scispace - formally typeset
Search or ask a question

Showing papers in "Journal of Muscle Research and Cell Motility in 1999"


Journal ArticleDOI
TL;DR: It is concluded that, compared with viscoelastic models, a cross-bridge-based explanation of passive muscle properties is better able to accommodate the currently known facts although, as new information becomes available, this view may need to be revised.
Abstract: The tension rise during stretch of passive skeletal muscle is biphasic, with an initial steep rise, followed by a subsequent more gradual change. The initial rise has been interpreted as being due to the presence of numbers of long-term, stable cross-bridges in resting muscle fibres. A point of weakness with the cross-bridge interpretation is that the initial stiffness reaches its peak value at muscle lengths beyond the optimum for myofilament overlap. To explain this result it has been suggested that despite the reduced overlap at longer lengths, the closer interfilament spacing and a higher sensitivity of the myofilaments to Ca2+ allows more stable cross-bridges to form. Recently the stretch responses of passive muscle have been re-examined and it has been suggested that it is not necessary to invoke cross-bridge mechanisms at all. Explanations based on a viscous resistance to interfilament sliding and mechanical properties of the elastic filaments, the gap filaments, were thought to adequately account for the observed tension changes. However, an important property of passive muscle, the dependence of stretch responses on the immediate history of contraction and length changes, thixotropy, cannot be explained simply in terms of viscous and viscoelastic properties. The review discusses the cross-bridge interpretation of muscle thixotropy and the relationship of passive stiffness to filament resting tension and latency relaxation. It is proposed that cross-bridges can exist in three states; one, responsible for the resting stiffness, requires resting levels of calcium. When, during activation, calcium levels rise, cross-bridges enter a low-force, high-stiffness state, signalled by latency relaxation, before they move to the third, force-generating state. It is concluded that, compared with viscoelastic models, a cross-bridge-based explanation of passive muscle properties is better able to accommodate the currently known facts although, as new information becomes available, this view may need to be revised.

177 citations


Journal ArticleDOI
TL;DR: The model was extended to account for slow-twitch muscle by comparing the fast-twitch model with previously published data and then changing the necessary parameters to fit the data, which accounts well for all previous findings including that of yielding.
Abstract: Interactions between physiological stimulus frequencies, fascicle lengths and velocities were analyzed in feline caudofemoralis (CF), a hindlimb skeletal muscle composed exclusively of fast-twitch fibers. Split ventral roots were stimulated asynchronously to produce smooth contractions at sub-tetanic stimulus frequencies. As described previously, the peak of the sub-tetanic force-length relationship was found to shift to longer lengths with decreases in stimulus frequency, indicating a length dependence for activation that is independent of filament overlap. The sub-tetanic force-velocity (FV) relationship was aAected strongly both by stimulus frequency and by length; decreases in either decreased the slope of the FV relationship around isometric. The shapes of the force transients following stretch or shortening revealed that these eAects were not due to a change in the instantaneous FV relationship; the relative shape of the force transients following stretch or shortening was independent of stimulus frequency and hardly aAected by length. The eAects of stimulus frequency and length on the sub-tetanic FV relationship instead appear to be caused by a time delay in the length-dependent changes of activation. In contrast to feline soleus muscle, which is composed exclusively of slow-twitch fibers, CF did not yield at sub-tetanic stimulus frequencies for the range of stretch velocities tested (up to 2 L0/s). The data presented here were used to build a model of muscle that accounted well for all of the eAects described. We extended our model to account for slow-twitch muscle by comparing our fast-twitch model with previously published data and then changing the necessary parameters to fit the data. Our slow-twitch model accounts well for all previous findings including that of yielding.

163 citations


Journal ArticleDOI
TL;DR: The main criterion in judging the success of gene therapy in Duchenne dystrophy is the correction of the pathological consequences of the absence of dystrophin, and it is of fundamental importance to obtain the largest possible and up-to-date inventory of the disorders, structural and functional, resulting from this absence.
Abstract: Duchenne muscular dystrophy is a very severe disease which a€ects the skeletal musculature in human patients. It produces muscle wasting. Necrotic ®bres are partially replaced by regeneration from dormant satellite cells, so that the disorder progresses slowly over the ®rst two decades of life. Eventually, the degenerative process takes over, muscle tissue is replaced by ®brous material and fat in®ltration, with severe deformations of the skeleton (spine) due to imbalance of muscle tone. Failure of the respiratory muscle is inevitably fatal. So far, no curative treatment is available. The disease a€ects only boys with a frequency of 1 in 3500 male births. It is one of the most frequent genetic diseases in the human race. The origin of the disease was elucidated in 1987 when it was shown that it is due to mutation(s) a€ecting a very large gene located in the p21 region of the X chromosome, and coding for a protein termed `dystrophin' (Ho€man et al., 1987a; Koenig et al., 1987). Mutations resulting in the absence of dystrophin are responsible for the very severe Duchenne phenotypes, while in milder diseases, called Becker dystrophy, dystrophin is expressed in low levels or in some truncated forms (Monaco et al., 1988; England et al., 1990). Duchenne-like dystrophies of various degrees of gravity have been observed in animal (mice, cats, dogs) and also result from mutations in the dystrophin gene. For further considerations of the relationship between mutations of the dystrophin gene and dystrophy phenotypes in human patients, see Baumbach et al. (1989) and Kunkel and Ho€man (1989). A comprehensive presentation of dystrophinopathies is found in Engel et al. (1994). In recent years various attempts have been made to restore in dystrophin-de®cient ®bres the expression of dystrophin (either complete or truncated) and of utrophin, a dystrophin-related protein. The ®nal criterion in judging the success of gene therapy in Duchenne dystrophy is the correction of the pathological consequences of the absence of dystrophin. Hence, it is of fundamental importance to obtain the largest possible and up-to-date inventory of the disorders, structural and functional, resulting from this absence. This is the speci®c purpose of the present review. However, before considering the situation in dystrophin-lacking muscles, a summary of the present knowledge on dystrophin and its associated or related proteins in normal muscles will be presented. This summary will be short as many recent and authoritative reviews are available on the structures, localizations and properties of these molecules, particularly in this journal (Winder, 1997; see also Matsumura and Campbell, 1994; Sunada and Campbell, 1995).

158 citations


Journal ArticleDOI
TL;DR: Caffeine is both a valuable tool for studying the coupling process, which remains in part to be elucidated, and a useful means for inducing a complete and reversible emptying of the Ca2‡ stores in the reticulum.
Abstract: Caffeine, the well-known extract of the coffee bean, is a 1,3,7 trimethylxanthine (Figure 5) with several cellular sites of action (e.g. adenosine receptors, phosphodiesterases, Ca2‡ storage systems) in different organs. For a general survey, see Daly (1993). In the present short review, we will concentrate on the action of this drug on processes which couple the electrical events at the surface membrane with the release of Ca2‡ from the sarcoplasmic reticulum (SR) in skeletal muscle ®bres. Here caffeine is both a valuable tool for studying the coupling process, which remains in part to be elucidated, and a useful means for inducing a complete and reversible emptying of the Ca2‡ stores in the reticulum. Caffeine has also direct and reversible effects on the contractile apparatus (Wendt and Stephenson, 1983), but these effects manifest themselves at caffeine concentrations that are normally higher (10±20 mM) than those needed for the study of the coupling process (1±7 mM). Considering the action of caffeine on the coupling process, we will ®rst present the current knowledge of the structure and function of subcellular components involved in excitation±contraction (E±C) coupling. This is followed by a presentation of early experiments with caffeine. Then, in the sections ``The interaction of caffeine and related compounds with the skeletal muscle ryanodine receptor (RyR1)'' and ``Caffeine in physiological experiments'' we will concentrate on speci®c points that are presently in the centre of muscle research activities and which illustrate the continuing interest in the action of this coffee bean ingredient. From the vast number of publications on caffeine and skeletal muscle function, only selected examples of papers will be considered. E±C coupling in skeletal muscle with special reference to the action of caffeine has been ®rst reviewed by Sandow (1965) and later on by Endo (1985), Palade et al. (1989) and by Melzer et al. (1995).

123 citations


Journal ArticleDOI
TL;DR: These integrated methods for the analysis of MHCs at the protein level demonstrate that equine skeletal muscle does not express the MHC-IIB, so type II fibres have been misclassified in numerous previous studies based upon the very traditional mATPase histochemistry.
Abstract: Combined methodologies of enzyme-linked immunosorbent assay (ELISA), sodium dodecyl sulphate polyacrilamide gel electrophoresis (SDS-PAGE), immunoblotting, traditional myofibrillar ATPase (mATPase) histochemistry and immunocytochemistry of whole biopsied samples were used to study myosin heavy chain (MHC) isoforms in the equine gluteus medius muscle. The ELISA technique allowed the quantification of the three MHC isoforms known to be present in different horse muscles: slow (MHC-I) and two fast (termed MHC-IIA and MCH-IIX). The SDS-PAGE method resolved MHCs in three bands: MHC-I, MHC-IIX and MHC-IIA from the fastest to the slowest migrating band and a quantification by densitometry for each MHC isoform was also possible. The identity of these three MHCs was confirmed by immunoblots with specific monoclonal antibodies. Five fibre types were defined immunohistochemically according to their MHC content: I, I + IIA, IIA, the hybrid IIAX and IIX. When quantitative data obtained with the four different methodologies were combined and compared, they were consistent and, when considered together, showed significant correlation. Nevertheless, the percentage of MHC-IIA histochemically derived was underestimated, while that of MHC-IIX was overestimated in comparison with the immunocytochemical determination of these MHC isoforms. The percentage of MHC-I obtained by ELISA technique was underestimated. In short, these integrated methods for the analysis of MHCs at the protein level demonstrate that equine skeletal muscle does not express the MHC-IIB, so type II fibres have been misclassified in numerous previous studies based upon the vary traditional mATPase histochemistry. They also offer new prospects for muscle fibre typing in equine experimental studies and veterinary medicine.

94 citations


Journal ArticleDOI
TL;DR: The interaction between purified CapZ and α-actinin, a major component of Z filaments and polar links in Z-discs, is reported here, found to be independent of actin and shown to be weakened in the presence of phosphoinositides.
Abstract: CapZ is a widely distributed and highly conserved, heterodimeric protein, that nucleates actin polymerization and binds to the barbed ends of actin filaments, preventing the addition or loss of actin monomers. CapZ interaction with actin filaments was shown to be of high affinity and decreased in the presence of PIP2. CapZ was located in nascent Z-lines during skeletal muscle myofibrillogenesis before the striated appearance of thin filaments in sarcomers. In this study, the stabilization and the anchorage of thin filaments were explored through identification of CapZ partners in the Z-line. Fish (sea bass) striated white muscle and its related Z-line proteins were selected since they correspond to the simplest Z-line organization. We report here the interaction between purified CapZ and α-actinin, a major component of Z filaments and polar links in Z-discs. Affinity of CapZ for α-actinin, estimated by fluorescence and immunochemical assays, is in the μ m range. This association was found to be independent of actin and shown to be weakened in the presence of phosphoinositides. Binding contacts on the α-actinin molecule lie in the 55 kDa repetitive domain. A model including CapZ/α-actinin/titin/actin interactions is proposed considering Luther's 3D Z-line reconstruction.

91 citations


Journal ArticleDOI
TL;DR: This investigation demonstrates that the aging process in human skeletal muscle is accompanied by a modification in the muscle phenotype which is both muscle and region specific; a transformation towards a fast and fetal phenotype concomitant with an increased number of fibres with a mixture of different MyHC isoforms in the masseter; and an opposite shift towards a slower phenotype in the biceps brachii.
Abstract: The myosin heavy chain (MyHC) content in functionally different parts of the human masseter muscle of six elderly and five young adult subjects (mean age 74 and 22 years, respectively) was determined, using gel electrophoresis. The MyHC composition of the old masseter was also studied by enzyme- and immunohistochemical methods and compared with previous data for young adults. For comparison, the biceps brachii muscle of the same subjects was also analysed. The old masseter contained smaller amounts of slow and larger amounts of fast and fetal MyHCs. These differences were region-dependent and were more pronounced in the superficial portion. There was also a larger proportion of “hybrid” fibres, containing two to four MyHC isoforms (42%), compared with the young adult masseter (23%). No such differences were observed between old and young biceps. In contrast to the masseter, the old biceps contained more slow MyHC and less fast MyHC. This investigation demonstrates that the aging process in human skeletal muscle is accompanied by a modification in the muscle phenotype which is both muscle and region specific; a transformation towards a fast and fetal phenotype concomitant with an increased number of fibres with a mixture of different MyHC isoforms in the masseter; and an opposite shift towards a slower phenotype in the biceps brachii. The results might reflect differences between jaw and limb muscles in genetic programs and adaptive responses to changed functional demands following aging.

78 citations


Journal ArticleDOI
TL;DR: Because complex structural differences in adult extraocular muscles may have physiological and pathophysiological significance, the three-dimensional pattern of myosin heavy chain (MHC) isoform expression within the orbital and global layers of the muscle bellies compared with the distal tendon ends was quantitatively assessed.
Abstract: Because complex structural differences in adult extraocular muscles may have physiological and pathophysiological significance, the three-dimensional pattern of myosin heavy chain (MHC) isoform expression within the orbital and global layers of the muscle bellies compared with the distal tendon ends was quantitatively assessed. Three of the six extraocular muscles of adult rabbits were examined for immunohistologic expression of all fast, fast IIA/X, slow, neonatal and developmental MHC isoforms. The percentages of myofibers positive for each of these 5 myosin isoforms were determined in the orbital and global layers. There were relatively similar patterns of fast and slow MHC expression in the orbital and global layers of each of the three muscles examined. There were high levels of developmental MHC in the orbital layers, but significantly fewer developmental MHC positive myofibers in the global layer. The most variable expression was found with the neonatal MHC. There were significant differences between the longitudinal expression of the various isoforms in the middle of each muscle compared with the tendon end. In the orbital layer of all three muscles examined, the large numbers of fibers positive for fast MHC in the middle of the muscle dramatically decreased at the tendon end, with a concomitant increase in expression of slow myosin. There was a greater number of developmental MHC-positive myofibers at the tendon end than in the middle of the muscle in all three muscles examined. In the global layer, the IIA/X-positive myofibers comprised only half of the total number of fast-positive myofibers whereas in the orbital layer they comprised all or almost all of the fast positive myofibers. The configuration of the extraocular muscles is more complex than might be indicated by previous studies. The lateral rectus muscle had the most individual pattern of MHC expression when compared with the inferior rectus and inferior oblique muscles. Together with dramatic cross-sectional MHC fiber type differences between the orbital and global layers of the muscles, there are pronounced longitudinal differences in the proportions of myofibers expressing these five MHC isoforms in the middle region of the muscles and those in the distal tendon ends. This longitudinal progression appears to occur both within single myofibers, as well as within the series of myofibers that comprise the length of the muscle. We also confirm that the number of myofibers is reduced at the tendonous end while the cross-sectional area of each of the remaining myofibers is proportionally increased with regard to those in the muscle belly. Future studies may yet require two additional schemes for anatomic classification of the named extraocular muscles. One will be based on immunohistochemical features of their constituent myofibers as a supplement to classifications based on their electron microscopic appearance, innervation patterns or relative position with regard to the globe and orbit. Another will be based on the proportional length and longitudinal position of individual myofibers within an individual extraocular muscle.

70 citations


Journal ArticleDOI
TL;DR: Length was found to have an effect independent of PAP on the shortening half of the FV relationship and on the rate of force relaxation during the later stages of isometric tetanic force decay (slower relaxation at longer lengths), confirming that PAP has other functionally important effects beyond a simple increase in sub-maximal isometric forces.
Abstract: Activation of mammalian fast-twitch skeletal muscle induces a persistent effect known as post-activation potentiation (PAP), classically defined as an increase in force production at sub-maximal levels of activation. The underlying mechanism is thought to be phosphorylation of the myosin regulatory light chain (MRLC), which leads to an increase in the rate constant for cross-bridge attachment (Sweeney et al., 1993). If true, this suggests the hypothesis that other contractile properties should be affected during PAP. Using a feline fast-twitch whole-muscle preparation (caudofemoralis) at 37 degrees C, we observed that PAP greatly increased tetanic forces during active lengthening decreased isometric tetanic rise times and delayed isometric tetanic force relaxation. The first two of these effects were length dependent with a greater effect occurring at shorter lengths. These findings confirmed that PAP has other functionally important effects beyond a simple increase in sub-maximal isometric forces. Furthermore, length was found to have an effect independent of PAP on the shortening half of the FV relationship (less force was produced at longer lengths) and on the rate of force relaxation during the later stages of isometric tetanic force decay (slower relaxation at longer lengths). All of these findings can be explained with a simplified, two-state model of cross-bridge dynamics that accounts for the interaction of both interfilament spacing and MRLC phosphorylation on the apparent rate constants for cross-bridge attachment and detachment. These findings are largely consistent with data collected previously from reduced preparations such as skinned fibers at cold, unphysiological temperatures (e.g. 5 degrees C). One finding that could not be explained by our model was that twitch fall times in the dispotentiated state were parabolically correlated with length, whereas in the potentiated state the relationship was linear. The time course of decay of this effect did not follow the time course of force dispotentiation, suggesting that there are other activation-dependent processes occurring in parallel with MRLC phosphorylation.

69 citations


Journal ArticleDOI
TL;DR: It is concluded that the correct assembly of a cytoskeletal scaffold is a prerequisite for correct thick filament assembly and for the integration of the contractile apparatus into the myofibril.
Abstract: The development of myofibrils involves the formation of contractile filaments and their assembly into the strikingly regular structure of the sarcomere. We analysed this assembly process in cultured human skeletal muscle cells and in rat neonatal cardiomyocytes by immunofluorescence microscopy using antibodies directed against cytoskeletal and contractile proteins. In particular, the question in which temporal order the respective proteins are integrated into developing sarcomeres was addressed. Although sarcomeric myosin heavy chain is expressed as one of the first myofibrillar proteins, its characteristic A band arrangement is reached at a very late stage. In contrast, titin, then myomesin and finally C-protein (MyBP-C) gradually form a regularly arranged scaffold on stress fiber-like structures (SFLS), on non-striated myofibrils (NSMF) and on nascent striated myofibrils (naSMF). Immediately subsequent to the completion of sarcomere cytoskeleton formation, the labeling pattern of myosin changes from the continuous staining of SFLS to the periodic staining characteristic for mature myofibrils. This series of events can be seen most clearly in the skeletal muscle cell cultures and – probably due to a faster developmental progression – less well in cardiomyocytes. We therefore conclude that the correct assembly of a cytoskeletal scaffold is a prerequisite for correct thick filament assembly and for the integration of the contractile apparatus into the myofibril.

62 citations


Journal ArticleDOI
TL;DR: Using a new assay for assessing the binding of CX with the light meromyosin (LMM) portion of myosin, it is demonstrated that recombinant CX, just as the full-length protein, is able to facilitate LMM polymerization.
Abstract: Myosin-binding protein-C (MyBP-C or C-protein) is a ca. 130 kDa protein present in the thick filaments of all vertebrate striated muscle. The protein contains ten domains, each of ca. 90-100 amino acids; seven are members of the IgI family of proteins, three of the fibronectin type III family. The motifs are arranged in the following order (from N- to C-terminus): Ig-Ig-Ig-Ig-Ig-Fn-Fn-Ig-Fn-Ig. The C-terminal Ig motif (domain X or CX) contains its light meromyosin-binding site. A recombinant form of CX, beginning at Met-1027, exhibits saturable binding to myosin with an affinity comparable to the C-terminal 13 kDa chymotryptic fragment of native MyBP-C. To identify the surface in CX involved in its interaction with myosin, nine site-directed mutants (R37E, K43E, N49D, E52R, D56K, R73E, R74E, G80D and R103E) were constructed. Using a new assay for assessing the binding of CX with the light meromyosin (LMM) portion of myosin, we demonstrate that recombinant CX, just as the full-length protein, is able to facilitate LMM polymerization. Moreover, we show that residues Arg-37, Glu-52, Asp-56, Arg-73, and Arg-74 are involved in this interaction with the myosin rod. All of these amino acids interact with negatively charged residues of LMM, since the mutants R37E, R73E and R74E are unable to bind myosin, whereas E52R and D56K bind myosin with higher affinity than wild-type CX. Residues Lys-43 and Arg-103 show a small but significant influence on the binding reaction; residues Asn-49 and Gly-80 seem not to be involved in this interaction. Based on these data, a model is proposed for the interaction between MyBP-C CX and myosin filaments. In this model, CX interacts with four molecules of LMM at four different sites of the binding protein, thus explaining the effects of MyBP-C on the critical concentration of myosin polymerization.

Journal ArticleDOI
TL;DR: In this article, the authors used confocal microscopy of isolated, living muscle fibres from Xenopus and mouse to visualise transverse tubules (t-tubules) and mitochondria and to relate possible fatigue induced morphological changes in these to force depression.
Abstract: Force recovery from fatigue in skeletal muscle may be very slow. Gross morphological changes with vacuole formation in muscle cells during the recovery period have been reported and it has been suggested that this is the cause of the delayed force recovery. To study this we have used confocal microscopy of isolated, living muscle fibres from Xenopus and mouse to visualise transverse tubules (t-tubules) and mitochondria and to relate possible fatigue- induced morphological changes in these to force depression. T-tubules were stained with either RH414 or sulforhodamine B and mitochondrial staining was with either rhodamine 123 or DiOC6(3). Fatigue was produced by repeated, short tetanic contractions. Xenopus fibres displayed a marked vacuolation which started to develop about 2min after fatiguing stimulation, reached a maximum after about 30min, and then receded in about 2h. Vacuoles were never seen during fatiguing stimulation. The vacuoles developed from localised swellings of t-tubules and were mostly located in rows of mitochondria. Mitochondrial staining, however, showed no obvious alterations of mitochondrial structure. There was no clear correlation between the presence of vacuoles and force depression; for instance, some fibres showed massive vacuole formation at a time when force had recovered almost fully. Vacuole formation was not reduced by cyclosporin A, which inhibits opening of the non-specific pore in the mitochondrial inner membrane. In mouse fibres there was no vacuole formation or obvious changes in mitochondrial structure after fatigue, but still these fibres showed a marked force depression at low stimulation frequencies (‘low-frequency fatigue’). Vacuoles could be produced in mouse fibres by glycerol treatment and these vacuoles were not associated with any force decline. In conclusion, vacuoles originating from the t-tubular system develop after fatigue in Xenopus but not in mouse fibres. These vacuoles are not the cause of the delayed force recovery after fatigue.

Journal ArticleDOI
TL;DR: Cell fractionation studies indicated that annexin VII is equally distributed between cytosol and membrane fractions in undifferentiated cells, while in differentiated cells it is exclusively present in the membrane fraction.
Abstract: Annexin VII exists in a 47 kDa and a 51 kDa isoform with the 51 kDa protein being the only isoform present in skeletal muscle. Expression of the 51 kDa isoform during myogenesis and localization was studied in cells after conversion into myogenic cells by transduction with MyoD and in mouse and human myogenic cell lines. MyoD expression in NIH3T3 and C3H10T1/2 fibroblasts led to disappearance of the mRNA specific for the 47kDa isoform and appearance of the 51 kDa isoform-specific mRNA. The overall amount of annexin VII protein was reduced in myogenic converted cells. Both in undifferentiated and differentiated cells annexin VII was localized by immunofluorescence microscopy to punctate structures which were distributed all over the cell. A GFP annexin VII fusion protein showed a similar distribution. Cell fractionation studies indicated that annexin VII is equally distributed between cytosol and membrane fractions in undifferentiated cells, while in differentiated cells it is exclusively present in the membrane fraction. By sucrose gradient centrifugation of postnuclear supernatants we identified two distinct annexin VII-containing membrane populations that cofractionated with caveolin 3- and sorcin-containing membranes.

Journal ArticleDOI
TL;DR: The observed sliding velocities of fluorescently labeled F-actin are essentially identical with those reported in the literature, suggesting that sliding velocity in the motility assay is limited by additional factors beyond those thought to limit velocity of unloaded shortening in muscle fibers.
Abstract: We describe an approach that allows us to form a micro in vitro motility assay with as little myosin as can be retrieved from a short ( ∼ 10 mm) segment of a single skinned skeletal muscle fiber (diameter some 100 μm). Myosin is directly extracted from the single fiber segment by a high ionic strength solution in the presence of MgATP, and the extracted myosin is immediately applied to a miniaturized flow cell that has been pretreated with BSA. The observed sliding velocities of fluorescently labeled F-actin are essentially identical with those reported in the literature. Since at the single fiber level most muscle fibers contain only a single myosin heavy chain isoform this approach allows us to determine without additional purification steps, the sliding velocity driven by myosins with different heavy chain isoforms. In addition, this approach can be used to directly correlate under identical experimental conditions unloaded shortening velocity measured in segments of skinned muscle fibers with the in vitro sliding velocity of fluorescently labeled F-actin by extraction of myosin from the same skinned fibers. Such direct correlation was performed with different myosin heavy chain isoforms as well as at different temperatures and ionic strengths. Under all conditions studied, unloaded shortening velocity was 4- to 8-fold faster than sliding velocity in the motility assay even at high temperature (22∘ C) and ionic strengths >50 mM. This suggests that sliding velocity in the motility assay is limited by additional factors beyond those thought to limit velocity of unloaded shortening in muscle fibers. One such factor might be unspecific ionic interactions between F-actin and the substrate in the motility assay resulting in somewhat higher sensitivity for ionic strength of sliding velocity in the motility assay. This might become of special relevance when using in vitro sliding velocity in assessing functional consequences of mutations involving charged residues of actin or myosin.

Journal ArticleDOI
TL;DR: The structure, tissue and cellular distribution of these myosin isoforms, as well as mutations detected within their genes that have been found to affect the hearing process, are described in this review.
Abstract: The discovery in the past few years of a huge diversity within the myosin superfamily has been coupled with an understanding of the role of these motor proteins in various cellular functions. Extensive studies have revealed that myosin isoforms are not only involved in muscle contraction but also in crucial functions of many specialized mammalian cells such as melanocytes, kidney and intestinal brush border microvilli, nerve growth cones or inner ear hair cells. A search for genes involved in the pathology of human genetic deafness resulted in identification of three novel myosins: myosin VI, myosin VIIA and, very recently, myosin XV. The structure, tissue and cellular distribution of these myosin isoforms, as well as mutations detected within their genes that have been found to affect the hearing process, are described in this review.

Journal ArticleDOI
TL;DR: The present study documents the binding interaction of skeletal muscle sarcoplasmic reticulum (SR) transmembrane protein triadin with peripheral histidine-rich, Ca2+-binding protein (HCP), and shows it to be expressed as a single protein species, both in fast-twitch and slow-twitch fibers, and to identically localize to the I-band.
Abstract: The present study documents the binding interaction of skeletal muscle sarcoplasmic reticulum (SR) transmembrane protein triadin with peripheral histidine-rich, Ca2+-binding protein (HCP). In addition to providing further evidence that HCP coenriches with RyR1, FKBP-12, triadin and calsequestrin (CS) in sucrose-density-purified TC vesicles, using specific polyclonal antibody, we show it to be expressed as a single protein species, both in fast-twitch and slow-twitch fibers, and to identically localize to the I-band. Colocalization of HCP and triadin at junctional triads is supported by the overlapping staining pattern using monoclonal antibodies to triadin. We show a specific binding interaction between digoxigenin-HCP and triadin, using ligand blot techniques. The importance of this finding is strengthened by the similarities in binding affinity and in Ca2+ dependence, (0.1–1 mM Ca2+) of the interaction of digoxigenin-HCP with immobilized TC vesicles. Suggesting that triadin dually interacts with HCP and with CS, at distinct sites, we have found that triadin-CS interaction in overlays does not require the presence of Ca2+. Consistent with the binding of CS to triadin luminal domain (Guo and Campbell, 1995), we show that binding sites for digoxigenin-CS, although not binding sites for digoxigenin-HCP, can be recovered in the 92 kDa triadin fragment, after chymotryptic cleavage of the NH2-terminal end of the folded molecule in intact TC vesicles. These differential effects form the basis for the hypothesis that HCP anchors to the junctional membrane domain of the SR, through binding to triadin short cytoplasmic domain at the NH2 terminus. Although the function of this interaction, as such, is not well understood, it seems of potential biological interest within the more general context of the structural-functional role of triadin at the triadic junction in skeletal muscle.

Journal ArticleDOI
TL;DR: The data indicate that after 3 days of passive stretch the central and the terminal parts of the soleus muscle adapt similarly by increasing the levels of the MRFs, by decreasing the overall Levels of the fast SERCA1-type of ATPase and by partially re-establishing a neonatal mode of alternativeSERCA1 transcript splicing resulting in an increased SERCA2+-transport ATPase/1a ratio.
Abstract: The mRNA levels of the adult and the neonatal sarcoplasmic/endoplasmic reticulum Ca(2+)-ATPases (SERCA1a and SERCA1b, respectively) and those of the muscle regulatory factors (MRFs: myoD, myf-5, myogenin, MRF4) have been assessed by RT PCR in rat soleus muscles immobilized for 3 days in an extended position (passive stretch). The transcript level of the fast type SERCA1a Ca(2+)-transport ATPase decreased to half of its normal value, whereas that of neonatal SERCA1b isoform increased 5-fold above control in stretched muscles. Immunostaining of muscle cross sections showed that the fraction of fibers expressing the SERCA1a protein was decreased evenly along the length of the stretched muscles indicating that a transformation occurred of fast fibers to slow ones. The mRNA levels of MRFs were elevated 3- to 6-fold above the normal level and were distributed evenly along the length of the stretched muscles. However in the controls these transcripts were more abundant at both ends of the muscle. The stretch increased the level of myoD and immunocytochemistry showed the expression of myoD protein in a number of nuclei of the stretched muscles whereas it was practically undetectable by this method in the control muscles. Western blotting did not indicate a significant stretch-induced increase in the level of the myogenin protein, in spite of the fact that immunocytochemistry tended to show more myogenin-positive nuclei in stretched muscles as compared to the controls. These data indicate that after 3 days of passive stretch the central and the terminal parts of the soleus muscle adapt similarly by increasing the levels of the MRFs, by decreasing the overall levels of the fast SERCA1-type of ATPase and by partially re-establishing a neonatal mode of alternative SERCA1 transcript splicing resulting in an increased SERCA1b/1a ratio.

Journal ArticleDOI
TL;DR: Data suggest that SM22 isoforms are differently assorted in MF vs. SMC; SM22 is a SMC-lineage marker inasmuch as its expression occurs in an experimental condition characterized by a time-related cell phenotypic transition from MF to SMC, and cell conversion ability of serosal cells in the adult might take place via the reactivation of a specific “foetal” gene programme.
Abstract: The E-11 and 1-B8 monoclonal antibodies raised to the smooth muscle (SM)-specific SM22 protein from pig stomach were used to study the in vivo and in vitro phenotypic characteristics of myofibroblasts (MF) and SM cells (SMC) from the bladder detrusor muscle and serosal thickening of male rabbit. The 22-kDa SM22 band found in the SM extract appeared to be composed of distinct isoforms when examined in non-equilibrium two-dimensional gel electrophoresis (2D-EF): α (the most basic), β, γ, and δ (the most acidic) in the ratio of 34(α):23(β):36(γ):8(δ). Western blots of 2D-electrophoresed bladder extracts treated with E-11 and 1-B8 showed that α, β, and δ, but not γ isoforms were labeled with E-11, whereas α, β, and γ isoforms were stained with 1-B8. This differential immunoreactivity was not influenced by phosphorylation. The tissue distribution of SM22 immunostaining was heterogeneous in the bladder SM and serosal thickening developed as a consequence of partial outflow obstruction of the urinary bladder. At the cellular level, the 1-B8 and E-11 antibodies stained the SMC in a “diffuse” (the whole cytoplasm) and “honeycomb” (the peripheral cytoplasm) manner, whereas MF immunostaining was quite homogeneous. The two antibodies also reacted with cultured primary bladder SMC and MF grown in low serum conditions showing a heterogeneous SM22 cell distribution but an identical subcellular localization, i.e., the actin-containing filamentous network, distinguishable in part from that found in vivo. The immunocytochemical, Western blotting and 2D-EF patterns of MF from thickened serosa indicated that the γ isoform alone is expressed in this tissue. This SM22 variant appeared before the completion of the cellular transition from MF to fully differentiated SMC. This pattern is reminiscent of bladder ontogenesis where SM22 expression in the developing bladder wall precedes that of SM myosin. Taken together these data suggest that: (i) SM22 isoforms are differently assorted in MF vs. SMC; (ii) SM22 is a SMC-lineage marker inasmuch as its expression occurs in an experimental condition characterized by a time-related cell phenotypic transition from MF to SMC, and (iii) cell conversion ability of serosal cells in the adult might take place via the reactivation of a specific “foetal” gene programme.

Journal ArticleDOI
TL;DR: Results suggest that thin-filament regulation may play a role in the activation of both unphosphorylated and phosphorylated smooth muscle myosins and suggest a possible mechanism for activation of slowly cycling unphosorylated cross bridges (i.e. latch-state) during tonic contractions of smooth muscle.
Abstract: Phosphorylation of the regulatory light chain subunit of smooth muscle myosin is sufficient, but not necessary for muscle contraction. It has been suggested that thin-filament regulation may also contribute to the regulation of contraction. A hallmark feature of regulated thin filaments, previously described for vertebrate skeletal muscle, is the capacity of strong-binding or rigor-like cross bridges to “turn-on” the actin filament. Turned-on thin filaments stimulate cross-bridge attachment even in the absence of calcium. The present study utilized an in vitro sliding-filament motility assay to test for thin-filament regulation of both unphosphorylated and phosphorylated smooth muscle myosins. Regulated thin-filaments were reconstituted from skeletal muscle actin and chicken gizzard smooth muscle tropomyosin (TmCG), and then turned-on either (1) by rigor cross bridges at low concentrations of MgATP, or (2) by adding N-ethyl-maleimide-modified skeletal subfragment S1(NEM-S1), which forms rigor-like bonds in the presence of MgATP. For control actin·TmCG filaments, force production by unphosphorylated myosin was 0.5% of that produced by thiophosphorylated myosin. The force exerted on actin·Tm filaments by both unphosphorylated and phosphorylated myosins was increased by reducing the [MgATP] to 10–100 μM MgATP (rigor-dependent activation). Force was also increased by actin·TmCG filaments that had been turned-on by NEM-S1 binding, with force production by unphosphorylated myosin increased 80-fold vs. 2.3-fold for thiophosphorylated myosin. TmCG was required for increased force production with both low MgATP and NEM-S1. Unloaded filament velocity for NEM-S1-activated thin filaments was 0.72 μm/sec with unphosphorylated myosin compared to 1.24 μm/sec with thiophosphorylated myosin. Taken together, these results suggest that thin-filament regulation may play a role in the activation of both unphosphorylated and phosphorylated smooth muscle myosins and suggest a possible mechanism for activation of slowly cycling unphosphorylated cross bridges (i.e. latch-state) during tonic contractions of smooth muscle.

Journal ArticleDOI
TL;DR: Investigation of the expression of an α-cardiac-like myosin heavy chain in rabbit tibialis anterior muscle undergoing fast-to-slow transition by continuous chronic low-frequency stimulation revealed a similar relationship between initial increase and successive decline of the MHCIα protein as seen at the␣mRNA level.
Abstract: The expression of an α-cardiac-like myosin heavy chain, MHCIα, was investigated at both the mRNA and protein levels in rabbit tibialis anterior muscle undergoing fast-to-slow transition by continuous chronic low-frequency stimulation (CLFS). According to sequence analyses of the PCR product, the MHCIα isoform was found to be identical to the α-cardiac MHC expressed in rabbit atrium. In muscles at different degrees of transformation, the upregulation of MHCIα mRNA preceded that of the MHCIβ mRNA. At more advanced stages of the transformation, MHCIα mRNA decayed while MHCIβ mRNA persisted at high levels. The expression of MHCIα, therefore, was transitory. Studies at the protein level were based on immunoblotting using a monoclonal antibody (F88 12F8,1), characterized to be specific to MHCIα in rabbit muscle. These studies revealed a similar relationship between initial increase and successive decline of the MHCIα protein as seen at the␣mRNA level. Immunohistochemistry of 30-day stimulated muscle revealed that up to 65% of the fibres expressed the MHCIα isoform in combination with other adult MHC isoforms. The most frequent patterns of coexistence were MHCIIa+MHCIα + MHCIβ (28%), MHCIα+MHCIβ (18%), and MHCIIa + MHCIα (11%). According to these combinations, the upregulation of MHCIα may be assigned as an intermediate step in the transformation of existing fibres during the␣MHCIIa → MHCIβ transition. A small fraction of fibres contained, in addition to the MHCIα + MHCIβ and MHCIIa + MHCIα combinations, developmental myosin, suggesting that MHCIα was also expressed in regenerating fibres originating from satellite cell-derived myotubes.

Journal ArticleDOI
TL;DR: The values of the rate constants that both provided the best fit through force–velocity data and correctly predicted crossbridge-dependent rate of energy output during an isometric contraction were found and used to predict the variation in rate ofEnergy liberation with shortening velocity.
Abstract: This study aimed to establish whether quantitatively accurate predictions of the rate of crossbridge-dependent energy output from shortening muscle could be made on the basis of a 2-state model of crossbridge kinetics incorporating weak coupling between mechanical cycles and ATP hydrolysis The model was based on Huxley's (1957) model but included rapid detachment, without ATP hydrolysis, of crossbridges when their strain energy increased sufficiently that crossbridge free energy exceeded that of the unbound state (Cooke et al, 1994) An expression was derived relating force to steady-state velocity in terms of the model's rate constants The values of the rate constants that both provided the best fit through force–velocity data and correctly predicted crossbridge-dependent rate of energy output during an isometric contraction were found and used to predict the variation in rate of energy liberation with shortening velocity The model predictions closely matched the estimated crossbridge energetics of frog sartorius muscle, including the decline in rate of enthalpy output at high shortening velocities Data from fast- and slow-twitch muscles of the mouse were also simulated The velocity-dependence of rate of energy liberation from fast-twitch EDL muscle was well described by the model The model overestimated crossbridge-dependent energy output from slow-twitch soleus at low shortening velocities but provided accurate predictions of energy output at high velocities In terms of this model, the distinctive energetics of fast and slow muscles cannot be explained exclusively by differences in cross-bridge detachment rate; differences in the relative rates of crossbridge attachment must also be considered to explain the different relations between energy output and shortening velocity

Journal ArticleDOI
TL;DR: A positive correlation between p94 expression level and fast glycolytic characteristics in porcine muscle is suggested.
Abstract: Members of the calpain proteinase family are present in all mammalian cells, although a novel calpain 94kDa isoform is found almost exclusively in skeletal muscle. p94 is difficult to purify from muscle and recombinant p94 autolyses rapidly when expressed in COS cells. However, in vivo the enzyme may be stabilised by interaction with titin, which has two well-characterised binding sites for p94 at the N2- and M-lines. Both these titin subdomains are subject to muscle-specific alternative splicing, which could be related to p94 expression level or stability in muscles of different fibre type. In this study, porcine longissimus dorsi (LD), trapezius (TZ) and adductor longus (AL) were characterised as fast, intermediate and slow using commercially available specific anti-human fast- and slow-myosin heavy chain mAbs and also by conventional histochemistry. p94 was quantified both in whole muscle preparations and single fibres by western blotting using an anti-p94 antiserum generated by expressing a recombinant p94 sequence as a GST fusion protein antigen.

Journal ArticleDOI
TL;DR: A novel actin binding protein has been isolated from chicken gizzard muscle and is likely to be a member of the synaptopodin family, which is a potent actin-bundling agent.
Abstract: A novel actin binding protein has been isolated from chicken gizzard muscle. When isolated, a pair of proteins with apparent molecular weights of 79 kDa and 103 kDa are obtained; both proteins have a pI near 9.3. Peptide mapping indicates that these proteins are related. Antibodies against this protein cross-reacted with proteins from other smooth muscle containing tissues as well as skeletal and heart muscle. Traces of cross-reactive material were also detected in brain and kidney tissue. The affinity of this protein for actin is ca. 1 × 106 M−1. Interestingly, this actin binding protein is a potent actin-bundling agent. A partial sequence analysis confirmed that there were no previously reported homologues in smooth muscle. However, considerable homology was found with the protein synaptopodin that is found in nervous tissue and kidney but is absent from muscle tissue. It is likely that the new protein is a member of the synaptopodin family. We call the smooth muscle actin binding protein fesselin.

Journal ArticleDOI
TL;DR: The results show that the relation between force and stiffness deviates from the direct proportionality less than previously reported, and suggest that myofilament compliance accounts for about 30% of the sarcomere compliance.
Abstract: The sarcomere stiffness was measured in single muscle fibres during the development of tetanic tension using a method insensitive to fibre intertia and viscosity. The stiffness was calculated by measuring the ratio between tension and sarcomere length during a period of fast sarcomere elongation at constant velocity. Tension changes were corrected for force truncation by the quick recovery mechanism. The results show that the relation between force and stiffness deviates from the direct proportionality less than previously reported. If the deviation is due to the presence of a linear myofilament compliance in series with the cross-bridges, our data suggest that myofilament compliance accounts for about 30% of the sarcomere compliance. This value is significantly smaller than 50–70% determined by X-ray diffraction measurements. These two different findings, however, may be reconciled by assuming that the myofilament compliance is non-linear increasing appropriately at low tension.

Journal ArticleDOI
TL;DR: Results show that the Drosophila mutant bentD mutation reduces oscillatory work in a way consistent with the proposed role of the connecting filaments in the stretch activation response of IFM.
Abstract: Projectin is a ca. 900 kDa protein that is a member of the titin protein superfamily. In skeletal muscle titins are involved in the longitudinal reinforcement of the sarcomere by connecting the Z-band to the M-line. In insect indirect flight muscle (IFM), projectin is believed to form the connecting filaments that link the Z-band to the thick filaments and is responsible for the high relaxed stiffness found in this muscle type. The Drosophila mutant bentD (btD) has been shown to have a breakpoint close to the carboxy-terminal kinase domain of the projectin sequence. Homozygotes for btD are embryonic lethal but heterozygotes (btD/+) are viable. Here we show that btD/+ flies have normal flight ability and a slightly elevated wing beat frequency (btD/+ 223+/-13 Hz; +/+ 203+/-5 Hz, mean +/- SD; P < 0.01). Electron microscopy of btD/+ IFM show normal ultrastructure but skinned fiber mechanics show reduced stretch activation and oscillatory work. Although btD/+ IFM power output was at wild-type levels, maximum power was achieved at a higher frequency of applied length perturbation (btD/+ 151+/-6 Hz; +/+ 102+/-14 Hz; P < 0.01). Results were interpreted in the context of a viscoelastic model of the sarcomere and indicate altered cross-bridge kinetics of the power-producing step. These results show that the btD mutation reduces oscillatory work in a way consistent with the proposed role of the connecting filaments in the stretch activation response of IFM.

Journal ArticleDOI
TL;DR: In this paper, the expression of MHCIα, an α-cardiac-like myosin heavy chain isoform, was studied in extensor digitorum longus (EDL) and tibialis anterior (TA) rat muscles undergoing fast-to-slow transition by chronic low-frequency stimulation (CLFS), a condition inducing a transient upregulation of mihIα in rabbit muscle.
Abstract: The expression of MHCIα, an α-cardiac-like myosin heavy chain isoform, was studied in extensor digitorum longus (EDL) and tibialis anterior (TA) rat muscles undergoing fast-to-slow transition by chronic low-frequency stimulation (CLFS), a condition inducing a transient upregulation of MHCIα in rabbit muscle. In order to enhance the transformation process, CLFS was applied to hypothyroid rats. mRNA analyses were performed by RT-PCR, and studies at the protein level by immunoblotting and immunohistochemistry, using the F88 antibody (F88 12F8,1) demonstrated in the accompanying paper to be specific for MHCIα. In total RNA preparations from slow- and fast-twitch muscles, MHCIα mRNA was present at minute levels, at least three orders of magnitude lower than in cardiac atrium. As verified immunohistochemically, MHCIα is present only in intrafusal fibres of rat muscle. Moreover, MHCIα is not expressed in extrafusal fibres and, contrary to the rabbit, was not upregulated at both the mRNA and protein levels by CLFS. These results support our notion of species-specific responses to CLFS. Another antibody reported to be specific to MHCIα, BA-G5, was also investigated by immunoblot and immunohistochemical analyses. Its specificity could not be validated for skeletal muscles of the rat. BG-A5 was shown to cross-react with MHCIIb and MHCIβ. These results question an upregulation of MHCIα in transforming rat muscles as reported in studies based on the use of this antibody.

Journal ArticleDOI
TL;DR: Evidence is reported for an additional role for cardiac troponin C (cTnC) in the association of force-generating myosin cross-bridges and to test this hypothesis actomyosin ATPase (NADH fluorescence change) and isometric force were measured in skinned cardiac preparations.
Abstract: It is well known that cardiac troponin C (cTnC) regulates the association of force-generating myosin cross-bridges. We report here evidence for an additional role for cTnC. This hypothesis states that Car2+ binds more strongly to cTnC when force-generating myosin cross-bridges are attached to actin and that removal of this bound Ca2+ accelerates the dissociation of force-generating myosin cross-bridges. Intact Fura-2-loaded rat papillary muscles and skinned (permeabilized) ventricular preparations were used. The preparations were mounted in the Guth Muscle Research System which is capable of measuring simultaneously fluorescence and force in response to length perturbations. All mechanical perturbations of muscle length (isotonic shortening, quick stretches and releases, and length vibrations) which cause dissociation of force-generating myosin cross-bridges during a twitch resulted in Ca2+ being released from troponin as judged from changes in the Ca2+ transients (Fura-2 (340/380) fluorescence ratio). Thus dissociation of force-generating myosin cross-bridges cause Ca2+ to be released from cTnC. Conversely, it would be expected that removal of strongly bound Ca2+ from cTnC would result in an increase in the rate of dissociation of force-generating myosin cross-bridges. To test this hypothesis actomyosin ATPase (NADH fluorescence change) and isometric force were measured in skinned cardiac preparations. The ratio of the ATPase/Force is proportional to the rate constant (gapp) for the dissociation of force-generating myosin cross-bridges. The data showed that decreasing the amount of Ca2+ bound to cTnC in skinned cardiac fibers caused an increase in the ratio of ATPase/Force, the rate of dissociation (gapp) of force-generating myosin cross-bridges.

Journal ArticleDOI
TL;DR: The juxtaposed fiber method more accurately approximates parvalbumin diffusion in undisturbed myoplasm because no artificial solutions were used and, upon fiber-to-fiber contact, a potentially confounding oil barrier at the interface rapidly disperses.
Abstract: The concentrations and diffusivity of two isoforms of parvalbumin, IVa and IVb, were measured using quantitative SDS PAGE in single fibers from semitendinosus muscles of the frog Rana temporaria. The concentrations of IVa and IVb were 2.9 ± 0.3 (SEM) and 4.5 ± 0.5 gl−1 total fiber volume, respectively. The total concentration of parvalbumin (7.4 ± 0.8 gl−1 total fiber) corresponds to a cytosolic concentration of 0.9 ± 0.1 mmoll−1 myoplasmic water. Estimates for the transverse and longitudinal diffusion coefficients for parvalbumin at 4°C were obtained in two ways: (1) by diffusion of parvalbumin out of skinned fibers into droplets of relaxing solution, and (2) by diffusion of parvalbumin between two juxtaposed skinned fibers under oil. The transverse diffusion coefficient obtained using the droplet method was significantly lower than that obtained using juxtaposed fibers, but the longitudinal diffusion coefficients obtained from both methods were similar. The juxtaposed fiber method more accurately approximates parvalbumin diffusion in undisturbed myoplasm because no artificial solutions were used and, upon fiber-to-fiber contact, a potentially confounding oil barrier at the interface rapidly disperses. The juxtaposed fiber method yielded values for transverse (4.27 ± 0.87 × 10−7 cm2 s−1) and longitudinal (3.20 ± 0.74 × 10−7 cm2 s−1) diffusion coefficients that were not significantly different, suggesting that diffusion of parvalbumin in myoplasm is essentially isotropic. The average diffusion coefficient of frog parvalbumin in myoplasm (3.74 ± 0.81 × 10−7 cm2 s−1; 4°C) is approximately a third of that estimated for frog parvalbumin diffusing in bulk water into and out of 3% agarose cylinders (10.6 × 10−7 cm2 s−1; 4°C). The reduced translational mobility of parvalbumin in myoplasm reflects an elevated effective viscosity due to tortuosity and viscous drag imposed by the fixed proteins of the cytomatrix and the numerous diffusible particles of the cytosol.

Journal ArticleDOI
TL;DR: At early stages of muscle development, skeletal muscles contract and relax slowly, regardless of whether they are destined to become fast- or slow-twitch, which leads to a paradoxical situation whereby at early stagesof muscle development the fast muscles have contraction dynamics of slow muscles but the slow muscles have activation profiles more characteristic of fast muscles.
Abstract: At early stages of muscle development, skeletal muscles contract and relax slowly, regardless of whether they are destined to become fast- or slow-twitch. In this study, we have characterised the activation profiles of developing fast- and slow-twitch muscles from a precocial species, the sheep, to determine if the activation profiles of the muscles are characteristically slow when both the fast- and slow-twitch muscles have slow isometric contraction profiles. Single skinned muscle fibres from the fast-twitch flexor digitorum longus (FDL) and slow-twitch soleus muscles from fetal (gestational ages 70, 90, 120 and 140 days; term 147 days) and neonatal (8 weeks old) sheep were used to determine the isometric force-pCa (pCa = -log10[Ca2+]) and force-pSr relations during development. Fast-twitch mammalian muscles generally have a greatly different sensitivity to Ca2+ and Sr2+ whereas slow-twitch muscles have a similar sensitivity to these divalent cations. At all ages studied, the force-pCa and force-pSr relations of the FDL muscle were widely separated. The mean separation of the mid-point of the curves (pCa50-pSr50) was approximately 1.1. This is typical of adult fast-twitch muscle. The force-pCa and force-pSr curves for soleus muscle were also widely separated at 70 and 90 days gestation (pCa50-pSr50 approximately 0.75); between 90 days and 140 days this separation decreased significantly to approximately 0.2. This leads to a paradoxical situation whereby at early stages of muscle development the fast muscles have contraction dynamics of slow muscles but the slow muscles have activation profiles more characteristic of fast muscles. The time course for development of the FDL and soleus is different, based on sarcomere structure with the soleus muscle developing clearly defined sarcomere structure earlier in gestation than the FDL. At 70 days gestation the FDL muscle had no clearly defined sarcomeres. Force (N cm-2) increased almost linearly between 70 and 140 days gestation in both muscle types and there was no difference between the Ca(2+)- and Sr(2+)-activated force throughout development.

Journal ArticleDOI
TL;DR: The results show that the rate of endothermic force generation in cardiac fibres is clearly different from that of either fast-twitch or slow-twitch mammalian skeletal muscle fibres; implication of such fibre type-specific differences is discussed in relation to the difficulty in identifying the biochemical step underlying endotherm cross-bridge force generation.
Abstract: Isometric tension responses to rapid temperature jumps (T-jumps) of 2–6°C were examined in skinned muscle fibre bundles isolated from papillary muscles of the rat heart. T-jumps were induced by an infra-red laser pulse (wave length 1.32 μm, pulse duration 0.2 ms) obtained from a Nd-YAG laser, which heated the fibres and bathing buffer solution in a 50 μl trough; the increased temperature by laser pulse was clamped at the high temperature by a Peltier system (see Ranatunga, 1996). In maximally Ca2+-activated (pCa ca. 4.5) fibres, the relationship between tension and temperature was non-linear, the increase of active tension with temperature being more pronounced at lower temperatures (below ca. 20°C). A T-jump at any temperature (range 3–35°C) induced an initial step decrease of tension of variable amplitude (Phase 1), probably due to thermal expansion, and it was followed by a tension transient which resulted in a net rise of tension above the pre-T-jump level. The rate of net rise of tension (Phase 2b or endothermic force generation) was 7–10/s at ca. 12°C and its Q10 was 6.3 (below 25°C). In cases where the step decrease of tension in Phase 1 was prominent, an initial quick tension recovery phase (Phase 2a, 70–100/s at 12°C) that did not contribute to a rise of tension above the pre-T-jump level, was also seen. This phase (Phase 2a) appeared to be similar to the quick tension recovery induced by a small length release and its rate increased with temperature with a Q10 of 1.8. In some cases where Phase 2a was present, a slower tension rise (Phase 3) was seen; its rate (ca. 5/s) was temperature-insensitive. The results show that the rate of endothermic force generation in cardiac fibres is clearly different from that of either fast-twitch or slow-twitch mammalian skeletal muscle fibres; implication of such fibre type-specific differences is discussed in relation to the difficulty in identifying the biochemical step underlying endothermic cross-bridge force generation.