha-actinin from the adductor muscles of bivalve
mollusks. Comp Biochem Physiol B Biochem 73: 655– 661, 1982.
560. Kiehl E and D’Haese J. A soluble calcium-binding protein
(SCBP) present in Drosophila melanogaster and Calliphora
erythrocephala muscle cells. Comp Biochem Physiol B Biochem
102: 475– 482, 1992.
561. Kier WM and Schachat FH. Biochemical comparison of fastand slow-contracting squid muscle. J Exp Biol 168: 41–56, 1992.
562. Kiff JE, Moerman DG, Schreifer LA, and Waterston RH.
Transposon-induced deletions in unc-22 of C. elegans associated
with almost normal gene activity. Nature 331: 631– 633, 1988.
563. Kimura I, Yoshitomi B, Konno K, and Arai KI. Preparation of
highly purified myosin from mantle muscle of squid, Ommastrephes sloani pacificus. Nippon Suisan Gakk 46: 885– 892, 1980.
564. Kimura K, Tanaka T, Nakae H, and Obinata T. Troponin from
nematode: purification and characterization of troponin from ascaris body wall muscle. Comp Biochem Physiol B Biochem 88:
399 – 407, 1987.
565. Kimura M, Narita M, Imamura T, Ushio H, and Yamanaka H.
High quality control of scallop adductor muscle by different modified atmosphere packaging. Nippon Suisan Gakk 66: 475– 480,
2000.
566. Kimura S, Miyaki T, Takema Y, and Kubota M. Electrophoretic analysis of connectin from the muscles of aquatic animals. Nippon Suisan Gakk 47: 787–791, 1981.
567. Kishimura H, Ojima T, and Nishita K. Hybridization experiments using fish myosin light chains and desensitized akazara
myosin. Nippon Suisan Gakk 52: 847– 851, 1986.
568. Kissinger JC, Hahn JH, and Raff RA. Rapid evolution in a
conserved gene family. Evolution of the actin gene family in the
sea urchin genus Heliocidaris and related genera. Mol Biol Evol
14: 654 – 665, 1997.
569. Kissinger JC and Raff RA. Evolutionary changes in sites and
timing of actin gene expression in embryos of the direct- and
indirect-developing sea urchins, Heliocidaris erythrogramma
and H. tuberculata. Dev Genes Evol 208: 82–93, 1998.
570. Ko J, Horiuchi S, and Yamaguchi M. Myosin from abdominal
flexor muscle in a crayfish, Procambarus clarki Girard. J Biochem
85: 541–548, 1979.
571. Koana T and Hotta Y. Isolation and characterization of flightless
mutants in Drosophila melanogaster. J Embryol Exp Morphol 45:
123–143, 1978.
572. Kobayashi C, Kobayashi S, Orii H, Watanabe K, and Agata K.
Identification of two distinct muscles in the planarian Dugesia
japonica by their expression of myosin heavy chain genes. Zool
Sci 15: 861– 869, 1998.
Physiol Rev • VOL
573. Kobayashi T, Kagami O, Takagi T, and Konishi K. Amino acid
sequence of horseshoe crab, Tachypleus tridentatus, striated
muscle troponin C. J Biochem 105: 823– 828, 1989.
574. Kobayashi T, Takagi T, Konishi K, and Cox JA. The primary
structure of a new Mr 18000 calcium vector protein from amphioxus. J Biol Chem 262: 2613–2623, 1987.
575. Kobayashi T, Takagi T, Konishi K, and Cox JA. Amino acid
sequence of crayfish troponin I. J Biol Chem 264: 1551–1557, 1989.
576. Kobayashi T, Takagi T, Konishi K, and Wnuk W. Amino acid
sequences of the two major isoforms of troponin C from crayfish.
J Biol Chem 264: 18247–18259, 1989.
577. Kobayashi T, Takasaki Y, Takagi T, and Konishi K. The amino
acid sequence of sarcoplasmic calcium-binding protein obtained
from sandworm, Perinereis vancaurica tetradentata. Eur J Biochem 144: 401– 408, 1984.
578. Kobe B, Heierhorst J, Feil SC, Parker MW, Benian GM,
Weiss KR, and Kemp BE. Giant protein kinases: domain interactions and structural basis of autoregulation. EMBO J 15: 6810 –
6821, 1996.
579. Kobe B, Heierhorst J, and Kemp BE. Intrasteric regulation of
protein kinases. Adv Second Messenger Phosphoprotein Res 31:
29 – 40, 1997.
580. Kodama S and Konno K. Isolation and biochemical properties of
myosin from squid brachial muscle. Nippon Suisan Gakk 49:
437– 442, 1983.
581. Koenders A, Lamey TM, Medler S, West JM, and Mykles DL.
Two fast-type fibers in claw closer and abdominal deep muscles of
the Australian freshwater crustacean, Cherax destructor, differ in
Ca2⫹ sensitivity and troponin-I isoforms. J Exp Zool 301: 588 –598,
2004.
582. Kohler L, Cox JA, and Stein EA. Sarcoplasmic calcium-binding
proteins in protochordate and cyclostome muscle. Mol Cell Biochem 20: 85–93, 1979.
583. Kohlmetz C, Epe C, and Schnieder T. In vitro expression of a
recombinant paramyosin of Ancylostoma caninum. Int J Parasitol 28: 1229 –1233, 1998.
584. Koike Y, Mita K, Suzuki MG, Maeda S, Abe H, Osoegawa K,
Dejong PJ, and Shimada T. Genomic sequence of a 320-kb
segment of the Z chromosome of Bombyx mori containing a
kettin ortholog. Mol Genet Genomics 269: 137–149, 2003.
585. Kojima S. Overview: from the horse experimentation by Prof.
Akira Fujinami to paramyosin. Parasitol Int 53: 151–162, 2004.
586. Kojima SI, Mishima M, Mabuchi I, and Hotta Y. A single
Drosophila melanogaster myosin light chain kinase gene produces multiple isoforms whose activities are differently regulated.
Genes Cells 1: 855– 871, 1996.
587. Kolmerer B, Clayton J, Benes V, Allen T, Ferguson C, Leonard K, Weber U, Knekt M, Ansorge W, Labeit S, and Bullard
B. Sequence and expression of the kettin gene in Drosophila
melanogaster and Caenorhabditis elegans. J Mol Biol 296: 435–
448, 2000.
588. Kolodziejska I, Pacana J, and Sikorski ZE. Effect of squid liver
extract on proteins and on the texture of cooked squid mantle. J
Food Biochem 16: 141–150, 1992.
589. Kolodziejska I, Sikorski ZE, and Sadowska M. Texture of
cooked mantle of squid Illex argentinus as influenced by specimen characteristics and treatments. J Food Sci 52: 932–935, 1987.
590. Kölsch B, Ziegler C, and Beinbrech G. Length determination of
synthetic thick filaments by cooperation of 2 myosin associated
proteins, paramyosin and projectin. Naturwissenschaften 82:
239 –241, 1995.
591. Kominz DR, Maruyama K, Levenbrook L, and Lewis M. Tropomyosin, myosin and actin from the blowfly, Phormia regina.
Biochim Biophys Acta 63: 106 –116, 1962.
592. Kondo S, Asakawa T, and Morita F. Difference UV-absorption
spectrums of scallop adductor myosin induced by ATP. J Biochem
86: 1567–1571, 1979.
593. Kondo S and Morita F. Smooth muscle of scallop adductor
contains at least two kinds of myosin. J Biochem 90: 673– 681,
1981.
594. Konno K. Thiols in scallop (Patinopecten yessoensis) myosin and
regulatory light chain binding. Comp Biochem Physiol B Biochem
99: 161–164, 1991.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
595. Konno K. Two calcium regulation systems in squid (Ommastrephes sloani pacificus) muscle. Preparation of calcium-sensitive
myosin and troponin-tropomyosin. J Biochem 84: 1431–1440,
1978.
596. Konno K, Arai K, and Watanabe S. Myosin-linked calcium
regulation in squid mantle muscle. J Biochem 86: 1639 –1650, 1979.
597. Konno K, Kodama S, Arai K, and Watanabe S. Changes in
reactivity of scallop adductor myosin with DTNB and with TNBS
accompanying dissociation and association of regulatory light
chain. J Biochem 94: 1399 –1407, 1983.
598. Konno K and Watanabe S. Effect of regulatory light chain on
chymotryptic digestion of scallop adductor myosin. J Biochem 97:
1645–1651, 1985.
599. Konno K and Watanabe S. Two different preparations of subfragment-1 from scallop adductor myosin. J Biochem 98: 141–148,
1985.
600. Korn ED. Coevolution of head, neck, and tail domains of myosin
heavy chains. Proc Natl Acad Sci USA 97: 12559 –12564, 2000.
601. Kortschak RD, Samuel G, Saint R, and Miller DJ. EST analysis
of the cnidarian Acropora millepora reveals extensive gene loss
and rapid sequence divergence in the model invertebrates. Curr
Biol 13: 2190 –2195, 2003.
602. Kovesdi I, Preugschat R, Stuerzl M, and Smith MJ. Actin
genes from the sea star Pisaster ochraceus. Biochim Biophys
Acta 782: 76 – 86, 1984.
603. Kovesdi I and Smith MJ. Actin gene number in the sea star
Pisaster ochraceus. Can J Biochem Cell Biol 63: 1145–1151, 1985.
604. Kovesdi I and Smith MJ. Quantitative assessment of actin transcript number in eggs, embryos, and tube feed of the sea star
Pisaster ochraceus. Mol Cell Biol 5: 3001–3008, 1985.
605. Kovilur S, Jacobson JW, Beach RL, Jeffery WR, and Tomlinson CR. Evolution of the chordate muscle actin gene. J Mol Evol
36: 361–368, 1993.
606. Kowbel DJ and Smith MJ. The genomic nucleotide sequences of
two differentially expressed actin coding genes from the sea star
Pisaster ochraceus. Gene 77: 297–308, 1989.
607. Kranewitter WJ, Ylanne J, and Gimona M. unc-87 is an actinbundling protein. J Biol Chem 276: 6306 – 6312, 2001.
608. Krause M and Hirsh D. A trans-spliced leader sequence on actin
mRNA in C. elegans. Cell 49: 753–761, 1987.
609. Krause M, Wild M, Rosenzweig B, and Hirsh D. Wild-type and
mutant actin genes in Caenorhabditis elegans. J Mol Biol 208:
381–392, 1989.
610. Kretsinger RH. Calcium coordination and the calmodulin fold:
divergent versus convergent evolution. Cold Spring Harb Symp
52: 499 –510, 1987.
611. Kretsinger RH. Structure and evolution of calcium-modulated
proteins. CRC Crit Rev Biochem 8: 119 –174, 1980.
612. Kretsinger RH, Rudnick SE, Sneden DA, and Schatz VB.
Calmodulin, S-100, and crayfish sarcoplasmic calcium-binding
protein crystals suitable for X-ray diffraction studies. J Biol Chem
255: 8154 – 8156, 1980.
613. Kreuz AJ, Simcox A, and Maughan D. Alterations in flight
muscle ultrastructure and function in Drosophila tropomyosin
mutants. J Cell Biol 135: 673– 687, 1996.
614. Krieger J, Raming K, Knipper M, Grau M, Mertens S, and
Breerm H. Cloning, sequencing, and expression of locust tropomyosin. Insect Biochem 20: 173–184, 1990.
615. Kronert WA, Acebes A, Ferrus A, and Bernstein SI. Specific
myosin heavy chain mutations suppress troponin I defects in
Drosophila muscles. J Cell Biol 144: 989 –1000, 1999.
616. Kronert WA, Edwards KA, Roche ES, Wells L, and Bernstein
SI. Muscle specific accumulation of Drosophila myosin heavy
chains—a splicing mutation in an alternative exon results in an
isoform substitution. EMBO J 10: 2479 –2488, 1991.
617. Kronert WA, O’Donnell PT, and Bernstein SI. A charge
change in an evolutionarily conserved region of the myosin globular head prevents myosin and thick filament accumulation in
Drosophila. J Mol Biol 236: 697–702, 1994.
618. Kronert WA, Odonnell PT, Fieck A, Lawn A, Vigoreaux JO,
Sparrow JC, and Bernstein SI. Defects in the Drosophila myosin rod permit sarcomere assembly but cause flight muscle degeneration. J Mol Biol 249: 111–125, 1995.
Physiol Rev • VOL
1043
619. Kubo S. Squid tropomyosins. Mem Fac Fish Hokkaido Univ 9:
57– 83, 1961.
620. Kugino M, Kugino K, and Ogawa T. Changes in microstructure
and rheological properties of squid mantle during storage. Food
Sci Technol Int Tokyo 3: 157–162, 1997.
621. Kulke M, Neagoe C, Kolmerer B, Minajeva A, Hinssen H,
Bullard B, and Linke WA. Kettin, a major source of myofibrillar
stiffness in Drosophila indirect flight muscle. J Cell Biol 154:
1045–1057, 2001.
622. Kurzawa-Goertz SE, Perreault-Micale CL, Trybus KM,
Szent-Györgyi AG, and Geeves MA. Loop I can modulate ADP
affinity, ATPase activity, and motility of different scallop myosins.
Transient kinetic analysis of S1 isoforms. Biochemistry 37: 7517–
7525, 1998.
623. Kusakabe R, Kusakabe T, Satoh N, Holland ND, and Holland
LZ. Differential gene expression and intracellular mRNA localization of amphioxus actin isoforms throughout development: implications for conserved mechanisms of chordate development. Dev
Genes Evol 207: 203–215, 1997.
624. Kusakabe R, Satoh N, Holland LZ, and Kusakabe T. Genomic
organization and evolution of actin genes in the amphioxus Branchiostoma belcheri and Branchiostoma floridae. Gene 227: 1–10,
1999.
625. Kusakabe T. Expression of larval type muscle actin encoding
genes in the ascidian Halocynthia roretzi. Gene 153: 215–218,
1995.
626. Kusakabe T. Ascidian actin genes: developmental regulation of
gene expression and molecular evolution. Zool Sci 14: 707–718,
1997.
627. Kusakabe T, Araki I, Satoh N, and Jeffery WR. Evolution of
chordate actin genes: evidence from genomic organization and
amino acid sequences. J Mol Evol 44: 289 –298, 1997.
628. Kusakabe T, Hikosaka A, and Satoh N. Coexpression and
promoter function in two muscle actin gene complexes of different structural organization in the ascidian Halocynthia roretzi.
Dev Biol 169: 461– 472, 1995.
629. Kusakabe T, Makabe KW, and Satoh N. Tunicate muscle actin
genes-structure and organization as a gene cluster. J Mol Biol 227:
955–960, 1992.
630. Kusakabe T, Suzuki J, Saiga H, Jeffery WR, Makabe KW, and
Satoh N. Temporal and spatial expression of a muscle actin gene
during embryogenesis of the ascidian Haolcynthia roretzi. Dev
Growth Differ 33: 227–234, 1991.
631. Kusakabe T, Swalla BJ, Satoh N, and Jeffery WR. Mechanism
of an evolutionary change in muscle cell differentiation in ascidians with different modes of development. Dev Biol 174: 379 –392,
1996.
632. Kusakabe T, Yoshida R, Kawakami I, Kusakabe R, Mochizuki
Y, Yamada L, Shin-i T, Kohara Y, Satoh N, Tsuda M, and
Satou Y. Gene expression profiles in tadpole larvae of Ciona
intestinalis. Dev Biol 242: 188 –203, 2002.
633. Labouesse M and Georges-Labouesse E. Cell adhesion: parallels between vertebrate and invertebrate focal adhesions. Curr
Biol 12: R528 –R530, 2003.
634. Labuhn M and Brack C. Age related changes in the mRNA
expression of actin isoforms in Drosophila melanogaster. Gerontology 43: 261–267, 1997.
635. Laclette JP, Landa A, Arcos L, Willms K, Davis AE, and
Shoemaker CB. Paramyosin is the Schistosoma mansoni (Trematoda) homolog of antigen-B from Taenia solium (Cestoda). Mol
Biochem Parasitol 44: 287–296, 1991.
636. Laclette JP, Shoemaker CB, Richter D, Arcos L, Pante N,
Cohen C, Bing D, and Nicholsonweller A. Paramyosin inhibits
complement C1. J Immunol 148: 124 –128, 1992.
637. Laclette JP, Skelly PJ, Merchant MT, and Shoemaker CB.
Aldehyde fixation dramatically alters the immunolocalization pattern of paramyosin in platyhelminth parasites. Exp Parasitol 81:
140 –143, 1995.
638. LaFramboise W, Griffis B, Bonner P, Warren W, Scalise D,
Guthrie RD, and Cooper RL. Muscle type specific myosin isoforms in crustacean muscle. J Exp Zool 286: 36 – 48, 1999.
639. Lakey A, Ferguson C, Labeit S, Reedy M, Larkins A, Butcher
G, Leonard K, and Bullard B. Identification and localization of
85 • JULY 2005 •
www.prv.org
1044
640.
641.
642.
643.
644.
645.
646.
647.
648.
649.
650.
651.
652.
653.
654.
655.
656.
657.
658.
659.
660.
661.
662.
663.
SCOTT L. HOOPER AND JEFFREY B. THUMA
high molecular weight proteins in insect flight and leg muscle.
EMBO J 9: 3459 –3467, 1990.
Lakey A, Labeit S, Gautel M, Ferguson C, Barlow DP, Leonard K, and Bullard B. Kettin, a large modular protein in the Z
disc of insect muscles. EMBO J 12: 2863–2871, 1993.
Laki K. A simple method for the isolation and crystallization of
tropomyosin from the muscles of the clam, Venus mercenaria.
Arch Biochem Biophys 67: 240 –242, 1957.
Laki K, Horváth B, and Klatzo I. On the relationship between
myosin and tropomyosin A. Biochim Biophys Acta 28: 656 – 657,
1958.
Lamers AE, Heiney JP, and Ram JL. Isolation and characterization of a cDNA encoding an actin protein from the zebra
mussel, Dreissena polymorpha. J Shellfish Res 17: 1215–1217,
1998.
Lanar DE, Pearce EJ, James SL, and Sher A. Identification of
paramyosin as a schistosome antigen recognized by intradermally
vaccinated mice. Science 234: 593–596, 1986.
Landa A, Laclette JP, Nicholson-Weller A, and Shoemaker
CB. cDNA cloning and recombinant expression of collagen-binding and complement inhibitor activity of Taenia solium paramyosin (AgB). Mol Biochem Parasitol 60: 343–347, 1993.
Landel CP, Krause MW, Waterston RH, and Hirsh D. DNA
rearrangements in the actin gene cluster in C. elegans accompany
reversion of three muscle mutants. J Mol Biol 180: 497–513, 1984.
Lang AB, Wyss C, and Epstein HF. Lack of actin III in fibrillar
flight muscle of flightless Drosophila mutant raised. Nature 291:
506 –508, 1981.
Langy S, Plichart C, and Luquiaud P. The immunodominant
Brugia malayi paramyosin as a marker of current infection with
Wuchereria bancrofti adult worms. Infect Immun 66: 2854 –2858,
1998.
Leake MC, Wilson D, Bullard B, and Simmons RM. The elasticity of single kettin molecules using a two-bead laser tweezers
assay. FEBS Lett 535: 55– 60, 2003.
Lee JJ, Calzone FJ, Britten RJ, Angerer RC, and Davidson
EH. Activation of sea urchin actin genes during embryogenesis. J
Mol Biol 188: 173–183, 1986.
Lee JJ, Shott RJ, Rose SJ III, Thomas TL, Britten RJ, and
Davidson EH. Sea urchin actin gene subtypes. Gene number,
linkage and evolution. J Mol Biol 172: 149 –176, 1984.
Lee JS, Lee J, Park SJ, and Yong TS. Analysis of the genes
expressed in Clonorchis sinensis adults using the expressed sequence tag approach. Parasitol Res 91: 283–289, 2003.
Lees-Miller JP and Helfman DM. The molecular basis for tropomyosin isoform diversity. Bioessays 13: 429 – 437, 1991.
Lehman W. Phylogenetic diversity of the proteins regulating muscular contraction. Int Rev Cytol 44: 55–92, 1976.
Lehman W. The ionic requirements for regulation by molluscan
thin filaments. Biochim Biophys Acta 745: 1–5, 1983.
Lehman W and Ferrell M. Phylogenetic diversity of troponin
subunit-C amino acid composition. FEBS Lett 121: 273–274, 1980.
Lehman W, Head JF, and Grant PW. The stoichiometry and
location of troponin I- and troponin C-like proteins in the myofibril of the bay scallop, Aequipecten irradians. Biochem J 187:
447– 456, 1980.
Lehman W, Kendrick-Jones J, and Szent-Györgyi AG. Myosin-linked regulatory systems: comparative studies. Cold Spring
Harb Symp 37: 319 –330, 1972.
Lehman W, Registein JM, and Ransom AL. The stoichiometry
of the components of arthropod thin filaments. Biochim Biophys
Acta 434: 215–222, 1976.
Lehman W and Szent-Györgyi AG. Regulation of muscular contraction. Distribution of actin control and myosin control in the
animal kingdom. J Gen Physiol 66: 1–30, 1975.
Lehrer SB, Ayuso R, and Reese G. Seafood allergy and allergens: a review. Mar Biotechnol 5: 339 –348, 2003.
Lehrer SB, Ayuso R, and Reese G. Current understanding of
food allergens. Ann NY Acad Sci 964: 69 – 85, 2002.
Lehrer SB and Reese G. Cross-reactivity between cockroach
allergens and arthropod, nematode, and mammalian allergens.
Rev Fr Allergol 38: 846 – 850, 1998.
Physiol Rev • VOL
664. Lei JY, Tang XX, Chambers TC, Pohl J, and Benian GM.
Protein kinase domain of twitchin has protein kinase activity and
an autoinhibitory region. J Biol Chem 269: 21078 –21085, 1994.
665. Leicht BG, Lyckegaard EMS, Benedict CM, and Clark AG.
Conservation of alternative splicing and genomic organization of
the myosin alkali light chain (MLC1) gene among Drosophila
species. Mol Biol Evol 10: 769 –790, 1993.
666. Leptin M, Bogaert T, Lehmann R, and Wilcox M. The function
of PS integrins during Drosophila embryogenesis. Cell 56: 401–
408, 1989.
667. Leroux MR and Candido EPM. Subunit characterization of the
Caenorhabditis elegans chaperonin containing TCP-1 and expression pattern of the gene encoding CCT-1. Biochem Biophys Res
Commun 241: 687– 692, 1997.
668. Leung PS, Chu KH, Chow WK, Ansari A, Bandea CI, Kwan
HS, Nagy SM, and Gershwin ME. Cloning, expression, and
primary structure of Metapenaeus ensis tropomyosin, the major
heat-stable shrimp allergen. J Allergy Clin Immunol 94: 882– 890,
1994.
669. Leung PSC, Chen YC, Gershwin ME, Wong H, Kwan HS, and
Chu KH. Identification and molecular characterization of Charybdis feriatus tropomyosin, the major crab allergen. J Allergy Clin
Immunol 102: 847– 852, 1998.
670. Leung PSC, Chen YC, Mykles DL, Chow WK, Li CP, and Chu
KH. Molecular identification of the lobster muscle protein tropomyosin as a seafood allergen. Mol Mar Biol Biotech 7: 12–20, 1998.
671. Leung PSC, Chow WK, Duffey S, Kwan HS, Gershwin ME,
and Chu KH. IgE reactivity against a cross-reactive allergen in
crustacea and mollusca: evidence for tropomyosin as the common
allergen. J Allergy Clin Immunol 98: 954 –961, 1996.
672. Leung PSC and Chu KH. cDNA cloning and molecular identification of the major oyster allergen from the Pacific oyster Crassostrea gigas. Clin Exp Allergy 31: 1287–1294, 2001.
673. Levin JZ and Horvitz HR. Three new classes of mutations in the
Caenorhabditis elegans muscle gene sup-9. Genetics 135: 53–70,
1993.
674. Levine RJC, Chantler PD, Kensler RW, and Woodhead JL.
Effects of phosphorylation by myosin light chain kinase on the
structure of Limulus thick filaments. J Cell Biol 113: 563–572,
1991.
675. Levine RJC, Dewey MM, and de Villafranca GW. Immunohistochemical localization of contractile proteins in Limulus striated
muscle. J Cell Biol 55: 221–235, 1972.
676. Levine RJC, Dewey MM, Elfvin M, and Walcott B. Lethocerus
flight muscle paramyosin: antibody localization and electrophoretic studies. Am J Anat 141: 453– 458, 1974.
677. Levine RJC, Elfvin M, Dewey MM, and Walcott B. Paramyosin
in invertebrate muscles. II. Content in relation to structure and
function. J Cell Biol 71: 273–279, 1976.
678. Lewis JA, Wu CH, Berg H, and Levine JH. The genetics of
levamisole resistance in the nematode Caenorhabditis elegans.
Genetics 95: 905–928, 1980.
679. Lewis JA, Wu CH, Levine JH, and Berg H. Levamisole resistant
mutants of the nematode Caenorhabditis elegans appear to lack
pharmacological acetylcholine receptors. Neuroscience 5: 967–
989, 1980.
680. Li BW, Chandrashekar R, Alvarez RM, Liftis W, and Weil GJ.
Identification of paramyosin as a potential protective antigen
against Brugia malayi infection in jirds. Mol Biochem Parasitol
49: 315–325, 1991.
681. Li BW, Chandrashekar R, and Weil GJ. Vaccination with recombinant filarial paramyosin induces partial immunity to Brugia
malayi infection in birds. J Immunol 150: 1881–1885, 1993.
682. Li BW, Hoppe PE, and Weil GJ. Cloning of an early immunodominant filarial antigen—a member of the Brugia malayi myosin heavy chain gene family. Int J Parasitol 25: 611– 619, 1995.
683. Li BW, Zhang SR, Curtis KC, and Weil GJ. Immune responses
to Brugia malayi paramyosin in rodents after DNA vaccination.
Vaccine 18: 76 – 81, 1999.
684. Li Y and Mykles DL. Analysis of myosins from lobster muscles:
fast and slow isozymes differ in heavy-chain composition. J Exp
Zool 255: 163–170, 1990.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
685. Li YS, Sleigh AC, Ross AGP, Li Y, Zhang XY, Williams GM, Yu
XL, Tanner M, and McManus DP. Human susceptibility to
Schistosoma japonicum in China correlates with antibody isotypes to native antigens. Trans R Soc Trop Med Hyg 95: 441– 448,
2001.
686. Lichter JB and Storti RV. In vitro transcription analysis of the
Drosophila tropomyosin and other muscle genes. Biochim Biophys Acta 1088: 419 – 424, 1991.
687. Lightner DV. Normal post mortem changes in the brown shrimp,
Panaeus aztecus. Fish Bull 72: 223–236, 1973.
688. Limberger RJ and McReynolds LA. Filarial paramyosin: cDNA
sequences from Dirofilaria immitis and Onchocerca volvulus.
Mol Biochem Parasitol 38: 271–280, 1990.
689. Lin MH, Nguyen HT, Dybala C, and Storti RV. Myocyte specific enhancer factor 2 acts cooperatively with a muscle activator
region to regulate Drosophila tropomyosin gene muscle expression. Proc Natl Acad Sci USA 93: 4623– 4628, 1996.
690. Lin SC and Storti RV. Developmental regulation of the Drosophila tropomyosin I (TmI) gene is controlled by a muscle activator
enhancer region that contains multiple cis-elements and binding
sites for multiple proteins. Dev Genet 20: 297–306, 1997.
691. Lin X, Qadota H, Moerman DG, and Williams BD. C. elegans
PAT-6/actopaxin plays a critical role in the assembly of integrin
adhesion complexes in vivo. Curr Biol 13: 922–932, 2003.
692. Littlefield KP, Swank DM, Sanchez BM, Knowles AF, Warshaw DM, and Bernstein SI. The converter domain modulates
kinetic properties of Drosophila myosin. Am J Physiol Cell
Physiol 284: C1031–C1038, 2003.
693. Liu FZ, Barral JM, Bauer CC, Ortiz I, Cook RG, Schmid MF,
and Epstein HF. Assemblases and coupling proteins in thick
filament assembly. Cell Struct Funct 22: 155–162, 1997.
694. Liu FZ, Bauer CC, Ortiz I, Cook RG, Schmid MF, and Epstein
HF. ␤-Filagenin, a newly identified protein coassembling with
myosin and paramyosin in Caenorhabditis elegans. J Cell Biol
140: 347–353, 1998.
695. Liu FZ, Ortiz I, Hutagalung A, Bauer CC, Cook RG, and
Epstein HF. Differential assembly of ␣- and ␥-filagenins into
thick filaments in Caenorhabditis elegans. J Cell Sci 113: 4001–
4012, 2000.
696. Liu HJ, Mardahl-Dumesnil M, Sweeney ST, O’Kane CJ, and
Bernstein SL. Drosophila paramyosin is important for myoblast
fusion and essential for myofibril formation. J Cell Biol 160:
899 –908, 2003.
697. Lizotte-Waniewski M, Tawe W, Guiliano DB, Lu WH, Liu J,
Williams SA, and Lustigman S. Identification of potential vaccine and drug target candidates by expressed sequence tag analysis and immunoscreening of Onchocerca volvulus larval cDNA
libraries. Infect Immun 68: 3491–3501, 2000.
698. Locker RH and Wild DJC. A comparative study of high molecular weight proteins in various types of muscle across the animal
kingdom. J Biochem 99: 1473–1484, 1986.
699. Lopez de Haro MS, Salgado LM, David CN, and Bosch TC.
Hydra tropomyosin TROP1 is expressed in head specific epithelial cells and is a major component of the cytoskeletal structure
that anchors nematocytes. J Cell Sci 107: 1403–1411, 1994.
700. López-Moreno HS, Correa D, Laclette JP, and Ortiz-Navarrete VF. Identification of CD4⫹ T cell epitopes of Taenia
solium paramyosin. Parasite Immunol 25: 513–516, 2003.
701. Lou SN. Purine content in grass shrimp during storage as related
to freshness. J Food Sci 63: 442– 444, 1998.
702. Loukas M and Kafatos FD. The actin loci in the genus Drosophila: establishment of chromosomal homologies among distantly
related species by in situ hybridization. Chromosoma 94: 297–308,
1986.
703. Lovato TL, Meadows SM, Baker PW, Sparrow JC, and Cripps
RM. Characterization of muscle actin genes in Drosophila virilis
reveals significant molecular complexity in skeletal muscle types.
Insect Mol Biol 10: 333–340, 2001.
704. Lowey S, Kucera J, and Holtzer A. On the structure of the
paramyosin molecule. J Mol Biol 7: 234 –244, 1963.
705. Luan-Rilliet Y, Milos M, and Cox JA. Thermodynamics of
cation binding to Nereis sarcoplasmic calcium-binding protein.
Physiol Rev • VOL
706.
707.
708.
709.
710.
711.
712.
713.
714.
715.
716.
717.
718.
719.
720.
721.
722.
723.
724.
725.
726.
1045
Direct binding studies, microcalorimitry, and conformational
changes. Eur J Biochem 208: 133–138, 1992.
MacGregor AN and Shore SJ. Immunocytochemistry of cytoskeletal proteins in adult Schistosoma mansoni. Int J Parasitol
20: 279 –284, 1990.
Machado C and Andrew DJ. Titin as a chromosomal protein.
Adv Exp Med Biol 481: 221–236, 2000.
Machado C and Andrew DJ. D-titin: a giant protein with dual
roles in chromosomes and muscles. J Cell Biol 151: 639 – 652,
2000.
Macias MT and Sastre L. Molecular cloning and expression of
four actin isoforms during Artemia development. Nucleic Acids
Res 18: 5219 –5225, 1990.
Mackenzie JM Jr and Epstein HF. Paramyosin is necessary for
determination of nematode thick filament length in vivo. Cell 22:
747–755, 1980.
Mackenzie JM Jr, Schachat F, and Epstein HF. Immunocytochemical localization of two myosins within the same muscle cells
in Caenorhabditis elegans. Cell 15: 413– 419, 1978.
Mackinnon AC, Qadota H, Norman KR, Moerman DG, and
Williams BD. C. elegans PAT-4/ILK functions as an adaptor protein within integrin adhesion complexes. Curr Biol 12: 787–797,
2002.
MacKrell AJ, Blumberg B, Haynes SR, and Fessler JH. The
lethal myospheroid gene of Drosophila encodes a membrane
protein homologous to vertebrate integrin ␤ subunits. Proc Natl
Acad Sci USA 85: 2633–2637, 1988.
MacLean DW, Meedel TH, and Hastings KEM. Tissue specific
alternative splicing of ascidian troponin I isoforms. J Biol Chem
272: 32115–32120, 1997.
MacLeod AR, Karn J, and Brenner S. Molecular analysis of the
unc-54 myosin heavy chain gene of Caenorhabditis elegans. Nature 291: 386 –390, 1981.
MacLeod AR, Waterston RH, and Brenner S. An internal deletion mutant of a myosin heavy chain in Caenorhabditis elegans.
Proc Natl Acad Sci USA 74: 5336 –5340, 1977.
MacLeod AR, Waterston RH, Fishpool RM, and Brenner S.
Identification of the structural gene for a myosin heavy-chain in
Caenorhabditis elegans. J Mol Biol 114: 133–140, 1977.
Mádi A, Mikkat S, Ringel B, Thiesen HJ, and Glocker MO.
Profiling stage-dependent changes of protein expression in Caenorhabditis elegans by mass spectrometric proteome analysis
leads to the identification of stage-specific marker proteins. Electrophoresis 24: 1809 –1817, 2003.
Mádi A, Mikkat S, Ringel B, Ulbrich M, Thiesen HJ, and
Glocker MO. Mass spectrometric proteome analysis for profiling
temperature-dependent changes of protein expression in wildtype Caenorhabditis elegans. Proteomics 3: 1526 –1534, 2003.
Maeda M, Saitoh H, Itami T, Takahashi Y, Mizuki E, and
Ohba M. Isolation of the actin-encoding cDNA of kuruma shrimp
Marsupenaeus japonicus. J Crustacean Biol 22: 704 –707, 2002.
Magnay JL, Holmes JM, Neil DM, and El Haj AJ. Temperaturedependent developmental variation in lobster muscle myosin
heavy chain isoforms. Gene 316: 119 –126, 2003.
Mahaffey JW, Coutu MD, Fyrberg EA, and Inwood W. The
flightless Drosophila mutant raised has two distinct genetic lesions affecting accumulation of myofibrillar proteins in flight muscles. Cell 40: 101–110, 1985.
Maita T, Konno K, Maruta S, Norisue H, and Matsuda G.
Amino acid sequence of the essential light chain of adductor
muscle myosin from Ezo giant scallop, Patinopecten yessoensis.
J Biochem 102: 1141–1149, 1987.
Maita T, Konno K, Ojima T, and Matsuda G. Amino acid
sequences of the regulatory light chains of striated adductor muscle myosins from Ezo giant scallop and Akazara scallop. J Biochem 95: 167–177, 1984.
Maita T, Tanaka H, Konno K, and Matsuda G. Amino acid
sequence of the regulatory light chain of squid mantle muscle
myosin. J Biochem 102: 1151–1157, 1987.
Maizels RM, Holland MJ, Falcone FH, Zang XX, and Yazdanbakhsh M. Vaccination against helminth parasites-the ultimate
challenge for vaccinologists? Immunol Rev 171: 125–147, 1999.
85 • JULY 2005 •
www.prv.org
1046
SCOTT L. HOOPER AND JEFFREY B. THUMA
727. Maizels RM, Selkirk ME, and Agnew A. Prospects for new
vaccines against helminth parasites. Trends Biotechnol 7: 316 –
321, 1989.
728. Makabe KW, Fujiwara S, Saiga H, and Satoh N. Specific
expression of myosin heavy chain gene in muscle lineage cells of
the ascidian embryo. Roux’s Arch Dev Biol 199: 307–313, 1990.
729. Makabe KW and Satoh N. Temporal expression of myosin heavy
chain gene during ascidian embryogenesis. Dev Growth Differ 31:
71–77, 1989.
730. Maki S, Kimura S, and Maruyama K. Localization of connectinlike proteins in the giant sarcomeres of barnacle muscle. Zool Sci
11: 821– 824, 1994.
731. Maki S, Ohtani Y, Kimura S, and Maruyama K. Isolation and
characterization of a kettin-like protein from crayfish claw muscle. J Muscle Res Cell Motil 16: 579 –585, 1995.
732. Manabe T, Kawamura Y, Higuchi H, Kimura S, and Maruyama K. Connectin, giant elastic protein, in giant sarcomeres of
crayfish claw muscle. J Muscle Res Cell Motil 14: 654 – 665, 1993.
733. Manseau LJ, Ganetzky B, and Craig EA. Molecular and genetic
characterization of the Drosophila melanogaster 87E actin gene
region. Genetics 119: 407– 420, 1988.
734. Manuel H, Zobel CR, and Siemankowski RF. Effect of divalent
cations on the proteolysis of vertebrate and invertebrate muscle
myosin. Biochim Biophys Acta 626: 88 –96, 1980.
735. Marden JH, Fitzhugh GH, Girgenrath M, Wolf MR, and Girgenrath S. Alternative splicing, muscle contraction, and intraspecific variation: associations between troponin T transcripts, Ca2⫹
sensitivity, and the force and power output of dragonfly flight
muscles during oscillatory contraction. J Exp Biol 204: 3457–3470,
2001.
736. Marden JH, Fitzhugh GH, and Wolf MR. From molecules to
mating success: integrative biology of muscle maturation in a
dragonfly. Am Zool 38: 528 –544, 1998.
737. Marden JH, Fitzhugh GH, Wolf MR, Arnold KD, and Rowan
B. Alternative splicing, muscle calcium sensitivity, and the modulation of dragonfly flight performance. Proc Natl Acad Sci USA
96: 15304 –15309, 1999.
738. Margulis BA, Bobrova IF, Mashanski VF, and Pinaev GP.
Major myofibrillar protein content and the structure of mollusc
adductor contractile apparatus. Comp Biochem Physiol A Physiol
64: 291–298, 1979.
739. Margulis BA, Galceran J, Podgornaya OI, and Pinaev GP.
Major contractile proteins of mollusc: tissue polymorphism of
actin, tropomyosin, and myosin light chains is absent. Comp
Biochem Physiol B Biochem 72: 473– 476, 1982.
740. Margulis BA and Pinaev GP. The species specifity of the contractile protein composition of the bivalve molluscs. Comp Biochem Physiol B Biochem 55: 189 –194, 1976.
741. Marı́n MC, Rodrı́iguez JR, and Ferrus A. Transcription of
Drosophila troponin I gene is regulated by two conserved, functionally identical, synergistic elements. Mol Biol Cell 15: 1185–
1196, 2004.
742. Mariol MC and Segalat L. Muscular degeneration in the absence
of dystrophin is a calcium-dependent process. Curr Biol 11: 1691–
1694, 2001.
743. Maroto M, Arredondo J, Goulding D, Marco R, Bullard B,
and Cervera M. Drosophila paramyosin/miniparamyosin gene
products show a large diversity in quantity, localization, and isoform pattern: a possible role in muscle maturation and function.
J Cell Biol 134: 81–92, 1996.
744. Maroto M, Arredondo JJ, Roman MS, Marco R, and Cervera
M. Analysis of the paramyosin/miniparamyosin gene. Miniparamyosin is an independently transcribed, distinct paramyosin
isoform, widely distributed in invertebrates. J Biol Chem 270:
4375– 4382, 1995.
745. Maroto M, Vinós J, Marco R, and Cervera M. Autophosphorylating protein kinase activity in titin-like arthropod projectin. J
Mol Biol 224: 287–291, 1992.
746. Marsden BJ, Shaw GS, and Sykes BD. Calcium binding proteins. Elucidating the contributions to calcium affinity from an
analysis of species variants and peptide fragments. Biochem Cell
Biol 68: 587– 601, 1990.
Physiol Rev • VOL
747. Martin RE, Masaracchia RA, and Donahue MJ. Ascaris suum:
regulation of myosin light chain phosphorylation from adult skeletal muscle. Exp Parasitol 61: 114 –119, 1986.
748. Martin RM, Chilton NB, Lightowlers MW, and Gasser RB.
Echinococcus granulosus myophilin-relationship with protein homologues containing “calponin-motifs.” Int J Parasitol 27: 1561–
1567, 1997.
749. Martin RM, Colebrook AL, Gasser RB, and Lightowlers MW.
Antibody response of patients with cystic hydatid disease to recombinant myophilin of Echinococcus granulosus. Acta Trop 61:
307–314, 1996.
750. Martin RM, Csar XF, Gasser RB, Felleisen R, and Lightowlers MW. Myophilin of Echinococcus granulosis: isoforms and
phosphorylation by protein kinase C. Parasitology 115: 205–211,
1997.
751. Martin RM, Gasser RB, Jones MK, and Lightowlers MW.
Identification and characterization of myophilin, a muscle specific
antigen of Echinococcus granulosus. Mol Biochem Parasitol 70:
139 –148, 1995.
752. Martı́nez A, Martı́nez J, Palacios R, and Panzani R. Importance of tropomyosin in the allergy to household arthropods.
Cross-reactivity with other invertebrate extracts. Allergol Immunopathol 25: 118 –126, 1997.
753. Martinez I, Friis TJ, and Careche M. Post mortem muscle
protein degradation during ice storage of Arctic (Pandalus borealis) and tropical (Penaeus japonicus and Penaeus monodon)
shrimps: a comparative electrophoretic and immunological study.
J Sci Food Agr 81: 1199 –1208, 2001.
754. Maruyama IN, Miller DM, and Brenner S. Myosin heavy chain
gene amplification as a suppressor mutation in Caenorhabditis
elegans. Mol Gen Genet 219: 113–118, 1989.
755. Maruyama K. Connectin/titin, giant elastic protein of muscle.
FASEB J 11: 341–345, 1997.
756. Maruyama K. Comparative aspects of muscle elastic proteins.
Rev Physiol Biochem Pharmacol 138: 1–18, 1999.
757. Maruyama K, Cage PE, and Bell JL. The role of connectin in
elastic properties of insect flight muscle. Comp Biochem Physiol
A Physiol 61: 623– 627, 1978.
758. Maruyama K and Kimura S. Connectin: from regular to giant
sizes of sarcomeres. Adv Exp Med Biol 481: 25–33, 2000.
759. Maruyama K and Nonomura Y. High molecular weight calcium
binding protein in the microsome of scallop striated muscle.
J Biochem 96: 859 – 870, 1984.
760. Mas JA, Garcı́a-Zaragoza E, and Cervera M. Two functionally
identical modular enhancers in Drosophila troponin T gene establish the correct protein levels in different muscle types. Mol
Biol Cell 15: 1931–1945, 2004.
761. Mathew D, Popescu A, and Budnik V. Drosophila amphiphysin
functions during synaptic fasciclin II membrane cycling. J Neurosci 23: 10710 –10716, 2003.
762. Matsumoto JJ. An electrophoretic study of the squid actomyosin. Nippon Suisan Gakk 25: 27–37, 1959.
764. Matsumoto JJ, Dyer WJ, Dingle JR, and Ellis DG. Protein in
extracts of prerigor and postrigor scallop striated muscle. J Fish
Res Board Can 24: 873– 882, 1967.
765. Matsumoto M and Yamanaka H. Studies on rigor mortis of
kurama prawn muscle. Nippon Suisan Gakk 57: 2121–2126, 1991.
766. Matsuno A, Kannda M, and Okuda M. Ultrastructural studies
on paramyosin core filaments from native thick filaments in catch
muscles. Tissue Cell 28: 501–505, 1996.
767. Matsuno A, Takanoohmuro H, Itoh Y, Matsuura T, Shibata
M, Nakae H, Kaminuma T, and Maruyama K. Anti-connectin
monoclonal antibodies that react with the unc-22 gene product
bind dense bodies of Caenorhabditis (nematode) body-wall muscle cells. Tissue Cell 21: 495–505, 1989.
768. Matulef K, Sirokman K, Perreault-Micale CL, and SzentGyörgyi AG. Amino-acid sequence of squid myosin heavy chain.
J Muscle Res Cell Motil 19: 705–712, 1998.
769. Mawuenyega KG, Kaji H, Yamauchi Y, Shinkawa T, Saito H,
Taoka M, Takahashi N, and Isobe T. Large-scale identification
of Caenorhabditis elegans proteins by multidimensional liquid
chromatography-tandem mass spectrometry. J Proteome Res 2:
23–35, 2003.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
770. Maxwell-Miller G, Josephson RV, Spindler AA, HollowayThomas D, Avery MW, and Phleger CF. Chilled (5°C) and
frozen (⫺18°C) storage stability of the purple-hinge rock scallop,
Hinnites multirugosus Gale. J Food Sci 1982: 47–1654, 1982.
771. McArdle K, Allen TS, and Bucher EA. Ca2⫹-dependent muscle
dysfunction caused by mutation of the Caenorhabditis elegans
troponin T-1 gene. J Cell Biol 143: 1201–1213, 1998.
772. McCann RO and Craig SW. The I/LWEQ module: a conserved
sequence that signifies F actin binding in functionally diverse
proteins from yeast to mammals. Proc Natl Acad Sci USA 94:
5679 –5684, 1997.
773. McCarter JP, Mitreva MD, Martin J, Dante M, Wylie T, Rao
U, Pape D, Bowers Y, Theising B, Murphy CV, Kloek AP,
Chiapelli BJ, Clifton SW, Bird DM, and Waterston RH. Analysis and functional classification of transcripts from the nematode
Meloidogyne incognita. Genome Biol 4: R26.1–R26.19, 2003.
774. McCombie WR, Adam MD, Kelley JM, FitzGerald MG, Utterback TR, Khan M, Dubnick M, Kerlavage AR, Venter JC, and
Fields C. Caenorhabditis elegans expressed sequence tags identify gene families and potential disease gene homologues. Nature
Genet 1: 124 –131, 1992.
775. McCubbin WD and Kay CM. The subunit structure of fibrous
muscle proteins as determined by osmometry. Biochim Biophys
Acta 154: 239 –241, 1968.
776. McLachlan AD. Structural implications of the myosin amino acid
sequence. Annu Rev Biophys Bioeng 13: 167–189, 1984.
777. McManus DP. The search for a vaccine against schistosomiasis—a difficult path but an achievable goal. Immunol Rev 171:
149 –161, 1999.
778. McManus DP and Bartley PB. A vaccine against Asian schistosomiasis. Parasitol Int 53: 163–173, 2004.
779. McManus DP, Liu S, Song G, Xu Y, and Wong JM. The vaccine
efficacy of native paramyosin A(Sj-97) against Chinese Schistosoma japonicum. Int J Parasitol 28: 1739 –1742, 1998.
780. McManus DP, Wong JYM, Zhou J, Cai C, Zeng Q, Smyth D, Li
Y, Kalinna BH, Duke MJ, and Yi X. Recombinant paramyosin
(rec-Sj-97) tested for immunogenicity and vaccine efficacy against
Schistosoma japonicum in mice and water buffaloes. Vaccine 20:
870 – 878, 2002.
781. McNeill PA and Hoyle G. Evidence for superthin filaments. Am
Zool 7: 483– 498, 1967.
782. Medler S, Lilley T, and Mykles DL. Fiber polymorphism in
skeletal muscles of the American lobster, Homarus americanus:
continuum between slow-twitch (S-1) and slow-tonic (S-2) fibers.
J Exp Biol 207: 2755–2767, 2004.
783. Medler S and Mykles DL. Analysis of myofibrillar protein and
transcripts in adult skeletal muscles of the American lobster
Homarus americanus: variable expression of myosins, actin, and
troponins in fast, slow-twitch, and slow-tonic fibres. J Exp Biol
206: 3557–3567, 2003.
784. Meedel TH. Myosin expression in the developing ascidian embryo. J Exp Zool 227: 203–211, 1983.
785. Meedel TH and Hastings KEM. Striated muscle type tropomyosin in a chordate smooth muscle, ascidian body-wall muscle.
J Biol Chem 268: 6755– 6764, 1993.
786. Melson GL and Cowgill RW. Comparison of the muscle protein
paramyosin from different molluscan species. Comp Biochem
Physiol B Biochem 55: 503–510, 1976.
787. Mercer KB, Flaherty DB, Miller RK, Qadota H, Tinley TL,
Moerman DG, and Benian GM. Caenorhabditis elegans unc-98,
a C2H2Zn finger protein, is a novel partner of unc-97/PINCH in
muscle adhesion complexes. Mol Biol Cell 14: 2492–2507, 2003.
788. Meredith J and Storti RV. Developmental regulation of the
Drosophila tropomyosin II gene in different muscles is controlled
by muscle type specific intron enhancer elements and distal and
proximal promoter control elements. Dev Biol 159: 500 –512, 1993.
789. Merrick JP and Johnson WH. Solubility properties of ␣-reduced
paramyosin. Biochemistry 16: 2260 –2264, 1977.
790. Miegel A, Kobayashi T, and Maeda Y. Isolation, purification,
and partial characterization of tropomyosin and troponin subunits
from the lobster tail muscle. J Muscle Res Cell Motil 13: 608 – 618,
1992.
Physiol Rev • VOL
1047
791. Miller BM, Nyitrai M, Bernstein SI, and Geeves MA. Kinetic
analysis of Drosophila muscle myosin isoforms suggests a novel
mode of mechanochemical coupling. J Biol Chem 278: 50293–
50300, 2003.
792. Miller DM, Ortiz I, Berliner GC, and Epstein HF. Differential
localization of two myosins within nematode thick filaments. Cell
34: 477– 490, 1983.
793. Miller DM, Stockdale FE, and Karn J. Immunological identification of the genes encoding the four myosin heavy chain isoforms of Caenorhabditis elegans. Proc Natl Acad Sci USA 83:
2305–2309, 1986.
794. Miller RC, Schaaf R, Maughan DW, and Tansey TR. A nonflight muscle isoform of Drosophila tropomyosin rescues an indirect flight muscle tropomyosin mutant. J Muscle Res Cell Motil
14: 85–98, 1993.
795. Milstein CP. Tryptic digestion of tropomyosin A, tropomyosin B,
and myosin. Nature 209: 614 – 615, 1966.
796. Milstein CP. Isolation of Aulacomya paramyosin. Biochim Biophys Acta 103: 634 – 640, 1967.
797. Milstein CP and Bailey K. Isolation and characterisation of a
tryptic core from the insoluble tropomyosin of Pinna nobilis.
Biochim Biophys Acta 49: 412– 413, 1961.
798. Mitreva M, McCarter JP, Martin J, Dante M, Wylie T, Chiapelli B, Pape D, Clifton SW, Nutman TB, and Waterston RH.
Comparative genomics of gene expression in the parasitic and
free-living nematodes Strongyloides stercoralis and Caenorhabditis elegans. Genome Res 14: 209 –220, 2004.
799. Miyamoto H, Hamaguchi M, and Okoshi K. Analysis of genes
expressed in the mantle of oyster Crassostrea gigas. Fisheries Sci
68: 651– 658, 2002.
800. Miyanishi T, Maita T, Morita F, Kondo S, and Matsuda G.
Amino acid sequences of the two kinds of regulatory light chains
of adductor smooth muscle myosin from Patinopecten yessoensis. J Biochem 97: 541–551, 1985.
801. Miyazaki JI, Hosoya M, Ishimoda-Takagi T, and Hirabayashi
T. Tissue specificity of tropomyosin from the crayfish, Cambarus
clarki. J Biochem 108: 59 – 65, 1990.
802. Miyazaki JI, Ishimoda-Takagi T, Sekiguchi K, and Hirbayashi R. Comparative study of horseshoe crab tropomyosin.
Comp Biochem Physiol B Biochem 93: 681– 687, 1989.
803. Miyazaki JI, Makioka T, Fujiwara Y, and Hirabayashi T.
Tissue specificity of crustacean tropomyosin. J Exp Zool 263:
235–244, 1992.
804. Miyazaki JI, Sekiguchi K, and Hirabayashi T. Tissue specificity of tropomyosin from a horseshoe crab, Tachypleus tridentatus. Comp Biochem Physiol B Biochem 85: 679 – 685, 1986.
805. Miyazaki JI, Yahata K, Makioka T, and Hirabayashi T. Tissue
specificity of arthropod tropomyosin. J Exp Zool 267: 501–509,
1993.
806. Miyazawa H, Fukamachi H, Inagaki Y, Reese G, Daul CB,
Lehrer SB, Inouye S, and Sakaguchi M. Identification of the
first major allergen of a squid (Todarodes pacificus). J Allergy
Clin Immunol 98: 948 –953, 1996.
807. Mochizuki Y, Mizuno H, Ogawa H, and Iso N. Determination of
the ratio of myosin and actin in raw meat by differential scanning
calorimetry. Fisheries Sci 61: 723–724, 1995.
808. Mochizuki Y, Mizuno H, Ogawa H, Ishimura K, Tsuchiya H,
Fukuzawa M, and Iso N. Rheological properties of cuttlefish and
squid raw meat. Fisheries Sci 60: 555–558, 1994.
809. Moerman DG and Baillie DL. Genetic organization in Caenorhabditis elegans: fine-structure analysis of the unc-22 gene. Genetics 91: 95–103, 1979.
810. Moerman DG, Benian GM, Barstead RJ, Schriefer LA, and
Waterston RH. Identification and intracellular localization of the
unc-22 gene product of Caenorhabditis elegans. Genes Dev 2:
93–105, 1988.
811. Moerman DG, Benian GM, and Waterston RH. Molecular cloning of the muscle gene unc-22 in Caenorhabditis elegans by Tc1
transposon tagging. Proc Natl Acad Sci USA 83: 2579 –2583, 1986.
812. Moerman DG, Plurad S, Waterston RH, and Baillie DL. Mutations in the unc-54 myosin heavy chain gene of Caenorhabditis
elegans that alter contractility but not muscle structure. Cell 29:
773–781, 1982.
85 • JULY 2005 •
www.prv.org
1048
SCOTT L. HOOPER AND JEFFREY B. THUMA
813. Moerman DG and Waterston RH. Spontaneous unstable unc-22
IV mutations in C. elegans var Bergerac. Genetics 108: 859 – 877,
1984.
814. Mogami K, Fujita SC, and Hotta Y. Identification of Drosophila
indirect flight muscle myofibrillar proteins by means of two-dimensional electrophoresis. J Biochem 91: 643– 650, 1982.
815. Mogami K and Hotta Y. Isolation of Drosophila flightless mutants which affect myofibrillar proteins of indirect flight muscle.
Mol Gen Genet 183: 409 – 417, 1981.
816. Mogami K, O’Donnell PT, Bernstein SI, Wright TR, and Emerson CP. Mutations of the Drosophila myosin heavy chain gene:
effects on transcription, myosin accumulation, and muscle function. Proc Natl Acad Sci USA 83: 1393–1397, 1986.
817. Molloy J, Kreuz A, Miller R, Tansey T, and Maughan D.
Effects of tropomyosin deficiency in flight muscle of Drosophila
melanogaster. Adv Exp Med Biol 332: 165–172, 1993.
818. Moncrief ND, Kretsinger RH, and Goodman M. Evolution of
EF-hand calcium-modulated proteins. I. Relationships based on
amino acid sequences. J Mol Evol 30: 522–562, 1990.
819. Montana ES and Littleton JT. Characterization of a hypercontraction-induced myopathy in Drosophila caused by mutations in
Mhc. J Cell Biol 164: 1045–1054, 2004.
820. Moore JR, Dickinson MH, Vigoreaux JO, and Maughan DW.
The effect of removing the N-terminal extension of the Drosophila
myosin regulatory light chain upon flight ability and the contractile dynamics of the indirect flight muscle. Biophys J 78: 1431–
1440, 2000.
821. Moore JR, Vigoreaux JO, and Maughan DW. The Drosophila
projectin mutant, bent(D), has reduced stretch activation and
altered indirect flight muscle kinetics. J Muscle Res Cell Motil 20:
797– 806, 1999.
822. Moreira JE, Dodane V, and Reese TS. Immunoelectronmicroscopy of soluble and membrane proteins with a sensitive postembedding method. J Histochem Cytochem 46: 847– 854, 1998.
823. Morgan NS. The myosin superfamily in Drosophila melanogaster. J Exp Zool 273: 104 –117, 1995.
824. Morita F and Kondo S. Regulatory light chain contents and
molecular species of myosin in catch muscle of scallop. J Biochem 92: 977–983, 1982.
825. Moulder GL, Huang MM, Waterston RH, and Barstead RJ.
Talin requires ␤-integrin, but not vinculin, for its assembly into
focal adhesion-like structures in the nematode Caenorhabditis
elegans. Mol Biol Cell 7: 1181–1193, 1996.
826. Mounier N, Gouy M, Mouchiroud D, and Prudhomme JC.
Insect muscle actins differ distinctly from invertebrate and vertebrate cytoplasmic actins. J Mol Evol 34: 406 – 415, 1992.
827. Mounier N and Prudhomme JC. Isolation of actin genes in
Bombyx mori: the coding sequence of a cytoplasmic actin gene
expressed in the silk gland is interrupted by a single intron in an
unusual position. Biochimie 68: 1053–1061, 1986.
828. Mounier N and Prudhomme JC. Differential expression of muscle and cytoplasmic actin genes during development of Bombyx
mori. Insect Biochem 21: 523–533, 1991.
829. Mounier N and Sparrow JC. Muscle actin genes in insects.
Comp Biochem Physiol B Biochem 105: 231–238, 1993.
830. Mühlschlegel F, Sygulla L, Frosch P, Massetti P, and Frosch
M. Paramyosin of Echinococcus granulosus: cDNA sequence and
characterization of a tegumental antigen. Parasitol Res 79: 660 –
666, 1993.
831. Müller W. Beobachtungen des patologischen Instituts zu Jena. 1.
über den Bau der Chorda dorsalis. Jen Z Naturwiss N: 327–353,
1871.
832. Muñoz D, Jimenez A, Marinotti O, and James AA. The AeAct-4 gene is expressed in the developing flight muscles of female
Aedes aegypti. Insect Biochem Mol Biol 13: 563–568, 2004.
833. Myers CD, Goh PY, Allen TS, Bucher EA, and Bogaert T.
Developmental genetic analysis of troponin T mutations in striated and nonstriated muscle cells of Caenorhabditis elegans.
J Cell Biol 132: 1061–1077, 1996.
834. Mykles DL. Heterogeneity of myofibrillar proteins in lobster fast
and slow muscles: variants of troponin, paramyosin, and myosin
light chains comprise four distinct protein assemblages. J Exp
Zool 234: 23–32, 1985.
Physiol Rev • VOL
835. Mykles DL. Histochemical and biochemical characterization of
two slow fiber types in decapod crustacean muscles. J Exp Zool
245: 232–243, 1988.
836. Mykles DL. Multiple variants of myofibrillar proteins in single
fibers of lobster claw muscles: evidence for two types of slow
fibers in the cutter closer muscle. Biol Bull 169: 476 – 483, 1985.
837. Mykles DL, Cotton JLS, Taniguchi H, Sano KI, and Maeda Y.
Cloning of tropomyosins from lobster (Homarus americanus)
striated muscles: fast and slow isoforms may be generated from
the same transcript. J Muscle Res Cell Motil 19: 105–115, 1998.
838. Mykles DL, Medler S, Koenders A, and Cooper R. Myofibrillar
protein isoform expression is correlated with synaptic efficacy in
slow fibres of the claw and leg opener muscles of crayfish and
lobster. J Exp Biol 205: 513–522, 2002.
839. Nagai T, Hamada T, Kai N, Tanouye Y, and Nagayama F.
Differential scanning calorimetry of several jellyfish mesogloea.
Fisheries Sci 63: 459 – 461, 1997.
840. Nagashima Y, Ebina H, Nagai T, Tanaka M, and Taguchi T.
Proteolysis affects thermal gelation of squid mantle muscle. J
Food Sci 57: 916 –922, 1992.
841. Naimi B, Harrison A, Cummins M, Nongthomba U, Clark S,
Canal I, Ferrus A, and Sparrow JC. A tropomyosin-2 mutation
suppresses a troponin I myopathy in Drosophila. Mol Biol Cell 12:
1529 –1539, 2001.
842. Nakada T, Nagano I, Wu Z, and Takahashi Y. Molecular cloning and expression of the full-length tropomyosin gene from
Trichinella spiralis. J Helminthol 77: 57– 63, 2003.
843. Nakae H and Obinata T. Immunocytochemical localization of
troponin I and C in the muscles of Caenorhabditis elegans. Zool
Sci 10: 375–379, 1993.
844. Nakamura T, Yanagisawa T, and Yamaguchi M. Studies on the
subunits of myosin from muscle layer of Ascaris lumbridoides
suum. Biochim Biophys Acta 412: 229 –240, 1975.
845. Nakauchi Y and Maruyama K. Immunoblot detection of vertebrate type of connectin (titin) in ascidian bodywall muscle and
tadpole. Zool Sci 9: 219 –221, 1992.
846. Nanduri J and Kazura JW. Paramyosin enhanced clearance of
Brugia malayi microfilaremia in mice. J Immunol 143: 3359 –
3363, 1989.
847. Nara T, Matsumoto N, Janecharut T, Natsuda H, Yamamoto
K, Irimura T, Nakamura K, Aikawa M, Oswald I, Sher A, Kita
K, and Kojima S. Demonstration of the target molecule of a
protective IgE antibody in secretory glands of Schistosoma japonicum larvae. Int Immunol 6: 963–971, 1994.
848. Nara T, Tanabe K, and Mahakunkijcharoen Y. The B cell
epitope of paramyosin recognized by a protective monoclonal IgE
antibody to Schistosoma japonicum. Vaccine 15: 79 – 84, 1997.
849. Nascimento E, Leao IC, Pereira VRA, Gomes YM, Chikhlikar
P, August T, Marques E, and Lucena-Silva N. Protective immunity of single and multi-antigen DNA vaccines against schistosomiasis. Mem Inst Oswaldo Cruz 97: 105–109, 2002.
850. Nave R, Furst D, Vinkemeier U, and Weber K. Purification and
physical properties of nematode mini-titins and their relation to
twitchin. J Cell Sci 98: 491– 496, 1991.
851. Nave R and Weber K. A myofibrillar protein of insect muscle
related to vertebrate titin connects Z band and A band: purification and molecular characterization of invertebrate mini-titin.
J Cell Sci 95: 535–544, 1990.
852. Neil DM, Fowler WS, and Tobasnick G. Myofibrillar protein
composition correlates with histochemistry in fibres of the abdominal flexor muscles of the Norway lobster Nephros norvegicus. J Exp Biol 183: 185–201, 1993.
853. Neuman S, Kaban A, Volk T, Yaffe D, and Nudel U. The
dystrophin/utrophin homologues in Drosophila and in sea urchin.
Gene 263: 17–29, 2001.
854. Newman SM and Wright TRF. A histological and ultrastructural
analysis of developmental defects produced by the mutation,
lethal(1)myospheroid, in Drosophila melanogaster. Dev Biol 86:
393– 402, 1981.
855. Newport GR, Harrison RA, McKerrow J, Tarr P, Kallestad J,
and Agabian N. Molecular cloning of Schistosoma mansoni
myosin. Mol Biochem Parasitol 26: 29 –38, 1987.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
856. Niki T, Kato Y, Nozawa H, and Seki N. Gelation of low salt
soluble proteins from scallop adductor muscle in relation to its
freshness. Fisheries Sci 68: 688 – 693, 2002.
857. Nishikata T, Mita-Miyazawa I, Deno T, and Satoh N. Muscle
cell differentiation in ascidian embryos analysed with a tissuespecific monoclonal antibody. Development 99: 163–171, 1987.
858. Nishino A, Satou Y, Morisawa M, and Satoh N. Muscle actin
genes and muscle cells in the appendicularian, Oikopleura longicauda: phylogenetic relationships among muscle tissues in the
urochordates. J Exp Zool 288: 135–150, 2000.
859. Nishita K. Preparation and biochemical properties of actin from
striated adductor muscle of scallop. Nippon Suisan Gakk 43:
805– 812, 1977.
860. Nishita K and Ojima T. American lobster troponin. J Biochem
108: 677– 683, 1990.
861. Nishita K, Ojima T, Takahashi A, and Inoue A. Troponin from
smooth adductor muscle of Ezo giant scallop. J Biochem 121:
419 – 424, 1997.
862. Nishita K, Ojima T, and Watanabe S. Myosin from striated
adductor muscle of Chlamys nipponensis akazara. J Biochem
86: 663– 673, 1979.
863. Nishita K, Tanaka H, and Ojima T. Amino acid sequence of
troponin C from scallop striated adductor muscle. J Biol Chem
269: 3464 –3468, 1994.
864. Nitao LK, Loo RRO, O’Neall-Hennessey E, Loo JA, SzentGyörgyi AG, and Reisler E. Conformation and dynamics of the
SH1-SH2 helix in scallop myosin. Biochemistry 42: 7663–7674,
2003.
865. Noguchi S. Reito surimi no kagaku-I. Suisan Neriseihin Gijutu
Kenkyo Kaishi 5: 356 –363, 1976.
866. Nongthomba U, Clark S, Cummins M, Ansari M, Stark M, and
Sparrow JC. Troponin I is required for myofibrillogenesis and
sarcomere formation in Drosophila flight muscle. J Cell Sci 117:
1795–1805, 2004.
867. Nongthomba U, Cummins M, lark S, igoreaux JO, and Parrow JC. Suppression of muscle hypercontraction by mutations in
the myosin heavy chain gene of Drosophila melanogaster. Genetics 164: 209 –222, 2003.
868. Nongthomba U, Pasalodos-Sanchez S, Clark S, Clayton JD,
and Sparrow JC. Expression and function of the Drosophila
ACT88F actin isoform is not restricted to the indirect flight muscles. J Muscle Res Cell Motil 22: 111–119, 2001.
869. Nongthomba U and Ramachandra NB. A direct screen identifies
new flight muscle mutants on the Drosophila second chromosome. Genetics 153: 261–274, 1999.
870. Nose A, Mahajan VB, and Goodman CS. Connectin: a homophilic cell adhesion molecule expressed on a subset of muscles
and the motoneurons that innervate them in Drosophila. Cell 70:
553–567, 1992.
871. Nose A, Takeichi M, and Goodman CS. Ectopic expression of
connectin reveals a repulsive function during growth cone guidance and synapse formation. Neuron 13: 525–539, 1994.
872. Nose A, Umeda T, and Takeichi M. Neuromuscular target recognition by a homophilic interaction of connectin cell adhesion
molecules in Drosophila. Development 124: 1433–1441, 1997.
873. Nyitray L, Goodwin EB, and Szent-Györgyi AG. Complete
primary structure of a scallop striated muscle myosin heavy chain.
Sequence comparison with other heavy chains reveals regions
that might be critical for regulation. J Biol Chem 266: 18469 –
18476, 1991.
874. Nyitray L, Jancso A, Ochiai Y, Graf L, and Szent-Györgyi AG.
Scallop striated and smooth muscle myosin heavy chain isoforms
are produced by alternative RNA splicing from a single gene. Proc
Natl Acad Sci USA 91: 12686 –12690, 1994.
875. Obinata T, Ooi A, and Takano-Ohmuro H. Myosin and actin
from ascidian smooth muscle and their interaction. Comp Biochem Physiol B Biochem 76: 437– 442, 1983.
876. Obinata T, Shirao T, and Murakami S. Sea urchin paramyosin.
Int J Biochem 6: 114 –126, 1975.
877. Obwaller A, Duchene M, Bruhn H, Steipe B, Tripp C, Kraft D,
Wiedermann G, Auer H, and Aspock H. Recombinant dissection of myosin heavy chain of Toxocara canis shows strong
clustering of antigenic regions. Parasitol Res 87: 383–389, 2001.
Physiol Rev • VOL
1049
878. Ochiai Y, Kariya Y, Watabe S, and Hashimoto K. Heat-induced
tendering of turban shell (Batillus cornutus) muscle. J Food Sci
50: 981–984, 1985.
879. O’Donnell PT and Bernstein SI. Molecular and ultrastructural
defects in a Drosophila myosin heavy chain mutant: differential
effects on muscle function produced by similar thick filament
abnormalities. J Cell Biol 107: 2601–2612, 1988.
880. O’Donnell PT, Collier VL, Mogami K, and Bernstein SI. Ultrastructural and molecular analyses of homozygous viable Drosophila melanogaster muscle mutants indicate there is a complex
pattern of myosin heavy chain isoform distribution. Genes Dev 3:
1233–1246, 1989.
881. Ogg SC, Anderson P, and Wickens MP. Splicing of a C. elegans
myosin pre-mRNA in a human nuclear extract. Nucleic Acids Res
18: 143–149, 1990.
882. Ohshima S, Komiya T, Takeuchi K, Endo T, and Obinata T.
Generation of multiple troponin T isoforms is a common feature
of the muscles in various chordate animals. Comp Biochem
Physiol B Biochem 90: 779 –784, 1988.
883. Ohshima S, Vallarimo C, and Gailey DA. Reassessment of 79B
actin gene expression in the abdomen of adult Drosophila melaogaster. Insect Mol Biol 6: 227–231, 1997.
884. Ohtani Y, Maki S, Kimura S, and Maruyama K. Localization of
connectin-like proteins in leg and flight muscles of insects. Tissue
Cell 28: 1– 8, 1996.
885. Ojima T. Studies on structure and function of molluscan troponin. Nippon Suisan Gakk 69: 326 –329, 2003.
886. Ojima T. Analysis of primary structure of proteins by cDNA
cloning. Nippon Suisan Gakk 60: 543–544, 1994.
887. Ojima T, Maita M, Inoue A, and Nishita K. Bacterial expression, purification, and characterization of akazara scallop troponin C. Fisheries Sci 63: 137–141, 1997.
888. Ojima T and Nishita K. Isolation of troponins from striated and
smooth adductor muscles of Akazara scallop. J Biochem 100:
821– 824, 1986.
889. Ojima T and Nishita K. Troponin from Akazara scallop striated
adductor muscles. J Biol Chem 261: 16749 –16754, 1986.
890. Ojima T and Nishita K. Separation of Akazara scallop and rabbit
troponin components by a single step chromatography on CMtoyopearl. J Biochem 104: 9 –11, 1988.
891. Ojima T and Nishita K. Comparative studies on biochemical
characteristics of troponins from Ezo giant scallop (Patinopecten
yessoensis) and Akazara scallop (Chlamys nipponensis akazara). Comp Biochem Physiol B Biochem 103: 727–732, 1992.
892. Ojima T, Ohta T, and Nishita K. Amino acid sequence of squid
troponin C. Comp Biochem Physiol B Biochem 129: 787–796,
2001.
893. Ojima T, Tanaka H, and Nishita K. Amino acid sequence of C
terminal 17-kDa CNBR fragment of Akazara scallop troponin I.
J Biochem 117: 158 –162, 1995.
894. Ojima T, Tanaka H, and Nishita K. Cloning and sequence of a
cDNA encoding Akazara scallop troponin C. Arch Biochem Biophys 311: 272–276, 1994.
895. Ojima T, Toyoguchi T, and Nishita K. Isolation and characterization of troponin from abdominal muscle of prawn Penaeus
japonicus. Fisheries Sci 61: 871– 875, 1995.
896. Ojima T, Yokomoto K, and Nishita K. Trinitrophenylation of
Akazara scallop striated adductor myosin. Nippon Suisan Gakk
55: 567–574, 1989.
897. Okamoto H, Hiromi Y, Ishikawa E, Yamada T, Isoda K,
Maekawa H, and Hotta Y. Molecular characterization of mutant
actin genes which induce heat-shock proteins in Drosophila flight
muscles. EMBO J 5: 589 –596, 1986.
898. Okkema PG, Harrison SW, Plunger V, Aryana A, and Fire A.
Sequence requirements for myosin gene expression and regulation in Caenorhabditis elegans. Genetics 135: 385– 404, 1993.
899. Olaechea RP, Ushio H, Watabe S, Takada K, and Hatae K.
Toughness and collagen content of abalone muscles. Biosci Biotechnol Biochem 57: 6 –11, 1993.
900. Olivas RR, Sandez OR, Haard NF, Aguilar RP, and Brauer
JME. Changes in firmness and thermal behavior of ice-stored
muscle of jumbo squid (Dosidicus gigas). Eur Food Res Technol
219: 312–315, 2004.
85 • JULY 2005 •
www.prv.org
1050
SCOTT L. HOOPER AND JEFFREY B. THUMA
901. Oliveira GC and Kemp WM. Cloning of two actin genes from
Schistosoma mansoni. Mol Biochem Parasitol 75: 119 –122, 1995.
902. Ono S. Purification and biochemical characterization of actin
from Caenorhabditis elegans: its difference from rabbit muscle
actin in the interaction with nematode ADF cofilin. Cell Motil
Cytoskeleton 43: 128 –136, 1999.
903. Oota S and Saitou N. Phylogenetic relationship of muscle tissues deduced from superimposition of gene trees. Mol Biol Evol
16: 856 – 867, 1999.
904. Orii H, Ito H, and Watanabe K. Anatomy of the planarian
Dugesia japonica. I. The muscular systems revealed by antisera
against myosin heavy chains. Zool Sci 19: 1123–1131, 2002.
905. Ortega MA, Macias MT, Martinez JL, Palmero I, and Sastre
L. Expression of actin isoforms in Artemia. Symp Soc Exp Biol
46: 131–137, 1992.
906. Oshino T, Shimamura J, Fukuzawa A, Maruyama K, and
Kimura S. The entire cDNA sequences of projectin isoforms of
crayfish claw closer and flexor muscles and their localization.
J Muscle Res Cell Motil 24: 431– 438, 2003.
907. Otwell WS and Hamann DD. Textural characterization of squid
Loligo pealei L: instrumental and panel evaluations. J Food Sci 44:
1636 –1642, 1979.
908. Otwell WS and Hamann DD. Textural characterization of squid:
scanning electron microscopy of cooked mantle. J Food Sci 44:
1629 –1635, 1979.
909. Overbeek PA, Merlino GT, Peters NK, Cohn VH, Moore GP,
and Kleinsmith LJ. Characterization of five members of the actin
gene family in the sea urchin. Biochim Biophys Acta 656: 195–205,
1981.
910. Papdopoulos LS, Smith SB, Wheeler TL, and Finne G. Muscle
ultrastructural changes in freshwater prawns, Macriobrachium
rosenbergii, during iced storage. J Food Sci 54: 1125–1128, 1989.
911. Paredi ME and Crupkin M. Physicochemical and biochemical
properties of actomyosin from striated adductor muscles of scallop (Zygochlamys patagonica) stored at 2– 4°C. J Aquatic Food
Prod Technol 11: 79 – 87, 2002.
912. Paredi ME and Crupkin M. Biochemical properties of actomyosin and expressible moisture of frozen stored adductor muscles
of patagonian scallop (Zygochlamys patagonica). J Food Biochem 27: 461– 470, 2003.
913. Paredi ME and Crupkin M. Biochemical properties of actomyosin from frozen stored mantles of squid (Illex argentinus) at
different sexual maturation stages. J Agric Food Chem 45: 1629 –
1632, 1997.
914. Paredi ME, Davidovich LA, and Crupkin M. Thermally induced
gelation of squid (Illex argentinus) actomyosin. Influence of sexual maturation stage. J Agric Food Chem 47: 3592–3595, 1999.
915. Paredi ME, De Vido de Mattio N, and Crupkin M. Biochemical
properties of actomyosin and expressible juice of cold stored
striated adductor muscles of Aulacomya ater ater (Molina): effects of ionic solutions. J Aquatic Food Prod Technol 1: 133–144,
1992.
916. Paredi ME, De Vido de Mattio N, and Crupkin M. Biochemical
properties of actomyosin of cold stored striated adductor muscles
of Aulaomya ater ater (Molina). J Food Sci 55: 1567–1570, 1990.
917. Paredi ME, deMattio NAD, and Crupkin M. Biochemical properties of actomyosin and expressible moisture of frozen stored
striated adductor muscles of Aulacomya ater ater (Molina): effects of polyphosphates. J Agric Food Chem 44: 3108 –3112, 1996.
918. Paredi ME, Tomas MC, Añon MC, and Crupkin M. Thermal
stability of myofibrillar proteins from smooth and striated muscles of scallop (Chlamys tehuelchus): a differential scanning calorimetric study. J Agric Food Chem 46: 3971–3976, 1998.
919. Paredi ME, Tomas MC, and Crupkin M. Thermal behavior of
myofibrillar proteins from the adductor muscles of scallops. A
differential scanning calorimetric study (DSC). Braz J Chem Eng
20: 153–159, 2003.
920. Paredi ME, Tomas MC, and Crupkin M. Thermal denaturation
of myofibrillar proteins of striated and smooth adductor muscles
of scallop (Zygochlamys patagonica). A differential scanning calorimetric study. J Agric Food Chem 50: 830 – 834, 2002.
921. Paredi ME, Tomas MC, Crupkin M, and Añon MC. Postmortem changes in adductor muscles from Aulacomya ater ater
Physiol Rev • VOL
922.
923.
924.
925.
926.
927.
928.
929.
930.
931.
932.
933.
934.
935.
936.
937.
938.
939.
940.
941.
942.
(Molina) stored at 2– 4°C. A differential scanning calorimetic
study. J Agric Food Chem 43: 1758 –1761, 1995.
Paredi ME, Tomas MC, Crupkin M, and Añon MC. Thermal
denaturation of Aulacomya ater ater (Molina) myofibrillar proteins—a differential scanning calorimetric study. J Agric Food
Chem 42: 873– 877, 1994.
Paredi ME, Tomas MC, Crupkin M, and Añon MC. Thermal
denaturation of muscle proteins from male and female squid (Illex
argentinus) at different sexual maturation stages. A differential
scanning calorimetric study. J Agric Food Chem 44: 3812–3816,
1996.
Parker VP, Falkenthal S, and Davidson N. Characterization of
the myosin light-chain-2 gene of Drosophila melanogaster. Mol
Cell Biol 5: 3058 –3068, 1985.
Parker-Thornburg J and Bonner JJ. Mutations that induce the
heat shock response of Drosophila. Cell 51: 763–772, 1987.
Parkinson J, Whitton C, Guiliano D, Daub J, and Blaxter M.
200,000 nematode expressed sequence tags on the Net. Trends
Parasitol 17: 394 –396, 2001.
Parkinson J, Mitreva M, Hall N, Blaxter M, and McCarter JP.
400,000 nematode ESTs on the Net. Trends Parasitol 19: 283–286,
2003.
Pascolini R, Panara F, Di Rosa I, Fagotti A, and Lorvik S.
Characterization and fine-structural localization of actin-and fibronectin-like proteins in planaria (Dugesia lugubris s. l). Cell
Tissue Res 267: 499 –506, 1992.
Pascolini R, Panara F, Fagotti A, DiRosa I, Simoncelli F, and
Gabbiani G. Cytoskeletal protein expression in planarians. Boll
Zool 60: 403– 406, 1993.
Patwary MU, Reith M, and Kenchington EL. Isolation and
characterization of a cDNA encoding an actin gene from sea
scallop (Placopecten magellanicus). J Shellfish Res 15: 265–270,
1996.
Patwary MU, Reith M, and Kenchington EL. Cloning and characterization of tropomyosin cDNAs from the sea scallop Placopecten magellanicus (Gmelin, 1791). J Shellfish Res 18: 67–70,
1999.
Pauls TL, Cox JA, Heizmann CW, and Hermann A. Sarcoplasmic calcium binding proteins in Aplysia nerve and muscle cells.
Eur J Neurosci 5: 549 –559, 1993.
Peachey LD. Muscle. Annu Rev Physiol 30: 401– 440, 1968.
Pearce EJ. Progress towards a vaccine for schistosomiasis. Acta
Tropica 86: 309 –313, 2003.
Pearce EJ, James SL, Dalton J, Barrall A, Ramos C, Strand
M, and Sher A. Immunochemical characterization and purification of Sm-97, a Schistosoma mansoni antigen monospecifically
recognized by antibodies from mice protectively immunized with
a nonliving vaccine. J Immunol 137: 3593–3600, 1986.
Pearce EJ, James SL, Hieny S, Lanar DE, and Sheer A.
Induction of protective immunity against Schistosoma mansoni
by vaccination with schistosome paramyosin (SM97), a nonsurface parasite antigen. Proc Natl Acad Sci USA 85: 5678 –5682,
1988.
Peckham M, Cripps RM, White DC, and Bullard B. Mechanics
and protein content of insect flight muscles. J Exp Biol 168: 57–76,
1992.
Pei-Ken S and Tien-Chin T. A comparative study of nucleotropomyosins from different sources. Scientia Sinica 4: 157–175,
1954.
Pérez-Pérez J, Fernández-Caldas E, Marañón F, Sastre J,
Bernal ML, Rodrı́guez J, and Bedate CA. Molecular cloning of
paramyosin, a new allergen of Anisakis simplex. Int Arch Allergy
Immunol 123: 120 –129, 2000.
Perreault-Micale CL, Jancso A, and Szent-Györgyi AG. Essential and regulatory light chains of Placopecten striated and
catch muscle myosins. J Muscle Res Cell Motil 17: 533–542, 1996.
Perreault-Micale CL, Kalabokis VN, Nyitray L, and SzentGyörgyi AG. Sequence variations in the surface loop near the
nucleotide binding site modulate the ATP turnover rates of molluscan myosins. J Muscle Res Cell Motil 17: 543–553, 1996.
Petrova TV, Comte M, Takagi T, and Cox JA. Thermodynamic
and molecular properties of the interaction between Amphioxus
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
943.
944.
945.
946.
947.
948.
949.
950.
951.
952.
953.
954.
955.
956.
957.
958.
959.
960.
961.
calcium vector protein and its 26 kDa target. Biochemistry 34:
312–318, 1995.
Petrova TV, Takagi T, and Cox JA. Phosphorylation of the IQ
domain regulates the interaction between Ca2⫹-vector protein and
its target in amphioxus. J Biol Chem 271: 26646 –26652, 1996.
Philpott DE, Kahlbrock M, and Szent-Györgyi AG. Filamentous organization of molluscan muscles. J Ultrastruct Res 3:
254 –269, 1960.
Pietrantonio PV, Holmes SP, Jagge C, and Frazier SK. Cloning of troponin C and other gene fragments from the red imported
fire ant Solenopsis invicta Buren (Hymenoptera, Formicidae).
Southwest Entolmol Suppl 25: 89 –96, 2002.
Platt J. Fibres connecting the central nervous system and the
chorda in amphioxus. Anat Anz 7: 282–291, 1892.
Plenefisch JD, Zhu XP, and Hedgecock EM. Fragile skeletal
muscle attachments in dystrophic mutants of Caenorhabditis
elegans: isolation and characterization of the mua genes. Development 127: 1197–1207, 2000.
Podgornaya OI, Drozdov AL, Margulis BA, and Pinaev GP.
Sarcoplasmic actin in muscle cells of ambulacral tubes of the sea
star. Biofizika 27: 285–287, 1982.
Podlubnaya ZA, Shpagina MD, Vikhlyantsev IM, Malyshev
SL, Udaltsov SN, Ziegler C, and Beinbrech G. Comparative
electron microscopic study on projectin and titin binding to F
actin. Insect Biochem Mol Biol 33: 789 –793, 2003.
Polyak E, Standiford DM, Yakopson V, Emerson CP, and
Franzini-Armstrong C. Contribution of myosin rod protein to
the structural organization of adult and embryonic muscles in
Drosophila. J Mol Biol 331: 1077–1091, 2003.
Popeijus M, Blok VC, Cardle L, Bakker E, Phillips MS,
Helder J, Smant G, and Jones JT. Analysis of genes expressed
in second stage juveniles of the potato cyst nematodes Globodera
rostochiensis and G. pallida using the expressed sequence tag
approach. Nematology 2: 567–574, 2000.
Prado A, Canal I, Barbas JA, Molloy J, and Ferrus A. Functional recovery of troponin I in a Drosophila heldup mutant after
a second site mutation. Mol Biol Cell 6: 1433–1441, 1995.
Prêcheur B, Cox JA, Petrova T, Mispelter J, and Craescu
CT. Nereis sarcoplasmic Ca2⫹-binding protein has a highly unstructured apo state which is switched to the native state upon
binding of the first Ca2⫹ ion. FEBS Lett 395: 89 –94, 1996.
Probst WC, Cropper EC, Heierhorst J, Hooper SL, Jaffe H,
Vilim F, Beushausen S, Kupfermann I, and Weiss KR. cAMPdependent phosphorylation of Aplysia twitchin may mediate
modulation of muscle contractions by neuropeptide cotransmitters. Proc Natl Acad Sci USA 91: 8487– 8491, 1994.
Prokop A, Martin-Bermudo MD, Bate M, and Brown NH.
Absence of PS integrins or laminin A affects extracellular adhesion, but not intracellular assembly, of hemiadherens and neuromuscular junctions in Drosophila embryos. Dev Biol 196: 58 –76,
1998.
Pulak RA and Anderson P. Structures of spontaneous deletions
in Caenorhabditis elegans. Mol Cell Biol 8: 3748 –3754, 1988.
Pureur RP, Coffe G, Soyer-Gobillard MO, de Billy F, and
Pudles J. A network of 2– 4 nm filaments found in sea urchin
smooth muscle. Exp Cell Res 162: 63–76, 1986.
Pyeun JH, Hashimoto K, and Matsuura F. Isolation and characterization of abalone paramyosin. Nippon Suisan Gakk 39:
395– 402, 1973.
Qiu F, Lakey A, Agianian B, Hutchings A, Butcher GW, Labeit S, Leonard K, and Bullard B. Troponin C in different insect
muscle types: identification of two isoforms in Lethocerus, Drosophila, and Anopheles that are specific to asynchronous flight
muscle in the adult insect. Biochem J 371: 811– 821, 2003.
Quigley MM and Mellon D Jr. Changes in myofibrillar gene
expression during fiber type transformation in the claw closer
muscles of the snapping shrimp, Alpheus heterochelis. Dev Biol
106: 262–265, 1984.
Raghavan J, Maina CV, Fitzgerald PC, Tuan RS, Slatko BE,
Ottesen EA, and Nutman TB. Characterization of a muscleassociated antigen from Wuchereria bancrofti. Exp Parasitol 75:
379 –389, 1992.
Physiol Rev • VOL
1051
962. Raghavan N, McReynolds LA, Maina CV, Feinstone SM, Jayaraman K, Ottesen EA, and Nutman TB. A recombinant clone
of Wuchereria bancrofti with DNA specificity for human lymphatic filarial parasites. Mol Biochem Parasitol 47: 63–71, 1991.
963. Raghavan S and White RAH. Connectin mediates adhesion in
Drosophila. Neuron 18: 873– 880, 1997.
964. Ram D, Grossman Z, Markovics A, Avivi A, Ziv E, Lantner F,
and Schechter I. Rapid changes in the expression of a gene
encoding a calcium-binding protein in Schistosoma mansoni. Mol
Biochem Parasitol 34: 167–176, 1989.
965. Ramı́rez BL, Kuris JD, Wiest PM, Arias P, Aligui F, Peters P,
and Olds GR. Paramyosin: a candidate vaccine antigen against
Schistosoma japonicum. Parasite Immunol 18: 49 –52, 1996.
966. Ramos JDA, Cheong N, Lee BW, and Chua KY. Peptide mapping of immunoglobulin E and immunoglobulin G immunodominant epitopes of an allergenic Blomia tropicalis paramyosin, Blo
t 11. Clin Exp Allergy 33: 511–517, 2003.
967. Ramos JDA, Nge C, Wah LB, and Yan CK. cDNA cloning and
expression of Blo t 11, the Blomia tropicalis allergen homologous
to paramyosin. Int Arch Allergy Immunol 126: 286 –293, 2001.
968. Ramos JDA, Teo ASM, Lee BW, Cheong N, and Chua KY. DNA
immunization for the production of monoclonal antibodies to Blo
t 11, a paramyosin homolog from Blomia tropicalis. Allergy 59:
539 –547, 2004.
969. Ramos JDA, Teo ASM, Ou KL, Tsai LC, Lee BW, Cheong N,
and Chua KY. Comparative allergenicity studies of native and
recombinant Blomia tropicalis paramyosin (Blo t 11). Allergy 58:
412– 419, 2003.
970. Rao PVS, Rajagopal D, and Ganesh KA. B- and T-cell epitopes
of tropomyosin, the major shrimp allergen. Allergy 53: 44 – 47,
1998.
971. Ravaux J, Hassanin A, Deutsch J, Gaill F, and MarkmannMulisch U. Sequence analysis of the myosin regulatory light chain
gene of the vestimentiferan Riftia pachyptila. Gene 263: 141–149,
2001.
972. Rayment I and Holden HM. The 3-dimensional structure of a
molecular motor. Trends Biochem Sci 19: 129 –134, 1994.
973. Razzaq A, Robinson IM, McMahon HT, Skepper JN, Su Y,
Zelhof AC, Jackson AP, Gay NJ, and O’Kane CJ. Amphiphysin
is necessary for organization of the excitation-contraction coupling machinery of muscles, but not for synaptic vesicle endocytosis in Drosophila. Genes Dev 15: 2967–2979, 2001.
974. Razzaq A, Schmitz S, Veigel C, Molloy JE, Geeves MA, and
Sparrow JC. Actin residue Glu93 is identified as an amino acid
affecting myosin binding. J Biol Chem 274: 28321–28328, 1999.
975. Reber-Müller S, Studer R, Muller P, Yanze N, and Schmid V.
Integrin and talin in the jellyfish Podocoryne carnea. Cell Biol Int
25: 753–769, 2001.
976. Reedy MC, Bullard B, and Vigoreaux JO. Flightin is essential
for thick filament assembly and sarcomere stability in Drosophila
flight muscles. J Cell Biol 151: 1483–1499, 2000.
977. Reedy MC, Reedy MK, Leonard KR, and Bullard B. Gold/Fab
immunoelectron microscopy localization of troponin H and troponin T in Lethocerus flight muscle. J Mol Biol 239: 52– 67, 1994.
978. Reese G, Ayuso R, Carle T, and Lehrer SB. IgE-binding
epitopes of shrimp tropomyosin, the major allergen Pen a 1. Int
Arch Allergy Immunol 118: 300 –301, 1999.
979. Reese G, Ayuso R, and Lehrer SB. Tropomyosin: an invertebrate pan-allergen? Int Arch Allergy Immunol 119: 247–258, 1999.
980. Reese G, Ayuso R, Leong-Kee SM, Plante M, and Lehrer SB.
The IgE binding regions of the major allergen Pen a 1: multiple
epitopes or intramolecular cross-reactivity? Int Arch Allergy Immunol 124: 103–106, 2001.
981. Reese G, Daul CB, and Lehrer SB. Antigenic analysis (IgE and
monoclonal antibodies) of the major shrimp allergen Pen a 1
(tropomyosin) from Penaeus aztecus. Int Arch Allergy Immunol
107: 245–247, 1995.
982. Reese G, Jeoung BJ, Daul CB, and Lehrer SB. Characterization of recombinant shrimp allergen Pen a 1 (tropomyosin). Int
Arch Allergy Immunol 113: 240 –242, 1997.
983. Reese G, Tracey D, Daul CB, and Lehrer SB. IgE and monoclonal antibody reactivities to the major shrimp allergen Pen i 1
85 • JULY 2005 •
www.prv.org
1052
984.
985.
986.
987.
988.
989.
990.
991.
992.
993.
994.
995.
996.
997.
998.
999.
1000.
1001.
1002.
SCOTT L. HOOPER AND JEFFREY B. THUMA
(tropomosin) and vertebrate tropomyosin. Adv Exp Med Biol 409:
225–230, 1996.
Regenstein JM. Lobster (Homarus americanus) striated muscle
myosin. Comp Biochem Physiol B Biochem 56: 239 –244, 1977.
Regenstein JM and Szent-Györgyi AG. Regulatory proteins of
lobster striated muscle. Biochemistry 14: 917–925, 1975.
Reifegerste R, Grimm S, Albert S, Lipski N, Heimbeck G,
Hofbauer A, Pfugfelder GO, Quack D, Reichmuth C, Schug B,
Zinsmaier KE, Buchner S, and Buchner E. An invertebrate
calcium-binding protein of the calbindin subfamily: protein structure, genomic organization, and expression pattern of the calbindin-32 gene of Drosophila. J Neurosci 13: 2186 –2198, 1993.
Reiner DJ, Weinshenker D, and Thomas JH. Analysis of dominant mutations affecting muscle excitation in Caenorhabditis
elegans. Genetics 141: 961–976, 1995.
Richter D and Harn DA. Candidate vaccine antigens identified
by antibodies from mice vaccinated with 15- or 50-kilorad-radiated cercariae of Schistosoma mansoni. Infect Immun 61: 146 –
154, 1993.
Richter D, Incani RN, and Harn DA. Isotype responses to
candidate vaccine antigens in protective sera obtained from mice
vaccinated with irradiated cercariae of Schistosoma mansoni.
Infect Immun 61: 3003–3011, 1993.
Richter D, Reynolds SR, and Harn DA. Candidate vaccine
antigens that stimulate the cellular immune response of mice
vaccinated with irradiated cercariae of Schistosoma mansoni.
J Immunol 151: 256 –265, 1993.
Riddiford LM. Solvent perturbation and ultraviolet optical rotatory dispersion studies of paramyosin. J Biol Chem 241: 2792–
2802, 1966.
Riddiford LM and Scheraga HA. Structural studies of paramyosin. I. Hydrogen ion equilibria. Biochemistry 1: 95–107, 1962.
Riddle DL and Brenner S. Indirect suppression in Caenorhabditis elegans. Genetics 89: 299 –314, 1978.
Riemer D, Dodemont H, and Weber K. Analysis of the cDNA
and gene encoding a cytoplasmic intermdiate filament (IF) protein
from the cephalochordate Branchiostoma lanceolatum: implications for the evolution of the IF protein family. Eur J Cell Biol 58:
128 –135, 1992.
Riemer D, Dodemont H, and Weber K. Cloning of the nonneuronal intermediate filament protein of the gastropod Aplysia
californica; identification of an amio acid residue essential for the
IFA epitope. Eur J Cell Biol 56: 351–357, 1991.
Riemer D, Karabinos A, and Weber K. Analysis of eight cDNAs
and six genes for intermediate filament proteins in the cephalochordate Branchiostoma reveals differences in the multigene
families of lower chordates and the vertebrates. Gene 211: 361–
373, 1998.
Riemer D and Weber K. Common and variant properties of
intermediate filament proteins from lower chordates and vertebrates: two proteins from the tunicate Styela and the identification of a type III homologue. J Cell Sci 111: 2967–2975, 1998.
Riparbelli MG, Callaini G, and Dallai R. A segment corresponding to amino acids Gln199-Lys208 of murine IL-1 alpha
cross-reacts with an antigenic determinant localized in the Z line
of Drosophila melanogaster myofibrils. Biol Cell 86: 139 –144,
1996.
Roberts RG and Bobrow M. Dystrophins in vertebrates and
invertebrates. Hum Mol Genet 7: 589 –595, 1998.
Rodrı́guez-Ortega MJ, Grøsvik BE, Rodrı́guez-Ariza AR,
Goksøyr A, and López-Barea J. Changes in protein expression
profiles in bivalve molluscs (Chamaelea gallina) exposed to four
model environmental pollutants. Proteomics 3: 1535–1543, 2003.
Rogalski TM, Gilbert MM, Devenport D, Norman KR, and
Moerman DG. DIM-1, a novel imnmunoglobulin superfamily protein in Caenorhabditis elegans, is necessary for maintaining bodywall muscle integrity. Genetics 163: 905–915, 2003.
Rogalski TM, Gilchrist EJ, Mullen GP, and Moerman DG.
Mutations in the unc-52 gene responsible for body wall muscle
defects in adult Caenorhabditis elegans are located in alternatively spliced exons. Genetics 139: 159 –169, 1995.
Physiol Rev • VOL
1003. Rogalski TM, Moerman DG, and Baillie DL. Essential genes
and deficiencies in the unc-22 region of Caenorhabditis elegans.
Genetics 102: 725–736, 1982.
1004. Rogalski TM, Mullen GP, Bush JA, Gilchrist EJ, and Moerman DG. unc-52/perlecan isoform diversity and function in Caenorhabditis elegans. Biochem Soc Trans 29: 171–176, 2001.
1005. Rogalski TM, Mullen GP, Gilbert MM, Williams BD, and
Moerman DG. The unc-112 gene in Caenorhabditis elegans
encodes a novel component of cell-matrix adhesion structures
required for integrin localization in the muscle cell membrane.
J Cell Biol 150: 253–264, 2000.
1006. Rogalski TM, Williams BD, Mullen GP, and Moerman DG.
Products of the unc-52 gene in Caenorhabditis elegans are homologous to the core protein of the mammalian basement membrane heparan sulfate proteoglycan. Genes Dev 7: 1471–1484,
1993.
1007. Rosenberg-Hasson Y, Renert-Pasca M, and Volk T. A Drosophila dystrophin-related protein, MSP-300, is required for embryonic muscle morphogenesis. Mech Dev 60: 83–94, 1996.
1008. Roulier EM, Fyrberg C, and Fyrberg E. Perturbations of Drosophila alpha-actinin cause muscle paralysis, weakness, and atrophy but do not confer obvious nonmuscle phenotypes. J Cell
Biol 116: 911–22, 1992.
1009. Rowlerson AM and Blackshaw SE. Fiber types in leech body
wall muscle. J Exp Biol 157: 299 –311, 1991.
1010. Royuela M, Astier C, Fraile B, and Paniagua R. Alpha-actinin
in different invertebrate muscle cell types of Drosophila melanogaster, the earthworm Eisenia foetida, and the snail Helix aspersa. J Muscle Res Cell Motil 20: 1–9, 1999.
1011. Royuela M, Astier C, Grandier-Vazeille X, Benyamin Y,
Fraile B, Paniagua R, and Duvert M. Immunohistochemistry of
chaetognath body wall muscles. Invert Biol 122: 74 – 82, 2003.
1012. Royuela M, Fraile B, Arenas MI, and Paniagua R. Characterization of several invertebrate muscle cell types: a comparison
with vertebrate muscles. Microsc Res Tech 48: 107–115, 2000.
1013. Royuela M, Fraile B, Cervera M, and Paniagua R. Immunocytochemical electron microscopic study and Western blot analysis
of myosin, paramyosin, and miniparamyosin in the striated muscle of the fruit fly Drosophila melangaster and in obliquely striated and smooth muscles of the earthworm Eisenia foetida.
J Muscle Res Cell Motil 18: 169 –177, 1997.
1014. Royuela M, Fraile B, De Miguel MP, Cervera M, and Paniagua R. Immunohistochemical study and Western blotting analysis of titin-like proteins in the striated muscle of Drosophila
melanogaster and in the striated and smooth muscle of the oligochaete Eisenia foetida. Microsc Res Tech 35: 349 –356, 1996.
1015. Royuela M, Fraile B, and Paniagua R. Nebulin-like protein in
the earthworm Eisenia foetida. Immunocytochemical electron
microscopic study and Western blot analysis of several muscle
cell types. Eur J Cell Biol 73: 276 –280, 1997.
1016. Royuela M, Fraile B, Paniagua R, and Meyer-Rochow VB.
Immunocytochemical observations on muscle cell proteins in the
antarctic mussel shrimp Acetabulastoma sp (Crustacea, Ostracoda). Invert Biol 118: 184 –189, 1999.
1017. Royuela M, Fraile B, Picazo ML, and Paniagua R. 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. Eur J Cell Biol 72: 90 –94, 1997.
1018. Royuela M, Garcia-Anchuelo R, Arenas MI, Cervera M,
Fraile B, and Paniagua R. Immunocytochemical electron microscopic study and Western blot analysis of paramyosin in different
invertebrate muscle cell types of the fruit fly Drosophila melanogaster, the earthworm Eisenia foetida, and the snail Helix aspersa. Histochem J 28: 247–255, 1996.
1019. Royuela M, Garcia-Anchuelo R, Paz de Miguel M, Arenas MI,
Fraile B, and Paniagua R. Immunocytochemical electron microscopic study and Western blot analysis of troponin in striated
muscle of the fruit fly Drosophila melanogaster and in several
muscle cell types of the earthworm Eisenia foetida. Anat Rec 244:
148 –154, 1996.
1020. Royuela M, Hugon G, Rivier F, Paniagua R, and Mornet D.
Dystrophin associated proteins in obliquely striated muscle of the
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
1021.
1022.
1023.
1024.
1025.
1026.
1027.
1028.
1029.
1030.
1031.
1032.
1033.
1034.
1035.
1036.
1037.
1038.
1039.
leech Pontobdella muricata (Annelida, Hirudinea). Histochem J
33: 135–139, 2001.
Royuela M, Meyer-Rochow VB, Fraile B, and Paniagua R.
Muscle cells in the tiny marine Antarctic mite Halacarellus thomasi: an ultrastructural and immunocytochemical study. Polar
Biol 23: 759 –765, 2000.
Royuela M, Paniagua R, Rivier F, Hugon G, Robert A, and
Mornet D. Presence of invertebrate dystrophin-like products in
obliquely striated muscle of the leech, Pontobdella muricata (Annelida, Hirudinea). Histochem J 31: 603– 608, 1999.
Rozek CE and Davidson N. Drosophila has one myosin heavychain gene with three developmentally regulated transcripts. Cell
32: 23–34, 1983.
Rozek CE and Davidson N. Differential processing of RNA
transcribed from the single copy Drosophila myosin heavy chain
gene produces four messenger RNAs that encode two polypeptides. Proc Natl Acad Sci USA 83: 2128 –2132, 1986.
Rubenstein PA. The functional importance of multiple actin
isoforms. Bioessays 12: 309 –315, 1990.
Rubin GM, Yandell MD, Wortman JR, Gabor Miklos GL,
Nelson CR, Hariharan IK, Fortini ME, Li PW, Apweiler R,
Fleischmann W, Cherry JM, Henikoff S, Skupski MP, Misra
S, Ashburner Mi Birney E, Boguski MS, Brody T, Brokstein
P, Celniker SE, Chervitz SA, Coates D, Cravchik A, Gabrielian A, Galle RF, Gelbart WM, George RA, Goldstein LS,
Gong F, Guan P, Harris NL, Hay BA, Hoskins RA, Li J, Li Z,
Hynes RO, Jones SJM, Kuehl PM, Lemaitre B, Littleton JT,
Morrison DK, Mungall C, O’Farrell PH, Pickeral OK, Shue C,
Vosshall LB, Zhang J, Zhao Q, Zheng XH, Zhong F, Zhong W,
Gibbs R, Venter JC, Adams MD, and Lewis S. Comparative
genomics of the eukaryotes. Science 287: 2204 –2215, 2000.
Rüegg JC. The proteins associated with contraction in lamellibranch “catch” muscle. Proc R Soc Lond B Biol Sci 154: 209 –223,
1961.
Ruiz-Trillo I, Paps J, Loukota M, Ribera C, Jondelius U,
Baguñà J, and Riutort M. A phylogenetic analysis of myosin
heavy chain type II sequences corroborates that Acoela and Nemertodermatida are basal bilaterians. Proc Natl Acad Sci USA 99:
11246 –11251, 2002.
Rushforth AM, Saari B, and Anderson P. Site-selected insertion of the transposon Tc1 into a C. elegans myosin light chain
gene. Mol Cell Biol 13: 902–910, 1993.
Rushforth AM, White CC, and Anderson P. Functions of the
Caenorhabditis elegans regulatory myosin light chain genes mlc-1
and mlc-2. Genetics 150: 1067–1077, 1998.
Rybakova IN, Patel JR, and Ervasti JM. The dystrophin complex forms a mechanically strong link between the sarcolemma
and costameric actin. J Cell Biol 150: 1209 –1214, 2000.
Saarne T, Kaiser L, Rasool O, Huecas S, Hage-Hamsten M,
and Gafvelin G. Cloning and characterisation of two IgE-binding
proteins, homologous to tropomyosin and ␣-tubulin, from the
mite Lepidoglyphus destructor. Int Arch Allergy Immunol 130:
258 –265, 2003.
Saide JD. Identification of a connecting filament protein in insect
fibrillar flight muscle. J Mol Biol 153: 661– 679, 1981.
Saide JD, Chin-Bow S, Hogan-Sheldon J, and BusquetsTurner L. Z band proteins in the flight muscle and leg muscle of
the honeybee. J Muscle Res Cell Motil 11: 125–136, 1990.
Saide JD, Chin-Bow S, Hogan-Sheldon J, Busquets-Turner L,
Vigoreaux JO, Valgeirsdottir K, and Pardue ML. Characterization of components of Z bands in the fibrillar flight muscle of
Drosophila melanogaster. J Cell Biol 109: 2157–2167, 1989.
Sailer M, Reuzelselke A, and Achazi RK. The calmodulin protein-kinase system of Mytilus edulis catch muscle. Comp Biochem Physiol B Biochem 96: 533–541, 1990.
Sainsbury GM and Bullard B. New proline-rich proteins in
isolated insect Z-discs. Biochem J 191: 333–339, 1980.
Sakai U, Okamoto H, Mogami K, Yamada T, and Hotta Y.
Actin with tumor related mutation is antimorphic in Drosophila
muscle: two distinct modes of myofibrillar disruption by antimorphic actins. J Biochem 107: 499 –505, 1990.
Sakai Y, Okamoto H, Mogami K, Matsuo H, and Hotta Y. Heat
shock gene activation by mutant actin is independent of myofibril
Physiol Rev • VOL
1040.
1041.
1042.
1043.
1044.
1045.
1046.
1047.
1048.
1049.
1050.
1051.
1052.
1053.
1054.
1055.
1056.
1057.
1058.
1053
degeneration in Drosophila muscle. J Biochem 109: 670 – 673,
1991.
Sakurai Y, Kanzawa N, and Maruyama K. Characterization of
myosin and paramyosin from crayfish fast and slow muscles.
Comp Biochem Physiol B Biochem 113: 105–111, 1996.
Sanchez F, Tobin SL, Rdest U, Zulauf E, and McCarthy BJ.
Two Drosophila actin genes in detail. Gene structure, protein
structure and transcription during development. J Mol Biol 163:
533–551, 1983.
Sano K, Maeda K, Taniguchi H, and Maeda Y. Amino-acid
replacements in an internal region of tropomyosin alter the properties of the entire molecule. Eur J Biochem 267: 4870 – 4877,
2000.
Sano T, Noguchi SF, Tsuchiya T, and Matsumoto JJ. Contribution of paramyosin to marine meat gel characteristics. J Food
Sci 51: 946 –950, 1986.
Sano T, Noguchi SF, Tsuchiya T, and Matsumoto JJ.
Paramyosin-myosin-actin interactions in gel formation of invertebrate muscle. J Food Sci 54: 796 – 842, 1989.
Santos ABR, Chapman MD, Aalberse RC, Vailes LD, Ferriani
VPL, Oliver C, Rizzo MC, Naspitz CK, and Arruda LK. Cockroach allergens and asthma in Brazil: identification of tropomyosin as a major allergen with potential cross-reactivity with mite
and shrimp allergens. J Allergy Clin Immunol 104: 329 –337, 1999.
Sato H and Kamiya H. Immunofluorescent localization of intermediate filaments (IFs) in helminths using anti-mammalian IFs
monoclonal antibody. J Parasitol 86: 711–715, 2000.
Satou Y and Satoh N. Two cis-regulatory elements are essential
for the muscle specific expression of an actin gene in the ascidian
embryo. Dev Growth Differ 38: 565–573, 1996.
Satou Y, Takatori N, Fujiwara S, Nishikata T, Saiga H,
Kusakabe T, Shin-i T, Kohara Y, and Satoh N. Ciona intestinalis cDNA projects: expressed sequence tag analyses and gene
expression profiles during embryogenesis. Gene 287: 83–96, 2002.
Satou Y, Takatori N, Yamada L, Mochizuki Y, Hamaguchi M,
Ishikawa H, Chiba S, Imai K, Kano S, Murakami SD, Nakayama A, Nishino A, Sasakura Y, Satoh G, Shimotori T,
Shin-i T, Shoguchi E, Suzuki MM, Takada N, Utsumi N,
Yoshida N, Saiga H, Kohara Y, and Satoh N. Gene expression
profiles in Ciona intestinalis tailbud embryos. Development 128:
2893–2904, 2001.
Satou Y, Yamada L, Mochizuki Y, Takatori N, Kawashima T,
Sasaki A, Hamaguchi M, Awazu S, Yagi K, Sasakura Y, Nakayama A, Ishikawa H, Inaba K, and Satoh N. A cDNA resource from the basal chordate Ciona intestinalis. Genesis 33:
153–154, 2002.
Sauter A, Staudenmann W, Hughes GJ, and Heizmann CW. A
novel EF-hand Ca2⫹-binding protein from abdominal muscle of
crustaceans with similarity to calcyphosine from dog thyroidea.
Eur J Biochem 227: 97–101, 1995.
Schachat FH, Harris HE, and Epstein HF. Actin from the
nematode, Caenorhabditis elegans, is a single electrofocusing
species. Biochim Biophys Acta 493: 304 –309, 1977.
Schachat FH, Harris HE, and Epstein HF. Two homogeneous
myosins in body-wall muscle of Caenorhabditis elegans. Cell 10:
721–728, 1977.
Schaller MD. unc112: a new regulator of cell-extracellular matrix adhesions? J Cell Biol 150: F9 –F11, 2000.
Scheller RH, McAllister LB, Crain WR, Durica DS, Posakony
JW, Thomas TL, Britten RJ, and Davidson EH. Organization
and expression of multiple actin genes in the sea urchin. Mol Cell
Biol 1: 609 – 628, 1981.
Schmidt J, Bodor O, Gohr L, and Kunz W. Paramyosin isoforms of Schistosoma mansoni are phosphorylated and localized
in a large variety of muscle types. Parasitology 112: 459 – 467,
1996.
Schmitz S, Clayton J, Nongthomba U, Prinz H, Veigel C,
Geeves M, and Sparrow J. Drosophila ACT88F indirect flight
muscle specific actin is not N-terminally acetylated: a mutation in
N-terminal processing affects actin function. J Mol Biol 295: 1201–
1210, 2000.
Schmitz S, Schankin CJ, Prinz H, Curwen RS, Ashton PD,
Caves LSD, Fink RHA, Sparrow JC, Mayhew PJ, and Veigel
85 • JULY 2005 •
www.prv.org
1054
1059.
1060.
1061.
1062.
1063.
1064.
1065.
1066.
1067.
1068.
1069.
1070.
1071.
1072.
1073.
1074.
1075.
1076.
1077.
1078.
1079.
1080.
SCOTT L. HOOPER AND JEFFREY B. THUMA
C. Molecular evolutionary convergence of the flight muscle protein arthrin in Diptera and Hemiptera. Mol Biol Evol 20: 2019 –
2033, 2003.
Schriefer LA and Waterston RH. Phosphorylation of the Nterminal region of Caenorhabditis elegans paramyosin. J Mol Biol
207: 451– 454, 1989.
Schrimpf SP, Langen H, Gomes AV, and Wahlestedt C. A
two-dimensional protein map of Caenorhabditis elegans. Electrophoresis 22: 1224 –1232, 2001.
Schuchert P, Reber-Muller S, and Schmid V. Life stage specific
expression of a myosin heavy chain in the hydrozoan Podocoryne
carnea. Differentiation 54: 11–18, 1993.
Schuler MA and Keller EB. The chromosomal arrangement of
two linked actin genes in the sea urchin S. purpuratus. Nucleic
Acids Res 9: 591– 604, 1981.
Schuler MA, McOsker P, and Keller EB. DNA sequence of two
linked actin genes of sea urchin. Mol Cell Biol 3: 448 – 456, 1983.
Schultz JR, Tansey T, Gremke L, and Storti RV. A muscle
specific intron enhancer required for rescue of indirect flight
muscle and jump muscle function regulates Drosophila tropomyosin I gene expression. Mol Cell Biol 11: 1901–1911, 1991.
Seeber F, Höfle W, Kern A, and Lucius R. Onchocerca volvulus
and Acanthocheilonema viteae: cloning of cDNAs for muscle-cell
intermediate filaments. Parasitol Res 80: 699 –702, 1994.
Ségalat L. Dystrophin and functionally related proteins in the
nematode Caenorhabditis elegans. Neuromuscular Disorders 12:
S105–S109, 2002.
Ségalat L and Neri C. C. elegans comme modèle pour les maladies dégénératives héréditaires humaines. Med Sci Paris 19:
1218 –1225, 2003.
Seki N, Niki T, Ishikawa D, Kimura M, and Nozawa H. Preservation of scallop adductor muscle in oxygenated artificial seawater. J Food Sci 69: FCT262–FCT267, 2004.
Sellers JR. Myosins: a diverse superfamily. Biochim Biophys
Acta 1496: 3–22, 2000.
Sellers JR, Goodson HV, and Wang F. A myosin family reunion.
J Muscle Res Cell Motil 17: 7–22, 1996.
Sellers JR and Harvey EV. Purification of myosin light chain
kinase from Limulus muscle. Biochemistry 23: 5821–5826, 1984.
Serwe M, Meyer HE, Craig AG, Carlhoff D, and D’Haese J.
Complete amino acid sequence of the regulatory light chain of
obliquely striated muscle myosin from earthworm, Lumbricus
terrestris. Eur J Biochem 211: 341–346, 1993.
Shanti KN, Martin BM, Nagpal S, Metcalfe DD, and Rao PV.
Identification of tropomyosin as the major shrimp allergen and
characterization of its IgE binding epitopes. J Immunol 151:
5354 –5363, 1993.
Shelud’ko N, Khaitlina SY, and Tskhovrebova LA. ␣-Actinin
of the scallop cross-striated muscle. Physico-chemical properties.
Biophysika 25: 164 –167, 1980.
Shelud’ko N, Permjakova T, Tuturova K, Neverkina O, and
Drozdov A. Myorod, a thick filament protein in molluscan smooth
muscles: isolation, polymerization, and interaction with myosin.
J Muscle Res Cell Motil 22: 91–100, 2001.
Shelud’ko N, Tuturova K, Permyakova T, Tyurina O, Matusovskaya G, and Matusovsky O. Proteolytic substructure of
myorod, a thick filament protein of molluscan smooth muscles.
Comp Biochem Physiol B Biochem 133: 69 –75, 2002.
Shelud’ko NS, Tuturova KF, Permyakova TV, Plotnikov SV,
and Orlova A. A novel thick filament protein in smooth muscles
of bivalve molluscs. Comp Biochem Physiol B Biochem 122:
277–285, 1999.
Sher A, James SL, Correa-Oliveira R, Hieny S, and Pearce E.
Schistosome vaccines: current progress and future prospects.
Parasitology 98: S61–S68, 1989.
Shi FH, Zhang YB, Lin JJ, Zuo X, Shen W, Cai YM, Ye P,
Bickle QD, and Taylor MG. Field testing of Schistosoma japonicum DNA vaccines in cattle in China. Vaccine 20: 3629 –3631,
2002.
Shi FH, Zhang YB, Ye P, Lin JJ, Cai YM, Shen W, Bickle QD,
and Taylor MG. Laboratory and field evaluation of Schistosoma
japonicum DNA vaccines in sheep and water buffalo in China.
Vaccine 20: 462– 467, 2001.
Physiol Rev • VOL
1081. Shima Y, Tsuchiya T, Lehman W, and Matsumoto JJ. The
characterization of invertebrate troponin C. Comp Biochem
Physiol B Biochem 79: 525–529, 1984.
1082. Shimamura J, Maruyama K, and Kimura S. Localization of
projectin in locust flight muscle. Comp Biochem Physiol B Biochem 136: 419 – 423, 2003.
1083. Shinoda Y, Yamada A, and Yagi K. Identification of troponin I of
crayfish myofibrils. J Biochem 103: 636 – 40, 1988.
1084. Shirakata M, Takagi T, and Konishi K. Isolation and characterization of a 29,000 dalton protein from ascidian (Halocynthia
roretzi) body wall muscle. Comp Biochem Physiol B Biochem 85:
71–76, 1986.
1085. Shott RJ, Lee JJ, Britten RJ, and Davidson EH. Differential
expression of the actin gene family of Strongylocentrotus purpuratus. Dev Biol 101: 295–306, 1984.
1086. Shukle RH. Molecular and cytological characterization of an
actin gene from Hessian fly (Diptera: Cecidomyiidae). Ann Entomol Soc Am 93: 1164 –1172, 2000.
1087. Shustov AV, Kotelkin AT, Sorokin AV, Ternovoi VA, and
Loktev VB. The Opisthorchis felineus paramyosin: cDNA sequence and characterization of its recombinant fragment. Parasitol Res 88: 724 –730, 2002.
1088. Siegman MJ, Funabara D, Kinoshita S, Watabe S, Hartshorne DJ, and Butler TM. Phosphorylation of a twitchin related protein controls catch and calcium sensitivity of force production in invertebrate smooth muscle. Proc Natl Acad Sci USA
95: 5383–5388, 1998.
1089. Siegman MJ, Mooers SU, Li C, Narayan S, Trinkle-Mulcahy
L, Watabe S, Hartshorne DJ, and Butler TM. Phosphorylation
of a high molecular weight (⬃600 kDa) protein regulates catch in
invertebrate smooth muscle. J Muscle Res Cell Motil 18: 655– 670,
1997.
1090. Siemankowski RF and Zobel CR. Comparative studies on the
structure and aggregative properties of the myosin molecule. I.
The structure of the lobster myosin molecule. J Mechanochem
Cell Motil 3: 171–184, 1976.
1091. Siemankowski RF and Zobel CR. Comparative studies on the
structure and aggregative properties of the myosin molecule. III.
The in vitro aggregative properties of the lobster myosin molecule. Biochim Biophys Acta 420: 406 – 416, 1976.
1092. Siemankowski RF, Zobel CR, and Manuel H. Comparative
studies on the structure and aggregative properties of the myosin
molecule. II. The substructure of the lobster myosin molecule.
Biochim Biophys Acta 622: 25–35, 1980.
1093. Sikorski ZE and Kolodziejska I. The composition and properties of squid meat. Food Chem 20: 213–224, 1986.
1094. Sillen A, Verheyden S, Delfosse L, Braem T, Robben J,
Volckaert G, and Engelborghs Y. Mechanism of fluorescence
and conformational changes of the sarcoplasmic calcium binding
protein of the sand worm Nereis diversicolor upon Ca2⫹ or Mg2⫹
binding. Biophys J 85: 1882–1893, 2003.
1095. Silva R, Sparrow JC, and Geeves MA. Isolation and kinetic
characterisation of myosin and myosin S1 from the Drosophila
indirect flight muscles. J Muscle Res Cell Motil 24: 489 – 498, 2003.
1096. Simoncelli F, Fagotti A, DiRosa I, Panara F, Chaponnier C,
Gabbiani G, and Pascolini R. Expression of an actin in protochordates and lower craniates defined by anti-alpha SM-1. Eur
J Cell Biol 69: 297–300, 1996.
1097. Small TM, Gernert KM, Flaherty DB, Mercer KB,
Borodovsky M, and Benian GM. Three new isoforms of Caenorhabditis elegans UNC-89 containing MLCK-like protein kinase
domains. J Mol Biol 342: 91–108, 2004.
1098. Sohma H, Inoue K, and Morita F. A cAMP-dependent regulatory protein for RLC—a myosin kinase catalyzing the phosphorylation of scallop smooth muscle myosin light chain. J Biochem
103: 431– 435, 1988.
1099. Sohma H and Morita F. Characterization of regulatory light
chain—a myosin kinase from smooth muscle of scallop, Patinopecten yessoensis. J Biochem 101: 497–502, 1987.
1100. Sohma H and Morita F. Purification of a protein kinase phosphorylating myosin regulatory light chain-a (RLC-a) from smooth
muscle of scallop, Patinopecten yessoensis. J Biochem 100: 1155–
1163, 1986.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
1101. Sohma H, Sasada H, Inoue K, and Morita F. Regulatory light
chain-a myosin kinase (aMK) catalyzes phosphorylation of
smooth muscle myosin heavy chains of scallop, Patinopecten
yessoensis. J Biochem 104: 889 – 893, 1988.
1102. Sohma H, Yazawa M, and Morita F. Phosphorylation of regulatory light chain A (RLC-A) in smooth muscle myosin of scallop,
Patinopecten yessoensis. J Biochem 98: 569 –572, 1985.
1103. Soisson LA, Reid GDF, Farah IO, Nyindo M, and Strand M.
Protective immunity in baboons vaccinated with a recombinant
antigen or radiation-attenuated cercariae of Schistosoma mansoni is antibody-dependent. J Immunol 151: 4782– 4789, 1993.
1104. Soisson LA and Strand M. Schistosoma mansoni-induction of
protective immunity in rats using a recombinant fragment of a
parasite surface antigen. Exp Parasitol 77: 492– 494, 1993.
1105. Soisson LMA, Masterson CP, Tom TD, Mcnally MT, Lowell
GH, and Strand M. Induction of protective immunity in mice
using a 62-Kda recombinant fragment of a Schistosoma mansoni
surface antigen. J Immunol 149: 3612–3620, 1992.
1106. Sotelo CG, Pineiro C, Perez-Martin RI, and Gallardo JM.
Analysis of fish and squid myofibrillar proteins by capillary sodium dodecyl sulfate gel electrophoresis: actin and myosin quantification. Eur Food Res Technol 211: 443– 448, 2000.
1107. Southgate R and Ayme-Southgate A. Alternative splicing of a
region generates multiple amino terminal PEVK-like isoforms of
Drosophila projectin. J Mol Biol 313: 1035–1043, 2001.
1108. Sparrow J, Drummond D, Peckham M, Hennessey E, and
White D. Protein engineering and the study of muscle contraction
in Drosophila flight muscles. J Cell Sci Suppl 14: 73–78, 1991.
1109. Sparrow J, Reedy M, Ball E, Kyrtatas V, Molloy J, Durston J,
Hennessey E, and White D. Functional and ultrastructural effects of a missense mutation in the indirect flight muscle specific
actin gene of Drosophila melanogaster. J Mol Biol 222: 963–982,
1991.
1110. Sparrow JC. Flight and phosphorylation. Nature 374: 592–593,
1995.
1111. Stafford WF III and Szent-Györgyi AG. Physical characterization of myosin light chains. Biochemistry 17: 607– 614, 1978.
1112. Stafford WF III, Szentkiralyi EM, and Szent-Györgyi AG.
Regulatory properties of single-headed fragments of scallop myosin. Biochemistry 18: 5273–5280, 1979.
1113. Stafford WF III and Yphantis DA. Existence and inhibition of
hydrolytic enzymes attacking paramyosin in myofibrillar extracts
of Mercenaria mercenaria. Biochem Biophys Res Commun 49:
848 – 854, 1972.
1114. Standiford DM, Davis MB, Miedema K, Franzini-Armstrong
C, and Emerson CP. Myosin rod protein: a novel thick filament
component of Drosophila muscle. J Mol Biol 265: 40 –55, 1997.
1115. Standiford DM, Davis MB, Sun WT, and Emerson CP. Splice
junction elements and intronic sequences regulate alternative
splicing of the Drosophila myosin heavy chain gene transcript.
Genetics 147: 725–741, 1997.
1116. Standiford DM, Sun WT, Davis MB, and Emerson CP. Positive
and negative intronic regulatory elements control muscle specific
alternative exon splicing of Drosophila myosin heavy chain transcripts. Genetics 157: 259 –271, 2001.
1117. Steel C, Limberger RJ, McReynolds LA, Ottesen EA, and
Nutman TB. B cell responses to paramyosin. Isotopic analysis
and epitope mapping of filarial paramyosin in patients with onchocerciasis. J Immunol 145: 3917–3923, 1990.
1118. Stewart TJ, Smith AL, and Havercroft JC. Analysis of the
complete sequence of a muscle calcium binding protein of Schistosoma mansoni. Parasitology 105: 399 – 408, 1992.
1119. Stitt AW, Fairweather I, Trudgett AG, and Johnston CF.
Localization of actin in the liver fluke, Fasciola hepatica. Parasitol Res 78: 96 –102, 1992.
1120. Strumpf D and Volk T. Kakapo, a novel cytoskeletal-associated
protein is essential for the restricted localization of the neuregulin-like factor, vein, at the muscle-tendon junction site. J Cell Biol
143: 1259 –1270, 1998.
1121. Sugino H, Kojima N, Nishita K, and Ojima T. Characterization
and partial amino acid sequence of CNBR fragments of scallop
troponin C. Nippon Suisan Gakk 55: 333–340, 1989.
Physiol Rev • VOL
1055
1122. Sutter SB, Raeker MO, Borisov AB, and Russell MW. Orthologous relationship of obscurin and Unc-89: phylogeny of a novel
family of tandem myosin light chain kinases. Dev Genes Evol 214:
352–359, 2004.
1123. Suzuki MG, Shimada T, and Kobayashi M. Bm kettin, homologue of the Drosophila kettin gene, is located on the Z chromosome in Bombyx mori and is not dosage compensated. Heredity
82: 170 –179, 1999.
1124. Suzuki MM and Satoh N. Genes expressed in the amphioxus
notochord revealed by EST analysis. Dev Biol 224: 168 –177, 2000.
1125. Suzuki S, Kawamura M, and Maruyama K. Particle length and
stability of natural F-actin from adductor muscle of the clam,
Meretrix meretrix. Comp Biochem Physiol A Physiol 38: 147–155,
1971.
1126. Swalla BJ, White ME, Zhou J, and Jeffery WR. Heterochronic
expression of an adult muscle actin gene during ascidian larval
development. Dev Genet 15: 51– 63, 1994.
1127. Swank DM, Bartoo ML, Knowles AF, Iliffe C, Bernstein SI,
Molloy JE, and Sparrow JC. Alternative exon-encoded regions
of Drosophila myosin heavy chain modulate ATPase rates and
actin sliding velocity. J Biol Chem 276: 15117–15124, 2001.
1128. Swank DM, Knowles AF, Kronert WA, Suggs JA, Morrill GE,
Nikkhoy M, Manipon GG, and Bernstein SI. Variable N-terminal regions of muscle myosin heavy chain modulate ATPase rate
and actin sliding velocity. J Biol Chem 278: 17475–17482, 2003.
1129. Swank DM, Knowles AF, Suggs JA, Sarsoza F, Lee A,
Maughan DW, and Bernstein SI. The myosin converter domain
modulates muscle performance. Nat Cell Biol 4: 312–316, 2002.
1130. Swank DM, Kronert WA, Bernstein SI, and Maughan DW.
Alternative N-terminal regions of Drosophila myosin heavy chain
tune muscle kinetics for optimal power output. Biophys J 87:
1805–1814, 2004.
1131. Swank DM and Maughan DW. Rates of force generation in
Drosophila fast and slow muscle types have opposite responses to
phosphate. Adv Exp Med Biol 538: 459 – 468, 2003.
1132. Swank DM, Wells L, Kronert WA, Morrill GE, and Bernstein
SI. Determining structure/function relationships for sarcomeric
myosin heavy chain by genetic and transgenic manipulation of
Drosophila. Microsc Res Tech 50: 430 – 442, 2000.
1133. Swanson CJ. Occurrence of paramyosin among the nematomorpha. Nat New Biol 232: 122–123, 1971.
1134. Sweeney HL and Holzbaur ELF. Mutational analysis of motor
proteins. Annu Rev Physiol 58: 751–792, 1996.
1135. Szent-Györgyi AG. Meromyosins, the subunits of myosin. Arch
Biochem Biophys 42: 305–320, 1953.
1136. Szent-Györgyi AG. Milestone in physiology—the early history of
the biochemistry of muscle contraction. J Gen Physiol 123: 631–
641, 2004.
1137. Szent-Györgyi AG, Cohen C, and Kendrick-Jones J. Paramyosin and the filaments of molluscan “catch” muscles. II. Native
filaments: isolation and characterization. J Mol Biol 56: 239 –258,
1971.
1138. Szent-Györgyi AG, Cohen C, and Philpott DE. Light meromyosin fraction. I. A helical molecule from myosin. J Mol Biol 2:
133–142, 1960.
1139. Szent-Györgyi AG, Szentkiralyi EM, and Kendrick-Jones J.
The light chains of scallop myosin as regulatory subunits. J Mol
Biol 74: 179 –203, 1973.
1140. Szentkiralyi EM. Tryptic digestion of scallop S1: evidence for a
complex between the two light-chains and a heavy-chain peptide.
J Muscle Res Cell Motil 5: 147–164, 1984.
1141. Takagi T and Cox JA. Amino acid sequences of four isoforms of
amphioxus sarcoplasmic calcium binding proteins. Eur J Biochem 192: 387–399, 1990.
1142. Takagi T and Cox JA. Primary structure of the target of calcium
vector protein of amphioxus. J Biol Chem 265: 19721–19727, 1990.
1143. Takagi T, Kobayashi T, and Konishi K. Amino-acid sequence of
sarcoplasmic calcium-binding protein from scallop (Patinopecten
yessoensis) adductor striated muscle. Biochim Biophys Acta 787:
252–257, 1984.
1144. Takagi T and Konishi K. Amino acid sequence of troponin C
obtained from ascidian (Halocynthia roretzi) body wall muscle.
J Biochem 94: 1753–1760, 1983.
85 • JULY 2005 •
www.prv.org
1056
SCOTT L. HOOPER AND JEFFREY B. THUMA
1145. Takagi T and Konishi K. Amino acid sequence of ␣ chain of
sarcoplasmic calcium binding protein obtained from shrimp tail
muscle. J Biochem 95: 1603–1615, 1984.
1146. Takagi T, Konishi K, and Cox JA. The amino acid sequence of
two sarcoplasmic calcium-binding proteins from the protochordate Amphioxus. Biochemistry 25: 3585–3592, 1986.
1147. Takagi T and Konishi Y. Amino acid sequence of the ␤ chain of
sarcoplasmic calcium binding protein (SCP) obtained from
shrimp tail muscle. J Biochem 96: 59 – 67, 1984.
1148. Takagi T, Kudoh S, and Konishi K. The amino acid sequence of
ascidian (Halocynthia roretzi) myosin light chains. Biochim Biophys Acta 874: 318 –325, 1986.
1149. Takagi T, Petrova T, Comte M, Kuster T, Heizmann CW, and
Cox JA. Characterization and primary structure of amphioxus
troponin C. Eur J Biochem 221: 537–546, 1994.
1150. Takagi T, Valette-Talbi L, and Cox JA. Primary structure of
three minor isoforms of amphioxus sacroplasmic calcium-binding
protein. FEBS Lett 302: 159 –160, 1992.
1151. Takagi T, Yasunaga H, and Nakamura A. Structure of 29-Kda
protein from ascidian (Halocynthia roretzi) body wall muscle.
J Biochem 113: 321–326, 1993.
1152. Takahashi M and Morita F. An activating factor (tropomyosin)
for the superprecipitation of actomyosin prepared from scallop
adductor muscles. J Biochem 99: 339 –347, 1986.
1153. Takahashi S and Maruyama K. Activity changes in myosin
ATPase during metamorphosis of fruitfly. Zool Sci 4: 833– 838,
1987.
1154. Takahashi S, Takano-Ohmuro H, and Maruyama K. Regulation of Drosophila myosin ATPase activity by phosphorylation of
myosin light chains. I. Wild-type fly. Comp Biochem Physiol B
Biochem 95: 179 –181, 1990.
1155. Takano-Ohmuro H, Hirose E, and Mikawa T. Separation and
identification of Drosophila myosin light chains. J Biochem 94:
967–974, 1983.
1156. Takano-Ohmuro H, Takahashi S, Hirose G, and Maruyama K.
Phosphorylated and dephosphorylated myosin light chains of
Drosophila fly and larva. Comp Biochem Physiol B Biochem 95:
171–177, 1990.
1157. Tanaka H, Maita T, Ojima T, Nishita K, and Matsuda G.
Amino acid sequence of the regulatory light chain of clam foot
muscle myosin. J Biochem 103: 572–580, 1988.
1158. Tanaka H, Ojima T, and Nishita K. Amino acid sequence of
troponin I from Akazara scallop striated adductor muscle. J Biochem 124: 304 –310, 1998.
1159. Tanaka Y. Identification of Drosophila thorax proteins and their
distribution in two types of muscles. Proc Jpn Acad B Physiol 69:
191–196, 1993.
1160. Tanaka Y, Maruyama K, Mikawa T, and Hotta Y. Identification
of calcium binding proteins in two dimensional gel electrophoretic pattern of Drosophila thorax and their distribution in
two types of muscles. J Biochem 104: 489 – 491, 1988.
1161. Tanii I, Osafune M, Arata T, and Inoue A. ATPase characteristics of myosin from nematode Caenorhabditis elegans purified
by an improved method. Formation of myosin-phosphate-ADP
complex and ATP-induced fluorescence enhancement. J Biochem
98: 1201–1209, 1985.
1162. Tanikawa M, Ueyama K, and Maruyama K. Instability of insect
myosin ATPase activity and its protection. Comp Biochem
Physiol B Biochem 86: 63– 65, 1987.
1163. Tansey T, Mikus MD, Dumoulin M, and Storti RV. Transformation and rescue of a flightless Drosophila tropomyosin mutant.
EMBO J N: 1375–1385, 1987.
1164. Tansey T, Schultz JR, Miller RC, and Storti RV. Small differences in Drosophila tropomyosin expression have significant effects on muscle function. Mol Cell Biol 11: 6337– 6342, 1991.
1165. Taylor MG, Huggins MC, Shi FH, Lin JJ, Tian E, Ye P, Shen
W, Qian CG, Lin BF, and Bickle QD. Production and testing of
Schistosoma japonicum candidate vaccine antigens in the natural
ovine host. Vaccine 16: 1290 –1298, 1998.
1166. Taylor MJ, Jenkins RJ, and Bianco AE. Protective immunity
induced by vaccination with Onchocerca volvulus tropomyosin in
rodents. Parasite Immunol 18: 219 –225, 1996.
Physiol Rev • VOL
1167. Teichmann SA and Chothia C. Immunoglobulin superfamily
proteins in Caenorhabditis elegans. J Mol Biol 296: 1367–1383,
2000.
1168. Terami H, Williams BD, Kitamura S, Sakube Y, Matsumoto S,
Doi S, Obinata T, and Kagawa H. Genomic organization, expression, and analysis of the troponin C gene pat-10 of Caenorhabditis elegans. J Cell Biol 146: 193–202, 1999.
1169. Thèret I, Baladi S, Cox JA, Gallay J, Sakamoto H, and
Craescu CT. Solution structure and backbone dynamics of the
defunct domain of calcium vector protein. Biochemistry 40:
13888 –13897, 2001.
1170. Thèret I, Baladi S, Cox JA, Sakamoto H, and Craescu CT.
Sequential calcium binding to the regulatory domain of calcium
vector protein reveals functional asymmetry and a novel mode of
structural rearrangement. Biochemistry 39: 7920 –7926, 2000.
1171. Thèret I, Cox JA, Mispelter J, and Craescu CT. Backbone
dynamics of the regulatory domain of calcium vector protein,
studied by N-15 relaxation at four fields, reveals unique mobility
characteristics of the intermotif linker. Protein Sci 10: 1393–1402,
2001.
1172. Thomas WR and Smith W. Towards defining the full spectrum of
important house dust mite allergens. Clin Exp Allergy 29: 1583–
1587, 1999.
1173. Thomas WR, Smith WA, Hales BJ, Mills KL, and O’Brien RM.
Characterization and immunobiology of house dust mite allergens. Int Arch Allergy Immunol 129: 1–18, 2002.
1174. Tien-Chin T, Pei-Hsing T, and Chia-Mu P. A comparative
physio-chemical study of tropomyosins from different sources.
Sci Sin 5: 91–111, 1955.
1175. Tobin SL, Cook PJ, and Burn TC. Transcripts of individual
Drosophila actin genes are differentially distributed during embryogenesis. Dev Genet 11: 15–26, 1990.
1176. Tobin SL, Zulauf E, Sanchez F, Craig EA, and McCarthy BJ.
Multiple actin related sequences in the Drosophila melanogaster
genome. Cell 19: 121–131, 1980.
1177. Tobita T, Hiraide F, Miyazaki J, and Ishimoda Takagi T.
Muscle type tropomyosin of sea urchin egg increases the actin
binding of nonmuscle type tropomyosin. J Biochem 120: 922–928,
1996.
1178. Toffenetti J, Mischke D, and Pardue ML. Isolation and characterization of the gene for myosin light chain two of Drosophila
melanogaster. J Cell Biol 104: 19 –28, 1987.
1179. Tohtong R, Rodriguez D, Maughan D, and Simcox A. Analysis
of cDNAs encoding Drosophila melanogaster myosin light chain
kinase. J Muscle Res Cell Motil 18: 43–56, 1997.
1180. Tohtong R, Yamashita H, Graham M, Haeberle J, Simcox A,
and Maughan D. Impairment of muscle function caused by mutations of phosphorylation sites in myosin regulatory light chain.
Nature 374: 650 – 653, 1995.
1181. Tomlinson CR, Bates WR, and Jeffery WR. Development of a
muscle actin specified by maternal and zygotic mRNA in ascidian
embryos. Dev Biol 123: 470 – 482, 1987.
1182. Tomlinson CR, Beach RL, and Jeffery WR. Differential expression of a muscle actin gene in muscle cell lineages of ascidian
embryos. Development 101: 751–765, 1987.
1183. Toyota N, Obinata R, and Terakado K. Isolation of troponintropomyosin-containing thin filaments from ascidian smooth muscle. Comp Biochem Physiol B Biochem 62: 433– 441, 1979.
1184. Trinick J. Elastic filaments and giant proteins in muscle. Curr
Opin Cell Biol 3: 112–119, 1991.
1185. Trinick J. Titin and nebulin: protein rulers in muscle? Trends
Biochem Sci 19: 405– 409, 1994.
1186. Trinick J. Titin as a scaffold and spring. Cytoskeleton Curr Biol
6: 258 –260, 1996.
1187. Tsai LC, Chao PL, Hung MW, Sun YC, Kuo IC, Chua KY, Liaw
SH, Chua KY, and Kuo IC. Protein sequence analysis and mapping of IgE and IgG epitopes of an allergenic 98-kDa Dermatophagoides farinae paramyosin, Der f 11. Allergy 55: 141–147,
2000.
1188. Tsai LC, Chao PL, Shen HD, Tang RB, Chang TC, Chang ZN,
Hung MW, Lee BL, and Chua KY. Isolation and characterization
of a novel 98-kd Dermatophagoides farinae mite allergen. J Allergy Clin Immunol 102: 295–303, 1998.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
1189. Tsai LC, Sun YC, Chao PL, Ng HP, Hung MW, Hsieh KH, Liaw
SH, and Chua KY. Sequence analysis and expression of a cDNA
clone encoding a 98-kDa allergen in Dermatophagoides farinae.
Clin Exp Allergy 29: 1606 –1613, 1999.
1190. Tskhovrebova L and Trinick J. Titin: properties and family
relationships. Nat Rev Mol Cell Biol 4: 679 – 689, 2003.
1191. Tsuchiya T. The muscle structure and its composition proteins of
invertebrates— especially about squid and octopus. Cook Sci 21:
159 –166, 1988.
1192. Tsuchiya T and Ehara T. Gel formability of the muscles of
marine invertebrates. Suisan Neriseihin Gijyutu Kenkyu Kaishi
17: 242–257, 1991.
1193. Tsuchiya T, Head JF, and Lehman W. The isolation and characterization of a troponin-C-like protein from the mantle muscle
of the squid Loligo pealei. Comp Biochem Physiol B Biochem 71:
507–509, 1982.
1194. Tsuchiya T, Ikeda N, Obara K, and Hartshorne DJ. A type 2A
protein phosphatase from clam smooth muscle. Use of 4-methylumbelliferyl phosphate as substrate. Comp Biochem Physiol B
Biochem 118: 17–21, 1997.
1195. Tsuchiya T, Obinata T, Sato M, Mori T, Suzuki E, and
Amemiya S. Non-catch contraction in paramyosin containing
muscle in an echinothuriid sea urchin Asthenosoma ijjimai. J
Exp Biol 162: 361–365, 1992.
1196. Tsuchiya T, Shidara H, Nomaguchi K, Sano T, and Matsumoto JJ. Isolation and some properties of squid actinin. Nippon
Suisan Gakk 51: 2059 –2065, 1985.
1197. Tsuchiya T, Shinohara T, and Matsumoto JJ. Physiochemical
properties of squid tropomyosin. Nippon Suisan Gakk 46: 893–
896, 1980.
1198. Tzolovsky G, Millo H, Pathirana S, Wood T, and Bownes M.
Identification and phylogenetic analysis of Drosophila melanogaster myosins. Mol Biol Evol 19: 1041–1052, 2002.
1199. Uddin M, Ahmad MU, Jahan P, and Sanguandeekul R. Differential scanning calorimetry of fish and shellfish meat. Asian
J Chem 13: 965–968, 2001.
1200. Uddin M, Ishizaki S, Ishida M, and Tanaka M. Assessing the
end-point temperature of heated fish and shellfish meats. Fish Sci
68: 768 –775, 2002.
1201. Umemiya T, Takeichi M, and Nose A. M-spondin, a novel ECM
protein highly homologous to vertebrate F-spondin, is localized at
the muscle attachment sites in the Drosophila embryo. Dev Biol
186: 165–176, 1997.
1202. Valette-Talbi L, Comte M, Chaponnier C, and Cox JA. Immunolocalization of calcium vector protein and its target protein in
amphioxus. Histochemistry 100: 73– 81, 1993.
1203. Vandekerckhove J and Weber K. At least six different actins are
expressed in a higher mammal: an analysis based on the amino
acid sequence of the amino-terminal tryptic peptide. J Mol Biol
126: 783– 802, 1978.
1204. Vandekerckhove J and Weber K. Chordate muscle actins differ
distinctly from invertebrate muscle actins. The evolution of the
different vertebrate muscle actins. J Mol Biol 179: 391– 413, 1984.
1205. Van Straaten M, Goulding D, Kolmerer B, Labeit S, Clayton
J, Leonard K, and Bullard B. Association of kettin with actin in
the Z disc of insect flight muscle. J Mol Biol 285: 1549 –1562, 1999.
1206. Varadaraj K, Kumari SS, and Skinner DM. Actin-encoding
cDNAs and gene expression during the intermolt cycle of the
Bermuda land crab Gecarcinus lateralis. Gene 171: 177–184, 1996.
1207. Vargas-Parada L and Laclette JP. Gene structure of Taenia
solium paramyosin. Parasitol Res 89: 375–378, 2003.
1208. Vásquez-Talavera J, Solı́s CF, Medina-Escutia E, MoralesLópex Z, Proaño J, Correa D, and Laclette JP. Human T and
B cell epitope mapping of Taenia solium paramyosin. Parasite
Immunol 23: 575–579, 2001.
1209. Vásquez-Talavera J, Solı́s CF, Terrazas LI, and Laclette JP.
Characterization and protective potential of the immune response
to Taenia solium paramyosin in a murine model of cisticercosis.
Infect Immun 69: 5412–5416, 2001.
1210. Venier P, Pallvicini A, De Nardi B, and Lanfranchi G. Towards
a catalogue of genes transcribed in multiple tissues of Mytilus
galloprovincialis. Gene 314: 29 – 40, 2003.
Physiol Rev • VOL
1057
1211. Venkatesh B, Tay BH, Elgar G, and Brenner S. Isolation,
characterization, and evolution of nine pufferfish (Fugu rubripes)
actin genes. J Mol Biol 259: 655– 665, 1996.
1212. Venolia L, Ao WY, Kim S, Kim C, and Pilgrim D. unc-45 gene
of Caenorhabditis elegans encodes a muscle specific tetratricopeptide repeat containing protein. Cell Motil Cytoskeleton 42:
163–177, 1999.
1213. Venolia L and Waterston RH. The unc-45 gene of Caenorhabditis elegans is an essential muscle affecting gene with maternal
expression. Genetics 126: 345–353, 1990.
1214. Verjovski-Almeida S, DeMarco R, Martins EAL, Guimaraes
PEM, Ojopi EPB, Paquola ACM, Piazza JP, Nishiyama MY,
Kitajima JP, Adamson RE, Ashton PD, Bonaldo MF, Coulson
PS, Dillon GP, Farias LP, Gregorio SP, Ho PL, Leite RA,
Malaquias LCC, Marques RCP, Miyasato PA, Nascimento
ALTO, Ohlweiler FP, Reis EM, Ribeiro MA, Sa RG, Stukart
GC, Soares MB, Gargioni C, Kawano T, Rodrigues V, Madeira AMBN, Wilson RA, Menck CFM, Setubal JC, Leite
LCC, and Dias-Neto E. Transcriptome analysis of the acoelomate human parasite Schistosoma mansoni. Nature Genet 35:
148 –157, 2003.
1215. Verjovski-Almeida S, Leite LCC, Dias-Neto E, Menck CFM,
and Wilson RA. Schistosome transcriptome: insights and perspective for functional genomics. Trends Parasitol 20: 304 –308,
2004.
1216. Verrez-Bagnis V and Jérôme M. Physicochemical properties
and rheological behavior of Patella caerula paramyosin. Z Lebensm Unters Forsch 201: 230 –235, 1995.
1217. Vibert P. Domains, motions, and regulation in the myosin head.
J Muscle Res Cell Motil 9: 296 –305, 1988.
1218. Vibert P and Castellani L. Substructure and accessory proteins
in scallop myosin filaments. J Cell Biol 109: 539 –547, 1989.
1219. Vibert P and Craig R. Three-dimensional reconstruction of thin
filaments decorated with a Ca2⫹-regulated myosin. J Mol Biol 157:
299 –319, 1982.
1220. Vibert P, Edelstein SM, Castellani L, and Elliott BW. Minititins in striated and smooth molluscan muscles—structure, location, and immunological cross-reactivity. J Muscle Res Cell Motil
14: 598 – 607, 1993.
1221. Vibert P, York ML, Castellani L, Edelstein S, Elliott B, and
Nyitrai L. Structure and distribution of mini-titins. Adv Biophys
33: 199 –209, 1996.
1222. Vibert PJ. Domain structure of the myosin head in correlation
averaged images of shadowed molecules. J Muscle Res Cell Motil
9: 147–155, 1988.
1223. Vigoreaux JO. Genetics of the Drosophila flight muscle myofibril: a window into the biology of complex systems. Bioessays 23:
1047–1063, 2001.
1224. Vigoreaux JO. Alterations in flightin phosphorylation in Drosophila flight muscles are associated with myofibrillar defects
engendered by actin and myosin heavy-chain mutant alleles. Biochem Genet 32: 301–314, 1994.
1225. Vigoreaux JO, Hernandez C, Moore J, Ayer G, and Maughan
D. A genetic deficiency that spans the flightin gene of Drosophila
melanogaster affects the ultrastructure and function of the flight
muscles. J Exp Biol 201: 2033–2044, 1998.
1226. Vigoreaux JO, Moore JR, and Maughan DW. Role of the elastic
protein projectin in stretch activation and work output of Drosophila flight muscles. Adv Exp Med Biol 481: 237–250, 2000.
1227. Vigoreaux JO and Perry LM. Multiple isoelectric variants of
flightin in Drosophila stretch-activated muscles are generated by
temporally regulated phosphorylations. J Muscle Res Cell Motil
15: 607– 616, 1994.
1228. Vigoreaux JO, Saide JD, and Pardue ML. Structurally different
Drosophila striated muscles utilize distinct variants of Z band
associated proteins. J Muscle Res Cell Motil 12: 340 –354, 1991.
1229. Vigoreaux JO, Saide JD, Valgeirsdottir K, and Pardue ML.
Flightin, a novel myofibrillar protein of Drosophila stretch activated muscles. J Cell Biol 121: 587–598, 1993.
1230. Vijay-Kumar S and Cook WJ. Structure of a sarcoplasmic calcium-binding protein from Nereis diversicolor refined at 2 Å
resolution. J Mol Biol 224: 413– 426, 1992.
85 • JULY 2005 •
www.prv.org
1058
SCOTT L. HOOPER AND JEFFREY B. THUMA
1231. Vinós J, Domingo A, Marco R, and Cervera M. Identification
and characterization of Drosophila melanogaster paramyosin. J
Mol Biol 220: 687–700, 1991.
1232. Vinós J, Maroto M, Garesse R, Marco R, and Cervera M.
Drosophila melanogaster paramyosin: developmental pattern,
mapping, and properties deduced from its complete coding sequence. Mol Gen Genet 231: 385–394, 1992.
1233. Volk T. A new member of the spectrin superfamily may participate in the formation of embryonic muscle attachments in Drosophila. Development 116: 721–730, 1992.
1234. Von Ebner V. über den Bau der Chorda dorsalis des Amphioxus
(Branchiostoma lanceolatus). Anz k Akad Wiss Wien 32: 213–214,
1895.
1235. Vorobiev S, Strokopytov B, Drubin DG, Frieden C, Ono S,
Condeelis J, Rubenstein PA, and Almo SC. The structure of
nonvertebrate actin: implications for the ATP hydrolytic mechanism. Proc Natl Acad Sci USA 100: 5760 –5765, 2003.
1236. Vyazunova I and Lan Q. Stage-specific expression of two actin
genes in the yellow fever mosquito, Aedes aegypti. Insect Mol Biol
13: 241–249, 2004.
1237. Wahlberg MH. Three main patterns in the expression of six actin
genes in the plerocercoid and adult Diphyllobothrium dendriticum tapeworm (Cestoda). Mol Biochem Parasitol 86: 199 –209,
1997.
1238. Wahlberg MH. The distribution of F actin during the development of Diphyllobothrium dendriticum (Cestoda). Cell Tissue
Res 291: 561–570, 1998.
1239. Wahlberg MH and Johnson MS. Isolation and characterization
of five actin cDNAs from the cestode Diphyllobothrium dendriticum: a phylogenetic study of the multigene family. J Mol Evol 44:
159 –168, 1997.
1240. Wahlberg MH, Karlstedt KA, and Paatero GIL. Cloning, sequencing and characterization of an actin cDNA in Diphyllobothrium dendriticum (Cestoda). Mol Biochem Parasitol 68: 334,
1994.
1241. Wahlberg MH, Karlstedt KA, and Paatero GIL. Cloning, sequencing, and characterization of an actin cDNA in Diphyllobothrium dendriticum (Cestoda). Mol Biochem Parasitol 65: 357–
360, 1994.
1242. Waine GJ, Becker M, Kalinna BH, Yang W, Scott J, Liui X,
Tiu W, and McManus D. Molecular vaccines against schistosomiasis: current status and the challenges ahead. Asia Pac J Mol
Biol Biotech 1: 26 –35, 1993.
1243. Walker ID and Stewart M. Paramyosin: chemical evidence for
chain heterogeneity. FEBS Lett 58: 16 –18, 1975.
1244. Walker M and Trinick J. Electron microscopy of negatively
stained scallop myosin molecules. Effect of regulatory light chain
removal on head structure. J Mol Biol 208: 469 – 475, 1989.
1245. Wallimann T, Hardwicke PMD, and Szent-Györgyi AG. Regulatory and essential light-chain interactions in scallop myosin II
Photochemical cross-linking of regulatory and essential lightchains by heterobifunctional reagents. J Mol Biol 156: 153–173,
1982.
1246. Wallimann T and Szent-Györgyi AG. An immunological approach to myosin light-chain function in thick filament linked
regulation. 1. Characterization, specificity, and cross-reactivity of
anti-scallop myosin heavy- and light-chain antibodies by competitive, solid-phase radioimmunoassay. Biochemistry 20: 1176 –
1187, 1981.
1247. Wallimann T and Szent-Györgyi AG. An immunological approach to myosin light-chain function in thick filament linked
regulation. 2. Effects of anti-scallop myosin light-chain antibodies.
Possible regulatory role for the essential light chain. Biochemistry 20: 1188 –1197, 1981.
1248. Wang F, Martin BM, and Sellers JR. Regulation of actomyosin
interactions in Limulus muscle proteins. J Biol Chem 268: 3776 –
3780, 1993.
1249. Wang J, Karabinos A, Zimek A, Meyer M, Riemer M, Riemer
D, Hudson C, Lemaire P, and Weber K. Cytoplasmic intermediate filament protein expression in tunicate development: a specific marker for the test cells. Eur J Cell Biol 81: 302–311, 2002.
Physiol Rev • VOL
1250. Wang J, Pansky A, Venuti JM, Yaffe D, and Nudel U. A sea
urchin gene encoding dystrophin-related proteins. Hum Mol
Genet 7: 581–588, 1998.
1251. Warmke J, Yamakawa M, Molloy J, Falkenthal S, and
Maughan D. Myosin light chain-2 mutation affects flight, wing
beat frequency, and indirect flight muscle contraction kinetics in
Drosophila. J Cell Biol 119: 1523–1539, 1992.
1252. Warmke JW, Kreuz AJ, and Falkenthal S. Co-localization to
chromosome bands 99E1–3 of the Drosophila melanogaster myosin light chain 2 gene and a haplo-insufficient locus that affects
flight behavior. Genetics 122: 139 –151, 1989.
1253. Wassenberg DR II, Kronert WA, O’Donnell PT, and Bernstein SI. Analysis of the 5⬘ end of the Drosophila muscle myosin
heavy chain gene. Alternatively spliced transcripts initiate at a
single site and intron locations are conserved compared with
myosin genes of other organisms. J Biol Chem 262: 10741–10747,
1987.
1254. Watabe S and Hashimoto K. Isolation and characterization of
scallop smooth adductor tropomyosin. Nippon Suisan Gakk 46:
1183–1188, 1980.
1255. Watabe S, Iwasaki K, Funabara D, Hirayama Y, Nakaya M,
and Kikuchi Y. Complete amino acid sequence of Mytilus anterior byssus retractor paramyosin and its putative phosphorylation
site. J Exp Zool 286: 24 –35, 2000.
1256. Watabe S, Kantha SS, Hashimoto K, and Kagawa H. Phosphorylation and immunological cross-reactivity of paramyosin: a
comparative study. Comp Biochem Physiol B Biochem 96: 81– 88,
1990.
1257. Watabe S, Tsuchiya T, and Hartshorne DJ. Phosphorylation of
paramyosin. Comp Biochem Physiol B Biochem 94: 813– 821,
1989.
1258. Watanabe B, Maita T, Konno K, and Matsuda G. Amino acid
sequence of LC-1 light chain of squid mantle muscle myosin. Biol
Chem Hoppe-Seyler 367: 1025–1032, 1986.
1259. Waterston R, Martin C, Craxton M, Huynh C, Coulson A,
Hillier L, Durbin R, Green P, Shownkeen R, Holloran N,
Metzstein M, Hawkins T, Wilson R, Berks M, Du Z, Thomas
K, Thierry-Mieg J, and Sulston J. A survey of expressed genes
in Caenorhabditis elegans. Nature Genet 1: 114 –123, 1992.
1260. Waterston RH. The minor myosin heavy chain, MHCA, of Caenorhabditis elegans is necessary for the initiation of thick filament assembly. EMBO J 8: 3429 –3436, 1989.
1261. Waterston RH. Molecular genetic approaches to the study of
motility in Caenorhabditis elegans. Cell Motil Cytoskeleton 14:
136 –145, 1989.
1262. Waterston RH, Epstein HF, and Brenner S. Paramyosin of
Caenorhabditis elegans. J Mol Biol 90: 285–290, 1974.
1263. Waterston RH, Fishpool RM, and Brenner S. Mutants affecting
paramyosin in Caenorhabditis elegans. J Mol Biol 117: 679 – 697,
1977.
1264. Waterston RH, Hirsh D, and Lane TR. Dominant mutations
affecting muscle structure in Caenorhabditis elegans that map
near the actin gene cluster. J Mol Biol 180: 473– 496, 1984.
1265. Waterston RH, Smith KC, and Moerman DG. Genetic fine
structure analysis of the myosin heavy chain gene unc-54 of
Caenorhabditis elegans. J Mol Biol 158: 1–15, 1982.
1266. Waterston RH, Thomson JN, and Brenner S. Mutants with
altered muscle structure of Caenorhabditis elegans. Dev Biol 77:
271–302, 1980.
1267. Weber K, Plessmann U, and Ulrich W. Cytoplasmic intermediate filament proteins of invertebrates are closer to nuclear lamins
than are vertebrate intermediate filament proteins; sequence characterization of two muscle proteins of a nematode. EMBO J 8:
3221–3227, 1989.
1268. Weber K, Riemer D, and Dodemont H. Aspects of the evolution
of the lamin intermediate filament protein family—a current analysis of invertebrate intermediate filament proteins. Biochem Soc
Trans 19: 1021–1023, 1991.
1269. Wedeen CJ and Figueroa NB. Expression of actin mRNA in
embryos of the leech Helobdella triserialis. Int J Dev Biol 42:
581–590, 1998.
85 • JULY 2005 •
www.prv.org
INVERTEBRATE MUSCLE GENES AND PROTEINS
1270. Weisel JW. Paramyosin segments: molecular orientation and interactions in invertebrate muscle thick filaments. J Mol Biol 98:
675– 681, 1975.
1271. Weitkamp B, Jurk K, and Beinbrech G. Projectin-thin filament
interactions and modulation of the sensitivity of the actomyosin
ATPase to calcium by projectin kinase. J Biol Chem 273: 19802–
19808, 1998.
1272. Wells L, Edwards KA, and Bernstein SI. Myosin heavy chain
isoforms regulate muscle function but not myofibril assembly.
EMBO J 15: 4454 – 4459, 1996.
1273. Welsch U. über den feinbau der chorda dorsalis von Branchiostoma lanceolatum. Z Zellforsch 87: 69 – 81, 1968.
1274. Werner C and Rajan TV. Characterization of a myosin heavy
chain gene from Brugia malayi. Mol Biochem Parasitol 50: 261–
268, 1992.
1275. Werner C and Rajan TV. Comparison of the body wall myosin
heavy chain sequences from Onchocerca volvulus and Brugia
malayi. Mol Biochem Parasitol 50: 255–260, 1992.
1276. Weston D, Schmitz J, Kemp WM, and Kunz W. Cloning and
sequence of a complete myosin heavy chain cDNA from Schistosoma mansoni. Mol Biochem Parasitol 58: 161–164, 1993.
1277. Weston DS and Kemp WM. Schistosoma mansoni: comparison
of cloned tropomyosin antigens shared between adult parasites
and Biomphalaria glabrata. Exp Parasitol 76: 358 –370, 1993.
1278. Westritschnig K, Sibanda E, Thomas W, Auer H, Aspock H,
Pittner G, Vrtala S, Spitzauer S, Kraft D, and Valenta R.
Analysis of the sensitization profile towards allergens in central
Africa. Clin Exp Allergy 33: 22–27, 2003.
1279. White GE, Petry CM, and Schachat F. The pathway of myofibrillogenesis determines the interrelationship between myosin
and paramyosin synthesis in Caenorhabditis elegans. J Exp Biol
206: 1899 –1906, 2003.
1280. White ME and Crother BI. Gene conversions may obscure actin
gene family relationships. J Mol Evol 50: 170 –174, 2000.
1281. White ME and Crother BI. Diagnostic amino acids in actin
genes: an idea whose time has gone. Mol Biol Evol 16: 876 – 879,
1999.
1282. Williams BD and Waterston RH. Genes critical for muscle
development and function in Caenorhabditis elegans identified
through lethal mutations. J Cell Biol 124: 475– 490, 1994.
1283. Winkelman DA, Almeda S, Vibert P, and Cohen C. A new
myosin fragment: visualization of the regulatory domain. Nature
307: 758 –760, 1984.
1284. Winkleman L. Comparative studies of paramyosins. Comp Biochem Physiol B Biochem 55: 391–397, 1976.
1285. Witteman AM, Akkerdaas JH, van Leeuwen J, van der Zee
JS, and Aalberse RC. Identification of a cross-reactive allergen
(presumably tropomyosin) in shrimp, mite, and insects. Int Arch
Allergy Immunol 105: 56 – 61, 1994.
1286. Wnuk W. Resolution and calcium-binding properties of the two
major isoforms of troponin C from crayfish. J Biol Chem 264:
18240 –18246, 1989.
1287. Wnuk W, Cox JA, Kohler LG, and Stein EA. Calcium and
magnesium binding properties of a high affinity calcium-binding
protein from crayfish sarcoplasm. J Biol Chem 254: 5284 –5289,
1979.
1288. Wnuk W, Cox JA, and Stein E. Structural changes induced by
calcium and magnesium in a high affinity calcium-binding protein
from crayfish sarcoplasm. J Biol Chem 256: 11538 –11544, 1981.
1289. Wnuk W and Jauregui-Adell J. Polymorphism in high-affinity
calcium-binding proteins from crustacean sarcoplasm. Eur J Biochem 131: 177–182, 1983.
1290. Woods EF. Subunit structure of oyster paramyosin. Biochem J
113: 39 – 47, 1969.
1291. Woods EF and Pont MJ. Characterization of some invertebrate
tropomyosins. Biochemistry 10: 270 –276, 1971.
1292. Wright TRF. The phenogenetics of the embryonic mutant, lethal
myospheroid, in Drosophila melanogaster. J Exp Zool 143: 77–99,
1960.
1293. Wylie T, Martin JC, Dante M, Mitreva MD, Clifton SW, Chinwalla A, Waterston RH, Wilson RK, and McCarter JP. Nematode net: a tool for navigating sequences from parasitic and
free-living nematodes. Nucleic Acids Res 32: D423–D426, 2004.
Physiol Rev • VOL
1059
1294. Xie X, Harrison DH, Schlichting I, Sweet RM, Kalabokis VN,
Szent-Györgyi AG, and Cohen C. Structure of the regulatory
domain of scallop myosin at 2.8 A resolution. Nature 368: 306 –
312, 1994.
1295. Xu H, Miller S, van Keulen H, Wawrzynski MR, Rekosh DM,
and LoVerde PT. Schistosoma mansoni tropomyosin: cDNA
characterization, sequence, expression, and gene product localization. Exp Parasitol 69: 373–392, 1989.
1296. Xu H, Rekosh DM, Andrews W, Higashi GI, Nicholson L, and
LoVerde PT. Schistosoma mansoni tropomyosin—production
and purification of the recombinant protein and studies on its
immunodiagnostic potential. Am J Trop Med Hyg 45: 121–131,
1991.
1297. Yagi R, Ishimaru S, Yano H, Gaul U, Hanafusa H, and Sabe H.
A novel muscle LIM-only protein is generated from the paxillin
gene locus in Drosophila. EMBO Rep 2: 814 – 820, 2001.
1298. Yamada A, Yoshio M, Kojima H, and Oiwa K. An in vitro assay
reveals essential protein components for the “catch” state of
invertebrate smooth muscle. Proc Natl Acad Sci USA 98: 6635–
6640, 2001.
1299. Yamada A, Yoshio M, Oiwa K, and Nyitray L. Catchin, a novel
protein in molluscan catch muscles, is produced by alternative
splicing from the myosin heavy chain gene. J Mol Biol 295: 169 –
178, 2000.
1300. Yamanaka H and Shimida R. Post-mortem biochemical changes
in the muscle of the Japanese spiny lobster during storage. Fish
Sci 62: 821– 824, 1996.
1301. Yamanobe T and Sugi H. Purification and characterization of a
Ca2⫹ binding 450-Kda protein (MCBP-450) in the plasma membrane enriched fraction from a molluscan smooth muscle. Biochim Biophys Acta 1149: 166 –174, 1993.
1302. Yamashita RA, Sellers JR, and Anderson JB. Identification
and analysis of the myosin superfamily in Drosophila: a database
approach. J Muscle Res Cell Motil 21: 491–505, 2000.
1303. Yang W, Jackson DC, Zeng QR, and McManus DP. Multiepitope schistosome vaccine candidates tested for protective immunogenicity in mice. Vaccine 19: 103–113, 2000.
1304. Yang W, Waine GJ, and McManus DP. Antibodies to Schistosoma japonicum (Asian bloodfluke) paramyosin induced by nucleic acid vaccination. Biochem Biophys Res Commun 212: 1029 –
1039, 1995.
1305. Yang W, Waine GJ, Sculley DG, Liu X, and McManus DP.
Cloning and partial nucleotide sequence of Schistosoma japonicum paramyosin: a potential vaccine candidate against schistosomiasis. Int J Parasitol 22: 1187–1191, 1992.
1306. Yang W, Zheng YZ, Jones MK, and McManus DP. Molecular
characterization of a calponin-like protein from Schistosoma japonicum. Mol Biochem Parasitol 98: 225–237, 1999.
1307. Yang YZ, Cun SJ, Xie XJ, Lin JH, Wei JW, Yang WL, Mou CY,
Yu CL, Ye LT, Lu Y, Fu ZY, and Xu AL. EST analysis of gene
expression in the tentacle of Cyanea capillata. FEBS Lett 538:
183–191, 2003.
1308. Yasuda E, Goto T, Makabe KW, and Satoh N. Expression of
actin genes in the arrow worm Paraspadella gotoi (Chaetognatha). Zool Sci 14: 953–960, 1997.
1309. Yazawa Y. Actin linked regulation in scallop striated muscle. Proc
Jpn Acad B Physiol 61: 497–500, 1985.
1310. Yazawa Y and Kamidochi M. The properties and function of
invertebrate new muscle protein. Mol Cell Biochem 190: 63– 66,
1999.
1311. Yeung AT and Cowgill RW. Structural difference between
␣-paramyosin and ␤-paramyosin of Mercenaria mercenaria. Biochemistry 15: 4654 – 4659, 1976.
1312. Yi FC, Cheong N, Shek PCL, Wang DY, Chua KY, and Lee BW.
Identification of shared and unique immunoglobulin E epitopes of
the highly conserved tropomyosins in Blomia tropicalis and Dermatophagoides pteronyssinus. Clin Exp Allergy 32: 1203–1210,
2002.
1313. Yokoyama Y, Takahashi S, Sakaguchi M, Kawai F, and Kanamori M. Postmorten changes of ATP and its related compounds and freshness indices in spear squid Doryteuthis bleekeri
muscles. Fish Sci 60: 583–587, 1994.
85 • JULY 2005 •
www.prv.org
1060
SCOTT L. HOOPER AND JEFFREY B. THUMA
1314. Yoneda C, Kasamatsu C, Hatae K, and Watabe S. Changes in
taste and textural properties of the foot of the Japanese cockle
(Fulvia mutica) by cooking and during storage. Fish Sci 68:
1138 –1144, 2002.
1315. Yoshida W, Kunimi O, Fufira M, Kimura M, Nozawa H, and
Seki N. Thermal gelation of salted paste from scallop striated
adductor muscle. Fish Sci 69: 1017–1025, 2003.
1316. Yoshioka T, Kinoshita Y, Yoshino H, Park S, Konno K, and
Seki N. Change in translucency of squid mantle muscle upon
storage. Fish Sci 69: 408 – 413, 2003.
1317. You M, Xuan XN, Tsuji N, Kamio T, Igarashi I, Nagasawa H,
Mikami T, and Fujisaki K. Molecular characterization of a troponin I-like protein from the hard tick Haemaphysalis longicornis. Insect Biochem Mol Biol 32: 67–73, 2001.
1318. Yu Q and Bernstein SI. UCS proteins: managing the myosin
motor. Curr Biol 12: R525–R527, 2003.
1319. Yuasa HJ, Cox JA, and Takagi T. Diversity of the troponin C
genes during chordate evolution. J Biochem 123: 1180 –1190, 1998.
1320. Yuasa HJ, Cox JA, and Takagi T. Genomic structure of the
amphioxus calcium vector protein. J Biochem 126: 572–577, 1999.
1321. Yuasa HJ, Kawamura K, Yamamoto H, and Takagi T. The
structural organization of ascidian Halocynthia roretzi troponin I
genes. J Biochem 132: 135–141, 2002.
1322. Yuasa HJ, Sato S, Yamamoto H, and Takagi T. Structure of the
ascidian, Halocynthia roretzi, troponin C gene. J Biochem 121:
671– 676, 1997.
1323. Yuasa HJ, Sato S, Yamamoto H, and Takagi T. Primary structure of troponin I isoforms from the ascidian Halocynthia roretzi.
J Biochem 122: 374 –380, 1997.
1324. Yuasa HJ, Suzuki T, and Yazawa M. Structural organization of
lower marine nonvertebrate calmodulin genes. Gene 279: 205–212,
2001.
1325. Yuasa HJ and Takagi T. The genomic structure of the scallop,
Patinopecten yessoensis, troponin C gene: a hypothesis for the
evolution of troponin C. Gene 245: 275–281, 2000.
1326. Yuasa HJ and Takagi T. Genomic structure of the sandworm,
Perinereis vancaurica tetradentata, troponin C. Gene 268: 17–22,
2001.
1327. Zelhof AC, Bao H, Hardy RW, Razzaq A, Zhang B, and Doe
CQ. Drosophila amphiphysin is implicated in protein localization
and membrane morphogenesis but not in synaptic vesicle endocytosis. Development 128: 5005–5015, 2001.
1328. Zeng W and Donelson JE. The actin genes of Onchocerca volvulus. Mol Biochem Parasitol 55: 207–216, 1992.
1329. Zengel JM and Epstein HF. Identification of genetic elements
associated with muscle structure in the nematode Caenorhabditis
elegans. Cell Motil 1: 73–97, 1980.
1330. Zervas CG and Brown NH. Integrin adhesion: when is a kinase
a kinase? Curr Biol 12: R350 –R351, 2002.
1331. Zervas CG, Gregory SL, and Brown NH. Drosophila integrinlinked kinase is required at sites of integrin adhesion to link the
Physiol Rev • VOL
1332.
1333.
1334.
1335.
1336.
1337.
1338.
1339.
1340.
1341.
1342.
1343.
1344.
1345.
1346.
1347.
cytoskeleton to the plasma membrane. J Cell Biol 152: 1007–1018,
2001.
Zhang B and Zelhof AC. Amphiphysins: raising the BAR for
synaptic vesicle recycling and membrane dynamics. Traffic 3:
452– 460, 2002.
Zhang D and Miller DJ. Immunological studies on an Onchocherca volvulus intermediate-filament protein. Parasite Immunol 17: 61– 69, 1995.
Zhang D and Miller DJ. Molecular and serological analysis of
cDNAs encoding an intermediate-filament protein from Onchocerca volvulus. Mol Biochem Parasitol 67: 175–178, 1994.
Zhang QP, Ragnauth C, Greener MJ, Shanahan CM, and
Roberts RG. The nesprins are giant actin-binding proteins, orthologous to Drosophila melanogaster muscle protein MSP-300.
Genomics 80: 473– 481, 2002.
Zhang SX and Bernstein SI. Spatially and temporally regulated
expression of myosin heavy chain alternative exons during Drosophila embryogenesis. Mech Dev 101: 35– 45, 2001.
Zhang Y, Featherstone D, Davis W, Rushton E, and Broadie
K. Drosophila D-titin is required for myoblast fusion and skeletal
muscle striation. J Cell Sci 113: 3103–3115, 2000.
Zhang Y, Taylor MG, Johansen MV, and Bickle QD. Vaccination of mice with a cocktail DNA vaccine induces a Th1-type
immune response and partial protection against Schistosoma japonicum infection. Vaccine 20: 724 –730, 2001.
Zhang YB, Taylor MG, and Bickle QD. Schistosoma japonicum
myosin: cloning, expression, and vaccination studies with the
homologue of the S. mansoni myosin fragment IrV-5. Parasite
Immunol 20: 583–594, 1998.
Zhang YB, Taylor MG, Gregoriadis G, McCrossan MV, and
Bickle QD. Immunogenicity of plasmid DNA encoding the 62 kDa
fragment of Schistosoma japonicum myosin. Vaccine 18: 2102–
2109, 2000.
Zhou S, Liu S, Song G, Xu Y, and Sun W. Protective immunity
induced by the full-length cDNA encoding paramyosin of Chinese
Schistosoma japonicum. Vaccine 18: 3196 –3204, 2000.
Zhu JG, Lin JJ, Feng XG, Wu XF, Zhou YC, and Cai YM.
Cloning and expression of actin gene of Schistosoma japonicum
Chinese strain. Acta Biochim Biophys Sin 32: 545–549, 2000.
Ziegler C. Titin-related proteins in invertebrate muscles. Comp
Biochem Physiol A Physiol 109: 823– 833, 1994.
Ziegler C. Evidence for two myosin types in indirect flight muscles. Comp Biochem Physiol B Biochem 107: 91–98, 1994.
Ziegler C, Jurk K, Kolsch B, and Beinbrech G. In vitro interactions of proteins from insect myosin filaments. Biofizika 41: 78,
1996.
Zobel CR. Proteolytic fragments from the lobster myosin molecule. Biochim Biophys Acta 536: 142–155, 1978.
Zulauf E, Sanchez F, Tobin SL, Rdest U, and McCarthy BJ.
Developmental expression of a Drosophila actin gene encoding
actin I. Nature 292: 556 –558, 1981.
85 • JULY 2005 •
www.prv.org
Scott L. Hooper and Jeffrey B. Thuma
Physiol Rev 85:1001-1060, 2005. doi:10.1152/physrev.00019.2004
You might find this additional information useful...
A corrigendum for this article has been published. It can be found at:
http://physrev.physiology.org/cgi/content/full/85/4/1417
This article cites 1343 articles, 342 of which you can access free at:
http://physrev.physiology.org/cgi/content/full/85/3/1001#BIBL
This article has been cited by 2 other HighWire hosted articles:
Titin in insect spermatocyte spindle fibers associates with microtubules, actin, myosin
and the matrix proteins skeletor, megator and chromator
L. Fabian, X. Xia, D. V. Venkitaramani, K. M. Johansen, J. Johansen, D. J. Andrew and A.
Forer
J. Cell Sci., July 1, 2007; 120 (13): 2190-2204.
[Abstract] [Full Text] [PDF]
Molecular Evolution of Ankyrin: Gain of Function in Vertebrates by Acquisition of an
Obscurin/Titin-Binding-Related Domain
A. A. Hopitzan, A. J. Baines and E. Kordeli
Mol. Biol. Evol., January 1, 2006; 23 (1): 46-55.
[Abstract] [Full Text] [PDF]
Medline items on this article's topics can be found at http://highwire.stanford.edu/lists/artbytopic.dtl
on the following topics:
Oncology .. Protein Assembly
Physiology .. Actin
Biochemistry .. Muscle Proteins
Biochemistry .. Ryanodine Receptor Calcium Release Channel
Physiology .. Tropomyosin
Physiology .. Excitation-Contraction Coupling
Updated information and services including high-resolution figures, can be found at:
http://physrev.physiology.org/cgi/content/full/85/3/1001
Additional material and information about Physiological Reviews can be found at:
http://www.the-aps.org/publications/prv
This information is current as of June 5, 2010 .
Physiological Reviews provides state of the art coverage of timely issues in the physiological and biomedical sciences. It is
published quarterly in January, April, July, and October by the American Physiological Society, 9650 Rockville Pike, Bethesda
MD 20814-3991. Copyright © 2005 by the American Physiological Society. ISSN: 0031-9333, ESSN: 1522-1210. Visit our
website at http://www.the-aps.org/.
Physiol Rev 85: 1417, 2005;
doi:10.1152/physrev.00027.2005.
CORRIGENDA
Volume 85, July 2005
Pages 1001-1060: Hooper, Scott L., and Jeffrey B. Thuma. “Invertebrate Muscles:
Muscle Specific Genes and Proteins” (http://physrev.physiology.org/cgi/content/
full/85/3/1001). On p. 1006, we regret that Table 1 was published without the last
four entries. The complete Table 1 is listed here.
TABLE
1.
Actin and myosin information for relatively little studied groups
Group/Protein
Pterobranchia actin
Annelida actin
Gastropod actin
Species
Identification Method
Muscle Versus
Cytoplasmic
Isoforms?
Reference Nos.
Saccoglossus kowalevskii
Glycera, Nereis, species
unreported
Helobdella triserialis
Two cDNAs
Protein purification
Unknown
Unknown
Unknown
Unknown
127
658
One cDNA, apparent
family
Unknown
1269
Aplysia californica
Aplysia californica
Protein purification
One cDNA, 3–5 gene
family
One cDNA
Cloning of 5⬘-flanking
and part of coding
region of one gene
Protein purification
Many tissues
including
muscle, but not
segmental
ganglia
Unknown
cDNA clone
muscle specific
Unknown
Unknown
Unknown
Unknown
401
246
Unknown
Unknown
643
366
Unknown
Unknown
658
Immunohistochemistry
Three genes
Unknown
Two genes muscle
specific
1011
1308
Protein purification or
immunohistochemistry
Expressed sequence tag
One cDNA
Unknown
Unknown
Differentially
expressed in
different muscles
Unknown
Zebra mussel
Haliotis rufescens
Lophophorata, Brachiopoda
actin
Chaetognatha actin
Glottidea, unreported
species
Sagitta friderici
Paraspadella gotoi
Chelicerata actin
Various
Cephalochordata myosin
Urochordata myosin
Amphioxus (notochord)
Ascidia
Annelida myosin
Earthworm, leech
Protein purification or
immunohistochemistry
Next column
Chaetognatha myosin
Sagitta friderici
Immunohistochemistry
Unknown
Chelicerata myosin
Various
Protein purification or
immunohistochemistry
Unknown
www.prv.org
Muscle or
Developmental Stage
Specific Expression?
Unknown
Next column
Unknown
Only in
differentiating
embryonic muscle
cells
Two isoforms in
muscle specific
fashion
Present in body wall
muscle, specificity
unknown
Unknown
0031-9333/05 $18.00 Copyright © 2005 the American Physiological Society
247, 658,
1021, 1204
1124
23, 728, 729,
784, 857
160, 1009,
1013
1011
247, 426,
1021
1417