Comprehensive Physiology Wiley Online Library

Modulation of Skeletal Muscle Contraction by Myosin Phosphorylation

Full Article on Wiley Online Library



ABSTRACT

The striated muscle sarcomere is a highly organized and complex enzymatic and structural organelle. Evolutionary pressures have played a vital role in determining the structure‐function relationship of each protein within the sarcomere. A key part of this multimeric assembly is the light chain‐binding domain (LCBD) of the myosin II motor molecule. This elongated “beam” functions as a biological lever, amplifying small interdomain movements within the myosin head into piconewton forces and nanometer displacements against the thin filament during the cross‐bridge cycle. The LCBD contains two subunits known as the essential and regulatory myosin light chains (ELC and RLC, respectively). Isoformic differences in these respective species provide molecular diversity and, in addition, sites for phosphorylation of serine residues, a highly conserved feature of striated muscle systems. Work on permeabilized skeletal fibers and thick filament systems shows that the skeletal myosin light chain kinase catalyzed phosphorylation of the RLC alters the “interacting head motif” of myosin motor heads on the thick filament surface, with myriad consequences for muscle biology. At rest, structure‐function changes may upregulate actomyosin ATPase activity of phosphorylated cross‐bridges. During activation, these same changes may increase the Ca2+ sensitivity of force development to enhance force, work, and power output, outcomes known as “potentiation.” Thus, although other mechanisms may contribute, RLC phosphorylation may represent a form of thick filament activation that provides a “molecular memory” of contraction. The clinical significance of these RLC phosphorylation mediated alterations to contractile performance of various striated muscle systems are just beginning to be understood. © 2017 American Physiological Society. Compr Physiol 7:171‐212, 2017.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1. Paradigm dependence for force potentiation of mouse fast muscle. Example of paradigm and contraction type dependence for potentiation of mouse EDL muscle (in vitro 25°C). (A) Original force records showing isometric (solid line) and concentric (dashed line) twitch forces obtained before (left) and after (right) a potentiating stimulus in WT (top) and skMLCK−/− (bottom) muscles. For both genotypes, the potentiating atimulus consisted of four brief volleys of 100 Hz in a 10‐s time window with the muscle at optimal length; peak tetanic force declined by ∼8% from the first to the last tetanus in both genotypes. In WT muscles, this protocol typically elevates RLC phosphorylation from ∼0.10 to 0.55 mol phos per mol RLC (52,127,128,129); in skMLCK−/− muscles this protocol does not change RLC phosphorylation form resting values (127,129). Isometric and concentric twitch forces were increased by ∼20% and 40%, respectively, in the WT muscles. In the skMLCK−/− muscle, isometric twitch force was not increased although concentric twitch force was increased ∼10%. (B) Original force records showing gradual increase in both isometric (solid line) and concentric (dashed line) twitch force during 15 s of 15‐Hz stimulation in WT (top) and skMLCK−/− (bottom) muscles. The first and last twitches of each experiment are shown at right. In the WT muscle, isometric and concentric twitch forces were as increased to ∼20% and 50% of initial levels, respectively. In the skMLCK−/− muscle, isometric twitch force was not increased but concentric twitch force was increased by ∼15%. In this protocol, RLC phosphorylation is typically elevated from ∼0.10 to 0.50 mol phos per mol RLC while in skMLCK−/− muscles this protocol does not change RLC phosphorylation from resting values (unpublished observation). In both A and B muscles, shortened at 0.45 Vmax to obtain concentric twitches; both isometric and conctric twitches were elicted near Lo. In each panel, post CS twitch amplitude is normalized to initial unpotentiated value.
Figure 2. Figure 2. Cartoon depicting structure of sarcomeric myosin. Schematic representation of the myosin II (sarcomeric) protein of vertebrate striated muscle. Cartoon shows how the hexamer molecule is constructed of two MyHC monomers (∼240 kDa), which dimerize along the length of the elongated tail domain (shown in gray) toward the COOH terminus but which separate toward the NH3 terminus to form twin globular head domains. In addition, each head contains a dumbbell‐shaped regulatory (RLC) and ELC (∼20 kDa each) protein subunit (shown in purple and yellow, respectively) bound to the LCBD. The approximate location of the nucleotide binding site on the head is shown (black dot), while the actin binding site is formed by the cleft between the upper and lower subdomains of the head (shown in blue). Digestion by papain or trypsin cleaves the MyHC into heavy (HMM) and light (LMM) meromyosin fragments where the HMM consists of both the S‐1 (blue) domain and the N‐terminus portion of the S‐2 (gray) domain. Note that individual domains are not shown to scale and relative positions are highly idealized. Separation of individual monomers near the N‐terminus of the S‐2 region may allow more flexibility of S‐1 head position than shown. Text at bottom depicts how S‐1 and S‐2 domains constitute LMM and HMM components, respectively, of the MyHC. Myosin dimensions: tail, ∼150 nm; head, ∼16.5 nm in length. For atomic details of myosin S‐1 or RLC structure, see Refs. 6,27,152,153,212,311,312,313,314,404.
Figure 3. Figure 3. Regulatory cascade for skeletal myosin RLC phosphorylation. Schematic pathway for the calcium (Ca2+) and CaM‐derived activation of skMLCK and myosin RLC phosphorylation in vertebrate striated muscle. During activity, rapid increases in free myoplasmic Ca2+ (top left) drives a reaction cascade that parallels Ca2+ binding to TnC. Saturation of Ca2+‐binding sites on CaM produces a four Ca2+‐CaM pcomplex that is capable of binding to the inactive form of skMLCK. In skeletal muscle, formation of and diffusion of the four Ca2+‐CaM complex is very rapid and not rate limiting. The slower rate of formation of the four Ca2+‐CaM‐skMLCK complex produces an active pool of skMLCK enzyme that catalyzes a phosphotransferase reaction in which the high energy nucleotide molecule ATP donates a phosphoryl group to a serine reside on the RLC. The reverse reaction is catalyzed by MLCP and is unregulated. The thickness of the vertical arrows depicts the 40‐ to 50‐fold greater catalytic rate of skMLCKa (down) versus MLCP (up). During inactivity, Ca2+ sequestration (bottom left) causes the active holoenzyme and Ca2+‐CaM complexes to dissolve. Note that when [Ca2+]i declines Ca2+ first dissociates from the skMLCK bound fraction of Ca2+‐CaM pool, thus allowing CaM to dissociate from skMLCK. The thickness of the horizontal arrows depicts the much greater rate of skMLCK activation (right) versus inactivation (left) during activity and inactivity, respectively. Differences in the rates of skMLCKi → skMLCKa and skMLCKa catalytic activity versus skMLCKi ← skMLCKa and MLCP activity account for the “molecular memomry” of myosin phosphorylation. Note that differences in expression of skMLCK to MLCP in fast ([skMLCK] > [MLCP]) and slow ([skMLCK] < [MLCP]) skeletal muscle fibers also contribute to regulation of RLC phosphorylation (269,337). Regulatory scheme based on Refs. 192,349,352.
Figure 4. Figure 4. Kinetics of force, myoplasmic Ca2+, skMLCK activity and RLC phosphorylation. Kinetic model describing the time course of change in force (black line), peak myoplasmic Ca2+ concentration (solid red line), skMLCK activity (blue line), MLCP active (green line), and RLC phosphorylation (purple line) as a result of 2 s of high‐frequency (100‐150 Hz) tetanic stimulation. Also shown is the proposed change in resting myoplasmic Ca2+ (dotted red line) as a result of stimulation. Due to the rapid time course for RLC phosphorylation during and the relatively slow time course for RLC dephosphorylation following stimulation, RLC phosphorylation is proposed to act as a “molecular memory” of muscle activity by increasing the Ca2+ sensitivity of the contractile element of fast twitch skeletal muscle. During lower‐frequency stimulation, skMLCK activity and RLC phosphorylation occur on a slower time scale but occur rapidly enough to cause staircase potentiation (not shown). The contribution of stimulus‐induced changes to resting myoplasmic Ca2+ is illustrated as relatively minor for PTP; this mechanism may be more prevalent during staircase potentiation. Figure redrawn, from Ref. 349 and modified, according to Ref. 337. Note that skMLCK catalytic activity is slower than is formation of the active skMLCK holoenzyme by Ca2+‐CaM binding (not shown). Note that all parameters are expressed relative to the maximal response for the time window depicted, not absolute levels.
Figure 5. Figure 5. Influence of RLC phosphorylation on arthropod thick filaments. (A) Three‐dimensional reconstruction of tarantula thick filament segment showing four helical tracks of myosin S‐1 heads. One myosin interacting‐heads motif is highlighted in yellow. Densities of interaction between heads are labelled as “a,” “b,” and “c.” In this configuration, individual myosin heads are less likely to be able to interact with actin, turnover ATP, and/or produce force. Disruption of this motif by phosphorylation may represent a type of thick filament activation that promotes force and ATPase activity. Vertical calibration bar represents 14.5 nm. Reprinted from Alamo et al. 2008 with permission (8). (B) Cartoon showing how progressive phosphorylation of the RLC (i.e., nonphosphorylated‐monophosphorylated‐diphosphorylated) influences the distribution of “free” and “blocked” myosin heads. In rested state (a), constitutive monophosphorylation of the free head RLC (yellow dot) is sufficient to produce a population of free heads that “sway” (double arrow) between close contact with their unphosphorylated and “blocked” dimeric partner head and weak contact with the thin filament. Because of their mobility, these heads bind readily to actin during rapid Ca2+ activation of the contractile apparatus (i.e., during a twitch) (b). In (c), activation of skMLCK and diphosphorylation of free heads or monophosphorylation of blocked heads increases the fraction of heads able to bind to actin. In this model, the greater the phosphorylation of the RLC by skMLCK, the greater the potentiation, either by biasing the probability that a free head will bind to actin or by allowing the skMLCK enzyme to access and phosphorylate the RLC of the blocked head. Although withdrawal of Ca2+ by SERCA pumping and reduction of [Ca2+]myo causes relaxation (d) the slow dephosphorylation of the RLC produces a “molecular memory” at the level of the RLC that will produce PTP when the muscle is reactivated (e). Note that the extent to which this scheme applies to mammalian thick filaments is not known. Colour Key: free and swaying heads are shown in light and dark blue, respectively; blocked heads are shown in green; phosphorylation sites on RLC are shown in yellow and red (Ser35 and Ser45, respectively). Monophosphorylation is produced by phosphorylation of either Ser35 or Ser45, while diphosphorylation requires both sites to be phosphorylated. Reprinted from Brito et al. (2015) with permission (42). The reader is referred to original papers for more details of the interacting head motif or the molecular mechanism of this thick filament activation by phosphorylation.
Figure 6. Figure 6. Influence of RLC phosphorylation on arthropod thick filaments. (A) Three‐dimensional reconstruction of tarantula thick filament segment showing four helical tracks of myosin S‐1 heads. One myosin interacting‐heads motif is highlighted in yellow. Densities of interaction between heads are labelled as “a,” “b,” and “c.” In this configuration, individual myosin heads are less likely to be able to interact with actin, turnover ATP, and/or produce force. Disruption of this motif by phosphorylation may represent a type of thick filament activation that promotes force and ATPase activity. Vertical calibration bar represents 14.5 nm. Reprinted from Alamo et al. 2008 with permission (8). (B) Cartoon showing how progressive phosphorylation of the RLC (i.e., nonphosphorylated‐monophosphorylated‐diphosphorylated) influences the distribution of “free” and “blocked” myosin heads. In rested state (a), constitutive monophosphorylation of the free head RLC (yellow dot) is sufficient to produce a population of free heads that “sway” (double arrow) between close contact with their unphosphorylated and “blocked” dimeric partner head and weak contact with the thin filament. Because of their mobility, these heads bind readily to actin during rapid Ca2+ activation of the contractile apparatus (i.e., during a twitch) (b). In (c), activation of skMLCK and diphosphorylation of free heads or monophosphorylation of blocked heads increases the fraction of heads able to bind to actin. In this model, the greater the phosphorylation of the RLC by skMLCK, the greater the potentiation, either by biasing the probability that a free head will bind to actin or by allowing the skMLCK enzyme to access and phosphorylate the RLC of the blocked head. Although withdrawal of Ca2+ by SERCA pumping and reduction of [Ca2+]myo causes relaxation (d) the slow dephosphorylation of the RLC produces a “molecular memory” at the level of the RLC that will produce PTP when the muscle is reactivated (e). Note that the extent to which this scheme applies to mammalian thick filaments is not known. Colour Key: free and swaying heads are shown in light and dark blue, respectively; blocked heads are shown in green; phosphorylation sites on RLC are shown in yellow and red (Ser35 and Ser45, respectively). Monophosphorylation is produced by phosphorylation of either Ser35 or Ser45, while diphosphorylation requires both sites to be phosphorylated. Reprinted from Brito et al. (2015) with permission (42). The reader is referred to original papers for more details of the interacting head motif or the molecular mechanism of this thick filament activation by phosphorylation.
Figure 7. Figure 7. Theoretical relationship between Ca2+ sensitivity and isometric twitch force. Theoretical relationship between skMLCK catalyzed RLC phosphorylation mediated increases in Ca2+ sensitivity and isometric twitch force in permeabilized skeletal muscle fibers and intact rodent skeletal muscle, respectively. Main graph: representative force‐pCa relationship of permeabilized fibers where red line is without RLC phosphorylation and blue line is with RLC phosphorylation. In the highly phosphorylated state (>0.70 mol phos mol RLC), the force‐pCa relationship is shifted to the left relative to unphosphrylated state (253,354,356) without effect on maximal Ca2+ activated force (Fmax). Top left panel insert: example traces from EDL muscle (in vitro, 25°C) showing how presence of skMLCK enzyme and stimulation induced elevations in RLC phosphorylation is associated with PTP of isometric twitch force (before stimulation, thick line; after stimulation, thin line). Bottom right panel insert: example traces from skMLCK−/− mouse EDL muscle (in vitro, 25°C) showing how the absence of skMLCK enzyme and absence of stimulation induced RLC phosphorylation is associated with posttetanic depression of force rather than PTP (before stimulation, thick line; after stimulation, thin line). Note that changes to twitch time course (i.e., decreased half‐relaxation time) are roughly similar for both genotypes and thus apparently independent of RLC phosphorylation and PTP per se. Repetitive low frequency stimulation leading to staircase potentiation may produce increases in isometric twitch force amplitude that approach that displayed by WT muscles, indicating the presence of a RLC phosphorylation‐independent mechanism.
Figure 8. Figure 8. Summary of RLC phosphorylation effects on permeabilized fibers. Flow chart depicting theoretical cascade of events during Ca2+ activation of permeabilized skeletal muscle fibers. Distinction is made between suboptimal and optimal Ca2+ activation. Cascade based on results from mammalian muscle fibers at low temperatures. The influence of RLC phosphorylation on αFS is assumed to be due to an increase in fapp without change to gapp. Other mechanisms such as temperature, ionic strength, or Ca2+ binding to cation sites on the RLC are not considered in this model. Abbreviations: fraction of cycling cross‐bridges in force‐generating state (αFS); AM ATPase; loaded shortening velocity (VL); rate constant for steady‐state tension development (ktr or kREDEV); rate of relaxation from steady state (RFR); unloaded shortening velocity as determined by the slack test (Vu).
Figure 9. Figure 9. Influence of potentiation on force‐frequency relation of mouse skeletal muscle. Illustration of the influence of PTP on the force frequency responses of WT mouse EDL muscle (in vitro, 25°C) when measured at the peak of potentiation typically observed 10 to 20 s after a tetanic CS. (A) Isometric force‐frequency responses at optimal length in the unpotentiated (red) and potentiated (blue) state. Unpotentiated and potentiated forces (F) are normalized to peak force (Fo) in the unpotentiated state. The relatively small portion of the activation frequency domain for isometric PTP relative to the frequency stimulus range for fast muscle in vivo is indicated at top. Note that when isometric contractions are very brief (i.e., <100 ms in duration) the relative effect of PTP on the peak force that is attained is much greater. (B) Concentric force‐frequency responses in the unpotentiated (red) and potentiated (blue) state. To construct this model a shortening velocity of 0.35 Vmax was used for both unpotentiated and potentiated conditions; faster and slower speeds of shortening will shift the potentiated curve upward and downward, respectively, relative to the unpoentiateed curve (based on Refs. 128 and 382). Unpotentiated and potentiated forces (F) are normalized to peak force (Fo) in the unpotentiated state. The congruence between the activation frequency domains for concentric PTP relative to the frequency stimulus range for fast muscle in vivo is indicated at top. Comparison of panels A and B reveals that the influence of PTP is much greater for concentric than for isometric contractions both in terms of the magnitude of increase and the stimulus frequency range over which potentiation is evident.
Figure 10. Figure 10. Influence of potentiation on force‐velocity relation of mouse skeletal muscle. Illustration of the influence of PTP on the force‐velocity responses of WT mouse EDL muscle (in vitro, 25°C) when measured at the peak of potentiation typically observed 10 s after a tetanic CS. Note that, because the influence of potentiation on concentric force is greatest at moderate shortening speeds and smallest at slow and fast shortening speeds, the potentiated force‐velocity curve is less concave than the unpotentiated force‐velocity curve. Curves represent situations in which muscles are stimulated at high frequency to produce maximal responses in both unpotentiated and potentiated state. Vertical arrow denotes effect on potentiated curve relative to unpotentiated curve if stimulus frequency is increased or decreased (down and up, respectively). Curves are hand drawn.
Figure 11. Figure 11. Effect of genotype on concentric force potentiation of mouse muscle. Hypothetical model for concentric force potentiation of EDL muscles from WT and skMLCK−/− mice (in vitro, 25°). (A) The effect of mouse genotype on the time dependence of concentric force potentiation. Responses shown are for moderate frequency contractions (i.e., 45 Hz).and a moderate speed of shortening (i.e., 0.35 Vmax). Note that although WT muscle (blue line) responses are always greater, skMLCK−/− muscle (red line) also displays a significant potentiation that cannot be attributed to RLC phosphorylation. The vertical arrows represent the difference in potentiation between genotypes hypothesized to be due to RLC phosphorylation (i.e., WT – skMLCK−/− responses). Use of lower or higher‐frequency stimulation will shift both curves up or down the y‐axis, respectively. Potentiation calculated as post‐CS response at select times divided by pre‐CS value and should be considered represent situation for either peak or mean force. Based on unpublished data. (B) The effect of mouse genotype on the stimulus frequency dependence of concentric force potentiation. Values represent the peak of potentiation as captured 10 to 20 s after a tetanic CS. Note that although WT responses are always greater (blue line), it is apparent that skMLCK−/− muscles also display a potentiation at all frequencies that cannot be attributed to RLC phosphorylation. The vertical arrows represent the difference in potentiation hypothized to be solely due to RLC phosphorylation (i.e., WT – skMLCK−/− responses). Curves are based on a shortening speed of 0.35 Vmax and assume maximal potentiation as determined 10 to 20 s after a tetanic CS. Potentiation calculated as post‐CS response divided by pre‐CS value and represent either peak or mean force or work/power. Based on Ref. 126 and unpublished data.
Figure 12. Figure 12. Summary of RLC phosphorylation influence on intact rodent muscle. Flow chart depicting theoretical cascade of events leading to myosin RLC phosphorylation and PTP or staircase potentiation in fast twitch rodent skeletal muscle. The proposed influence of RLC phosphorylation on muscle force, work, and power is illustrated for twitch and tetanic (Pt and Po, respectively) as well as during isometric and dynamic contractions. The possible influence of estrogen on skMLCK activity and upregulating RLC phosphorylation is included. Potentiation due to alternate mechanisms is not considered (dotted lines). Abbreviatons: force‐time integral (FTI); rate of isometric force development (RFD); rate of isometric force relaxation (RFR); isometric twitch (Pt), submaximal force (<Po); peak isometric force (Po); and high‐energy phosphate turnover (HEPT). Economy refers to ratio between HEPT and FTI during isometric contractions while efficiency refers to ratio between HEPT and work performed during concentric contractions. Displacement refers to the magnitude of shortening during isotonic type contractions. Vmax refers to maximum shortening velocity as determined by extrapolation to zero load while Vu refers to unloaded shortening velocity as detrmined by the slack test. The proposed fiber type dependence of myosin RLC phosphorylation mediated potentiation demonstrated in skMLCK overexprssors (323) is recognized by the bifurcation of the flow chart and the dotted arrow pointing toward type IIa and IIa fibers. Dotted lines depict speculative link between estrogen signaling and skMLCK activity (66). Dashed lines depict speculative link between skMLCK catalyzed RLC phosphorylation, the SRX and resting muscle thermogenesis (70,247).
Figure 13. Figure 13. Summary of RLC phosphorylation influence on human skeletal muscle. Flow chart depicting theoretical cascade of events leading to myosin RLC phosphorylation postactivation potentiation (PAP) (staircase or PTP) in human skeletal muscle. A theoretical distinction is made between evoked and voluntary condition contractions as well as between isometric and ballistic contraction types. Studies showing that the magnitude of PAP of evoked contractions tend to be smaller in magnitude than for voluntary contractions is recognized and accounted for by the presence of signals arising in the periphery that act as a negative feedback signal to upper centers to dampen PAP (dashed lines). Dotted lines depict speculative link between reductions in estrogen levels and reductions in phosphorylation of fast and slow RLC isoforms noted in older women relative to either young women or age matched males (259). The potential influence of resting RLC phosphorylation levels on muscle thermogenesis, via redistribution of cross‐bridges from SRX to CRX, is included (dotted lines). Ballistic contractions refer to isokinetic contractions at fast (>180°/s) or slow (<180°/s) speeds of shortening. Abbreviations: evoked twitch (Pt); evoked submaximal force (Pt‐Po); evoked peak force (Po); rate of isometric force development (RFD); maximal voluntary contraction (MVC); submaximal voluntary contraction (SVC); and motor unit discharge rate (MUDR). Due to technical limitations, the influence of PAP on MUDR during ballistic contractions has not been determined. Neither metabolic factors inhibiting or alternate mechanisms for PAP are considered in this flowchart (although they may well exist).


Figure 1. Paradigm dependence for force potentiation of mouse fast muscle. Example of paradigm and contraction type dependence for potentiation of mouse EDL muscle (in vitro 25°C). (A) Original force records showing isometric (solid line) and concentric (dashed line) twitch forces obtained before (left) and after (right) a potentiating stimulus in WT (top) and skMLCK−/− (bottom) muscles. For both genotypes, the potentiating atimulus consisted of four brief volleys of 100 Hz in a 10‐s time window with the muscle at optimal length; peak tetanic force declined by ∼8% from the first to the last tetanus in both genotypes. In WT muscles, this protocol typically elevates RLC phosphorylation from ∼0.10 to 0.55 mol phos per mol RLC (52,127,128,129); in skMLCK−/− muscles this protocol does not change RLC phosphorylation form resting values (127,129). Isometric and concentric twitch forces were increased by ∼20% and 40%, respectively, in the WT muscles. In the skMLCK−/− muscle, isometric twitch force was not increased although concentric twitch force was increased ∼10%. (B) Original force records showing gradual increase in both isometric (solid line) and concentric (dashed line) twitch force during 15 s of 15‐Hz stimulation in WT (top) and skMLCK−/− (bottom) muscles. The first and last twitches of each experiment are shown at right. In the WT muscle, isometric and concentric twitch forces were as increased to ∼20% and 50% of initial levels, respectively. In the skMLCK−/− muscle, isometric twitch force was not increased but concentric twitch force was increased by ∼15%. In this protocol, RLC phosphorylation is typically elevated from ∼0.10 to 0.50 mol phos per mol RLC while in skMLCK−/− muscles this protocol does not change RLC phosphorylation from resting values (unpublished observation). In both A and B muscles, shortened at 0.45 Vmax to obtain concentric twitches; both isometric and conctric twitches were elicted near Lo. In each panel, post CS twitch amplitude is normalized to initial unpotentiated value.


Figure 2. Cartoon depicting structure of sarcomeric myosin. Schematic representation of the myosin II (sarcomeric) protein of vertebrate striated muscle. Cartoon shows how the hexamer molecule is constructed of two MyHC monomers (∼240 kDa), which dimerize along the length of the elongated tail domain (shown in gray) toward the COOH terminus but which separate toward the NH3 terminus to form twin globular head domains. In addition, each head contains a dumbbell‐shaped regulatory (RLC) and ELC (∼20 kDa each) protein subunit (shown in purple and yellow, respectively) bound to the LCBD. The approximate location of the nucleotide binding site on the head is shown (black dot), while the actin binding site is formed by the cleft between the upper and lower subdomains of the head (shown in blue). Digestion by papain or trypsin cleaves the MyHC into heavy (HMM) and light (LMM) meromyosin fragments where the HMM consists of both the S‐1 (blue) domain and the N‐terminus portion of the S‐2 (gray) domain. Note that individual domains are not shown to scale and relative positions are highly idealized. Separation of individual monomers near the N‐terminus of the S‐2 region may allow more flexibility of S‐1 head position than shown. Text at bottom depicts how S‐1 and S‐2 domains constitute LMM and HMM components, respectively, of the MyHC. Myosin dimensions: tail, ∼150 nm; head, ∼16.5 nm in length. For atomic details of myosin S‐1 or RLC structure, see Refs. 6,27,152,153,212,311,312,313,314,404.


Figure 3. Regulatory cascade for skeletal myosin RLC phosphorylation. Schematic pathway for the calcium (Ca2+) and CaM‐derived activation of skMLCK and myosin RLC phosphorylation in vertebrate striated muscle. During activity, rapid increases in free myoplasmic Ca2+ (top left) drives a reaction cascade that parallels Ca2+ binding to TnC. Saturation of Ca2+‐binding sites on CaM produces a four Ca2+‐CaM pcomplex that is capable of binding to the inactive form of skMLCK. In skeletal muscle, formation of and diffusion of the four Ca2+‐CaM complex is very rapid and not rate limiting. The slower rate of formation of the four Ca2+‐CaM‐skMLCK complex produces an active pool of skMLCK enzyme that catalyzes a phosphotransferase reaction in which the high energy nucleotide molecule ATP donates a phosphoryl group to a serine reside on the RLC. The reverse reaction is catalyzed by MLCP and is unregulated. The thickness of the vertical arrows depicts the 40‐ to 50‐fold greater catalytic rate of skMLCKa (down) versus MLCP (up). During inactivity, Ca2+ sequestration (bottom left) causes the active holoenzyme and Ca2+‐CaM complexes to dissolve. Note that when [Ca2+]i declines Ca2+ first dissociates from the skMLCK bound fraction of Ca2+‐CaM pool, thus allowing CaM to dissociate from skMLCK. The thickness of the horizontal arrows depicts the much greater rate of skMLCK activation (right) versus inactivation (left) during activity and inactivity, respectively. Differences in the rates of skMLCKi → skMLCKa and skMLCKa catalytic activity versus skMLCKi ← skMLCKa and MLCP activity account for the “molecular memomry” of myosin phosphorylation. Note that differences in expression of skMLCK to MLCP in fast ([skMLCK] > [MLCP]) and slow ([skMLCK] < [MLCP]) skeletal muscle fibers also contribute to regulation of RLC phosphorylation (269,337). Regulatory scheme based on Refs. 192,349,352.


Figure 4. Kinetics of force, myoplasmic Ca2+, skMLCK activity and RLC phosphorylation. Kinetic model describing the time course of change in force (black line), peak myoplasmic Ca2+ concentration (solid red line), skMLCK activity (blue line), MLCP active (green line), and RLC phosphorylation (purple line) as a result of 2 s of high‐frequency (100‐150 Hz) tetanic stimulation. Also shown is the proposed change in resting myoplasmic Ca2+ (dotted red line) as a result of stimulation. Due to the rapid time course for RLC phosphorylation during and the relatively slow time course for RLC dephosphorylation following stimulation, RLC phosphorylation is proposed to act as a “molecular memory” of muscle activity by increasing the Ca2+ sensitivity of the contractile element of fast twitch skeletal muscle. During lower‐frequency stimulation, skMLCK activity and RLC phosphorylation occur on a slower time scale but occur rapidly enough to cause staircase potentiation (not shown). The contribution of stimulus‐induced changes to resting myoplasmic Ca2+ is illustrated as relatively minor for PTP; this mechanism may be more prevalent during staircase potentiation. Figure redrawn, from Ref. 349 and modified, according to Ref. 337. Note that skMLCK catalytic activity is slower than is formation of the active skMLCK holoenzyme by Ca2+‐CaM binding (not shown). Note that all parameters are expressed relative to the maximal response for the time window depicted, not absolute levels.


Figure 5. Influence of RLC phosphorylation on arthropod thick filaments. (A) Three‐dimensional reconstruction of tarantula thick filament segment showing four helical tracks of myosin S‐1 heads. One myosin interacting‐heads motif is highlighted in yellow. Densities of interaction between heads are labelled as “a,” “b,” and “c.” In this configuration, individual myosin heads are less likely to be able to interact with actin, turnover ATP, and/or produce force. Disruption of this motif by phosphorylation may represent a type of thick filament activation that promotes force and ATPase activity. Vertical calibration bar represents 14.5 nm. Reprinted from Alamo et al. 2008 with permission (8). (B) Cartoon showing how progressive phosphorylation of the RLC (i.e., nonphosphorylated‐monophosphorylated‐diphosphorylated) influences the distribution of “free” and “blocked” myosin heads. In rested state (a), constitutive monophosphorylation of the free head RLC (yellow dot) is sufficient to produce a population of free heads that “sway” (double arrow) between close contact with their unphosphorylated and “blocked” dimeric partner head and weak contact with the thin filament. Because of their mobility, these heads bind readily to actin during rapid Ca2+ activation of the contractile apparatus (i.e., during a twitch) (b). In (c), activation of skMLCK and diphosphorylation of free heads or monophosphorylation of blocked heads increases the fraction of heads able to bind to actin. In this model, the greater the phosphorylation of the RLC by skMLCK, the greater the potentiation, either by biasing the probability that a free head will bind to actin or by allowing the skMLCK enzyme to access and phosphorylate the RLC of the blocked head. Although withdrawal of Ca2+ by SERCA pumping and reduction of [Ca2+]myo causes relaxation (d) the slow dephosphorylation of the RLC produces a “molecular memory” at the level of the RLC that will produce PTP when the muscle is reactivated (e). Note that the extent to which this scheme applies to mammalian thick filaments is not known. Colour Key: free and swaying heads are shown in light and dark blue, respectively; blocked heads are shown in green; phosphorylation sites on RLC are shown in yellow and red (Ser35 and Ser45, respectively). Monophosphorylation is produced by phosphorylation of either Ser35 or Ser45, while diphosphorylation requires both sites to be phosphorylated. Reprinted from Brito et al. (2015) with permission (42). The reader is referred to original papers for more details of the interacting head motif or the molecular mechanism of this thick filament activation by phosphorylation.


Figure 6. Influence of RLC phosphorylation on arthropod thick filaments. (A) Three‐dimensional reconstruction of tarantula thick filament segment showing four helical tracks of myosin S‐1 heads. One myosin interacting‐heads motif is highlighted in yellow. Densities of interaction between heads are labelled as “a,” “b,” and “c.” In this configuration, individual myosin heads are less likely to be able to interact with actin, turnover ATP, and/or produce force. Disruption of this motif by phosphorylation may represent a type of thick filament activation that promotes force and ATPase activity. Vertical calibration bar represents 14.5 nm. Reprinted from Alamo et al. 2008 with permission (8). (B) Cartoon showing how progressive phosphorylation of the RLC (i.e., nonphosphorylated‐monophosphorylated‐diphosphorylated) influences the distribution of “free” and “blocked” myosin heads. In rested state (a), constitutive monophosphorylation of the free head RLC (yellow dot) is sufficient to produce a population of free heads that “sway” (double arrow) between close contact with their unphosphorylated and “blocked” dimeric partner head and weak contact with the thin filament. Because of their mobility, these heads bind readily to actin during rapid Ca2+ activation of the contractile apparatus (i.e., during a twitch) (b). In (c), activation of skMLCK and diphosphorylation of free heads or monophosphorylation of blocked heads increases the fraction of heads able to bind to actin. In this model, the greater the phosphorylation of the RLC by skMLCK, the greater the potentiation, either by biasing the probability that a free head will bind to actin or by allowing the skMLCK enzyme to access and phosphorylate the RLC of the blocked head. Although withdrawal of Ca2+ by SERCA pumping and reduction of [Ca2+]myo causes relaxation (d) the slow dephosphorylation of the RLC produces a “molecular memory” at the level of the RLC that will produce PTP when the muscle is reactivated (e). Note that the extent to which this scheme applies to mammalian thick filaments is not known. Colour Key: free and swaying heads are shown in light and dark blue, respectively; blocked heads are shown in green; phosphorylation sites on RLC are shown in yellow and red (Ser35 and Ser45, respectively). Monophosphorylation is produced by phosphorylation of either Ser35 or Ser45, while diphosphorylation requires both sites to be phosphorylated. Reprinted from Brito et al. (2015) with permission (42). The reader is referred to original papers for more details of the interacting head motif or the molecular mechanism of this thick filament activation by phosphorylation.


Figure 7. Theoretical relationship between Ca2+ sensitivity and isometric twitch force. Theoretical relationship between skMLCK catalyzed RLC phosphorylation mediated increases in Ca2+ sensitivity and isometric twitch force in permeabilized skeletal muscle fibers and intact rodent skeletal muscle, respectively. Main graph: representative force‐pCa relationship of permeabilized fibers where red line is without RLC phosphorylation and blue line is with RLC phosphorylation. In the highly phosphorylated state (>0.70 mol phos mol RLC), the force‐pCa relationship is shifted to the left relative to unphosphrylated state (253,354,356) without effect on maximal Ca2+ activated force (Fmax). Top left panel insert: example traces from EDL muscle (in vitro, 25°C) showing how presence of skMLCK enzyme and stimulation induced elevations in RLC phosphorylation is associated with PTP of isometric twitch force (before stimulation, thick line; after stimulation, thin line). Bottom right panel insert: example traces from skMLCK−/− mouse EDL muscle (in vitro, 25°C) showing how the absence of skMLCK enzyme and absence of stimulation induced RLC phosphorylation is associated with posttetanic depression of force rather than PTP (before stimulation, thick line; after stimulation, thin line). Note that changes to twitch time course (i.e., decreased half‐relaxation time) are roughly similar for both genotypes and thus apparently independent of RLC phosphorylation and PTP per se. Repetitive low frequency stimulation leading to staircase potentiation may produce increases in isometric twitch force amplitude that approach that displayed by WT muscles, indicating the presence of a RLC phosphorylation‐independent mechanism.


Figure 8. Summary of RLC phosphorylation effects on permeabilized fibers. Flow chart depicting theoretical cascade of events during Ca2+ activation of permeabilized skeletal muscle fibers. Distinction is made between suboptimal and optimal Ca2+ activation. Cascade based on results from mammalian muscle fibers at low temperatures. The influence of RLC phosphorylation on αFS is assumed to be due to an increase in fapp without change to gapp. Other mechanisms such as temperature, ionic strength, or Ca2+ binding to cation sites on the RLC are not considered in this model. Abbreviations: fraction of cycling cross‐bridges in force‐generating state (αFS); AM ATPase; loaded shortening velocity (VL); rate constant for steady‐state tension development (ktr or kREDEV); rate of relaxation from steady state (RFR); unloaded shortening velocity as determined by the slack test (Vu).


Figure 9. Influence of potentiation on force‐frequency relation of mouse skeletal muscle. Illustration of the influence of PTP on the force frequency responses of WT mouse EDL muscle (in vitro, 25°C) when measured at the peak of potentiation typically observed 10 to 20 s after a tetanic CS. (A) Isometric force‐frequency responses at optimal length in the unpotentiated (red) and potentiated (blue) state. Unpotentiated and potentiated forces (F) are normalized to peak force (Fo) in the unpotentiated state. The relatively small portion of the activation frequency domain for isometric PTP relative to the frequency stimulus range for fast muscle in vivo is indicated at top. Note that when isometric contractions are very brief (i.e., <100 ms in duration) the relative effect of PTP on the peak force that is attained is much greater. (B) Concentric force‐frequency responses in the unpotentiated (red) and potentiated (blue) state. To construct this model a shortening velocity of 0.35 Vmax was used for both unpotentiated and potentiated conditions; faster and slower speeds of shortening will shift the potentiated curve upward and downward, respectively, relative to the unpoentiateed curve (based on Refs. 128 and 382). Unpotentiated and potentiated forces (F) are normalized to peak force (Fo) in the unpotentiated state. The congruence between the activation frequency domains for concentric PTP relative to the frequency stimulus range for fast muscle in vivo is indicated at top. Comparison of panels A and B reveals that the influence of PTP is much greater for concentric than for isometric contractions both in terms of the magnitude of increase and the stimulus frequency range over which potentiation is evident.


Figure 10. Influence of potentiation on force‐velocity relation of mouse skeletal muscle. Illustration of the influence of PTP on the force‐velocity responses of WT mouse EDL muscle (in vitro, 25°C) when measured at the peak of potentiation typically observed 10 s after a tetanic CS. Note that, because the influence of potentiation on concentric force is greatest at moderate shortening speeds and smallest at slow and fast shortening speeds, the potentiated force‐velocity curve is less concave than the unpotentiated force‐velocity curve. Curves represent situations in which muscles are stimulated at high frequency to produce maximal responses in both unpotentiated and potentiated state. Vertical arrow denotes effect on potentiated curve relative to unpotentiated curve if stimulus frequency is increased or decreased (down and up, respectively). Curves are hand drawn.


Figure 11. Effect of genotype on concentric force potentiation of mouse muscle. Hypothetical model for concentric force potentiation of EDL muscles from WT and skMLCK−/− mice (in vitro, 25°). (A) The effect of mouse genotype on the time dependence of concentric force potentiation. Responses shown are for moderate frequency contractions (i.e., 45 Hz).and a moderate speed of shortening (i.e., 0.35 Vmax). Note that although WT muscle (blue line) responses are always greater, skMLCK−/− muscle (red line) also displays a significant potentiation that cannot be attributed to RLC phosphorylation. The vertical arrows represent the difference in potentiation between genotypes hypothesized to be due to RLC phosphorylation (i.e., WT – skMLCK−/− responses). Use of lower or higher‐frequency stimulation will shift both curves up or down the y‐axis, respectively. Potentiation calculated as post‐CS response at select times divided by pre‐CS value and should be considered represent situation for either peak or mean force. Based on unpublished data. (B) The effect of mouse genotype on the stimulus frequency dependence of concentric force potentiation. Values represent the peak of potentiation as captured 10 to 20 s after a tetanic CS. Note that although WT responses are always greater (blue line), it is apparent that skMLCK−/− muscles also display a potentiation at all frequencies that cannot be attributed to RLC phosphorylation. The vertical arrows represent the difference in potentiation hypothized to be solely due to RLC phosphorylation (i.e., WT – skMLCK−/− responses). Curves are based on a shortening speed of 0.35 Vmax and assume maximal potentiation as determined 10 to 20 s after a tetanic CS. Potentiation calculated as post‐CS response divided by pre‐CS value and represent either peak or mean force or work/power. Based on Ref. 126 and unpublished data.


Figure 12. Summary of RLC phosphorylation influence on intact rodent muscle. Flow chart depicting theoretical cascade of events leading to myosin RLC phosphorylation and PTP or staircase potentiation in fast twitch rodent skeletal muscle. The proposed influence of RLC phosphorylation on muscle force, work, and power is illustrated for twitch and tetanic (Pt and Po, respectively) as well as during isometric and dynamic contractions. The possible influence of estrogen on skMLCK activity and upregulating RLC phosphorylation is included. Potentiation due to alternate mechanisms is not considered (dotted lines). Abbreviatons: force‐time integral (FTI); rate of isometric force development (RFD); rate of isometric force relaxation (RFR); isometric twitch (Pt), submaximal force (<Po); peak isometric force (Po); and high‐energy phosphate turnover (HEPT). Economy refers to ratio between HEPT and FTI during isometric contractions while efficiency refers to ratio between HEPT and work performed during concentric contractions. Displacement refers to the magnitude of shortening during isotonic type contractions. Vmax refers to maximum shortening velocity as determined by extrapolation to zero load while Vu refers to unloaded shortening velocity as detrmined by the slack test. The proposed fiber type dependence of myosin RLC phosphorylation mediated potentiation demonstrated in skMLCK overexprssors (323) is recognized by the bifurcation of the flow chart and the dotted arrow pointing toward type IIa and IIa fibers. Dotted lines depict speculative link between estrogen signaling and skMLCK activity (66). Dashed lines depict speculative link between skMLCK catalyzed RLC phosphorylation, the SRX and resting muscle thermogenesis (70,247).


Figure 13. Summary of RLC phosphorylation influence on human skeletal muscle. Flow chart depicting theoretical cascade of events leading to myosin RLC phosphorylation postactivation potentiation (PAP) (staircase or PTP) in human skeletal muscle. A theoretical distinction is made between evoked and voluntary condition contractions as well as between isometric and ballistic contraction types. Studies showing that the magnitude of PAP of evoked contractions tend to be smaller in magnitude than for voluntary contractions is recognized and accounted for by the presence of signals arising in the periphery that act as a negative feedback signal to upper centers to dampen PAP (dashed lines). Dotted lines depict speculative link between reductions in estrogen levels and reductions in phosphorylation of fast and slow RLC isoforms noted in older women relative to either young women or age matched males (259). The potential influence of resting RLC phosphorylation levels on muscle thermogenesis, via redistribution of cross‐bridges from SRX to CRX, is included (dotted lines). Ballistic contractions refer to isokinetic contractions at fast (>180°/s) or slow (<180°/s) speeds of shortening. Abbreviations: evoked twitch (Pt); evoked submaximal force (Pt‐Po); evoked peak force (Po); rate of isometric force development (RFD); maximal voluntary contraction (MVC); submaximal voluntary contraction (SVC); and motor unit discharge rate (MUDR). Due to technical limitations, the influence of PAP on MUDR during ballistic contractions has not been determined. Neither metabolic factors inhibiting or alternate mechanisms for PAP are considered in this flowchart (although they may well exist).
References
 1.Abbate F, Bruton JD, De Haan A, Westerblad H. Prolonged force increase following a high‐frequency burst is not due to a sustained elevation of [Ca2+]i. Am J Physiol Cell Physiol 283(1): C42‐C47, 2002.
 2.Abbate F, Sargeant AJ, Verdijk PW, de Haan A. Effects of high‐frequency initial pulses and posttetanic potentiation on power output of skeletal muscle. J Appl Physiol 88: 35‐40, 2000.
 3.Abbate F, Van Der Velden J, Stienen GJ, De Haan A. Posttetanic potentiation increases energy cost to a higher extent than work in rat fast skeletal muscle. J Musc Res Cell Motil 22: 703‐710, 2001.
 4.Adam A, De Luca CJ. Firing rates of motor units in human vastus lateralis muscle during fatiguing isometric contractions. J Appl Physiol 99: 268‐280, 2005.
 5.Adhikari BB, Somerset J, Stull JT, Fajer PG. Dynamic modulation of the regulatory domain of myosin heads by pH, ionic strength, and RLC phosphorylation in synthetic myosin filaments. Biochemistry 38(10): 3127‐3132, 1999.
 6.Alamo L, Li XE, Espinoza‐Fonseca LM, Pinto A, Thomas DD, Lehman W, Padrón R. Tarantula myosin free head regulatory light chain phosphorylation stiffens N‐terminal extension, releasing it and blocking its docking back. Mol Biosyst 11: 2180‐2189, 2015.
 7.Alamo L, Qi D, Wriggers W, Pinto A, Zhu J, Bilbao A, Gillilan RE, Hu S, Padron R. Conserved intramolecular interactions maintain myosin interacting‐heads motifs explaining tarantula muscle super‐relaxed state structural basis. J Mol Biol 428, 1142‐1164, 2016.
 8.Alamo L, Wriggers W, Pinto A, Bartoli F, Salazar L, Zhao FQ, Craig R, Padron R. Three‐dimensional reconstruction of tarantula myosin filaments suggests how phosphorylation may regulate myosin activity. J Mol Biol 384(4): 780‐797, 2008.
 9.Allen DG, Lamb GD, Westerblad H. Skeletal muscle fatigue: Cellular mechanisms. Physiol Rev 88(1): 287‐332, 2008.
 10.Allen DG, Lee JA, Westerblad H. Intracellular calcium and tension during fatigue in isolated single muscle fibres from Xenopus laevis. J Physiol 415: 433‐458, 1989.
 11.Bagshaw CR, Kendrick‐Jones J. Identification of the divalent metal ion binding domain of myosin regulatory light chains using spin‐labelling techniques. J Mol Biol 140(3):411‐433, 1980.
 12.Bagshaw CR, Reed GH. The significance of the slow dissociation of divalent metal ions from myosin ‘regulatory’ light chains. FEBS Lett 81: 386‐390, 1977.
 13.Bagshaw CR, Trentham DR. The reversibility of adenosine triphosphate cleavage by myosin. Biochem J 133(2):323‐328, 1973.
 14.Bagust J, Lewis DM, Luck JC. Post‐tetanic effects in motor units of fast and slow twitch muscle of the cat. J Physiol 237(1): 115‐121, 1974.
 15.Bárány K, Bárány M. Phosphorylation of the 18,000‐dalton light chain of myosin during a single tetanus of frog muscle. J Biol Chem 252(14): 4752‐4754, 1977.
 16.Bárány K, Bárány M, Gillis JM, Kushmerick MJ. Phosphorylation‐dephosphorylation of the 18,000‐dalton light chain of myosin during the contraction‐relaxation cycle of frog muscle. J Biol Chem 254(9): 3617‐3623, 1979.
 17.Bárány M. ATPase activity of myosin correlated with speed of muscle shortening. J Gen Physiol 50(6): S197‐S218, 1967.
 18.Bárány M, Close RI. The transformation of myosin in cross‐innervated rat muscles. J Physiol 213(2):455‐474, 1971.
 19.Barclay CJ. Effect of fatigue on rate of isometric force development in mouse fast‐ and slow‐twitch muscles. Am J Physiol 263(32): 1065‐1072, 1992.
 20.Barclay CJ. Quantifying Ca2+ release and inactivation of Ca2+ release in fast‐ and slow‐twitch muscles. J Physiol 590(23): 6199‐6212, 2012.
 21.Barsotti RJ, Butler TM. Chemical energy usage and myosin light chain phosphorylation in mammalian skeletal muscle. J Musc Res Cell Motil 5(1): 45‐64, 1984.
 22.Baudry S, Duchateau J. Postactivation potentiation in human muscle is not related to the type of maximal conditioning contraction. Muscle Nerve 30(3): 328‐336, 2004.
 23.Baudry S, Duchateau J. Postactivation potentiation in a human muscle: Effect on rate of torque development of tetanic and voluntary isometric contractions. J App Physiol 102: 1394‐1401, 2007.
 24.Baudry S, Duchateau J. Postactivation potentiation in a human muscle: Effect on the load‐velocity relation of tetanic and voluntary shortening contractions. J Appl Physiol 103(4): 1318‐1325, 2007.
 25.Baudry S, Klass M, Duchateau J. Postactivation potentiation influences differently the nonlinear summation of contractions in young and elderly adults. J Appl Physiol 98(4): 1243‐1250, 2005.
 26.Baudry S, Klass M, Duchateau J. Postactivation potentiation of short tetanic contractions is differently influenced by stimulation frequency in young and elderly adults. Eur J Appl Physiol 103(4): 449‐459, 2008.
 27.Bauer CB, Holden HM, Thoden JB, Smith R, Rayment I. X‐ray structures of the apo and MgATP‐bound states of Dictyostelium discoideum myosin motor domain. J Biol Chem 275(49): 38494‐3899, 2000.
 28.Baumann BA, Hambly BD, Hideg K, Fajer PG. The regulatory head of the myosin head acts as a rigid lever. Biochemistry 40(26):7868‐7873, 2001.
 29.Bernhard CG, von Euler US, Skoglund CR. Post‐tetanic action potentials in mammalian muscle. Acta Physiol Scand 2: 284‐288, 1941.
 30.Bicer S, Reiser PJ. Myosin light chain isoform expression among single mammalian skeletal muscle fibres: Species variations. J Musc Res Cell Motil 25: 623‐633, 2004.
 31.Bigland‐Ritchie B, Johansson R, Lippold OC, Smith S, Woods J. Changes in motoneurone firing rates during sustained maximal voluntary contractions. J Physiol 340, 335‐346, 1983.
 32.Bigland‐Ritchie B, Jones D, Woods J. Excitation frequency and muscle fatigue: Electrical responses during human voluntary and stimulated contractions. Exp Neurol 64: 414‐427, 1979.
 33.Blaauw B, Schiaffino S, Reggiani C. Mechanisms modulating skeletal muscle phenotype. Compr Physiol 3(4): 1645‐1687, 2013.
 34.Blinks JR, Rudel R, Taylor SR. Calcium transients in isolated amphibian skeletal muscle fibres: Detection with aequorin. J Physiol 277: 291‐323, 1978.
 35.Blumenthal DK, Stull JT. Activation of skeletal muscle myosin light chain kinase by calcium(2+) and calmodulin. Biochemistry 19(24): 5608‐5614, 1980.
 36.Blumenthal DK, Stull JT. Effects of pH, ionic strength, and temperature on activation by calmodulin and catalytic activity of myosin light chain kinase. Biochemistry 21(10): 2386‐2391, 1982.
 37.Bozzo C, Spolaore B, Toniolo L, Stevens L, Bastide B, Cieniewski‐ Bernard C, Fontana A, Mounier Y, Reggiani C. Nerve influence on myosin light chain phosphorylation in slow and fast skeletal muscles. FEBS J 272: 5771‐5785, 2005.
 38.Bozzo C, Stevens L, Toniolo L, Mounier Y, Reggiani C. Increased phosphorylation of myosin light chain associated with slow‐to‐fast transition in rat soleus. Am J Physiol (Cell Physiol) 285: 575‐583, 2003.
 39.Brenner B. Effect of Ca2+ on cross‐bridge turnover kinetics in skinned single rabbit psoas fibers: Implications for regulation of muscle contraction. Proc Nati Acad Sci USA 85: 3265‐3269, 1988.
 40.Brenner B, Eisenberg E. Rate of force generation in muscle: Correlation with actomyosin ATPase activity in solution. Proc Natl Acad Sci U S A 83(10): 3542‐3546, 1986.
 41.Brenner B. Muscle mechanics II: Skinned muscle fibers. In: Sugi H, editor. Current Methods in Muscle Physiology: Advantages, Problems and Limitations. Oxford University Press: New York, 1998.
 42.Brito R, Alamo L, Lundberg U, Guerrero JR, Pinto A, Sulbaran G, Gawinowicz MA, Craig R, Padron R. A molecular model of phosphorylation‐based activation and potentiation of tarantula muscle thick filaments. J Mol Biol 414(1): 44‐61, 2011.
 43.Brown IA, Loeb GE. Post‐activation potentiation: A clue for simplifying models of muscle dynamics. Am Zool 38: 743‐754, 1988.
 44.Brown IE, Loeb GE. Measured and modeled properties of mammalian skeletal muscle. I. The effects of post‐activation potentiation on the time course and velocity dependencies of force production. J Musc Res Cell Motil 20: 443‐456, 1999.
 45.Brown LT, Tuttle WW. The phenomenon of treppe in intact human skeletal muscle. Am J Physiol 77: 483‐490, 1926.
 46.Brown GL, von Euler US. The after effects of a tetanus on mammalian muscle. J Physiol 93: 39‐60, 1938.
 47.Brunello E, Caremani M, Melli L, Linari M, Fernandez‐Martinez M, Narayanan T, Irving M, Piazzesi G, Lombardi V, Reconditi M. The contributions of filaments and cross‐bridges to sarcomere compliance in skeletal muscle. J Physiol 592(Pt 17): 3881‐399, 2014.
 48.Buller AJ, Kean CJ, Ranatunga KW, Smith JM. Post‐tetanic depression of twitch tension in the cat soleus muscle. Exp Neurol 73(1): 78‐89, 1981.
 49.Butler TM, Seigman MJ, Mooers SV, Barsotti RJ. Myosin light chain phosphorylation does not modulate cross bridge cycling in mouse skeletal muscle. Science 220: 1167‐1169, 1983.
 50.Calderón JC, Bolaños P, Caputo C. Tetanic Ca2+ transient differences between slow‐ and fast‐twitch mouse skeletal muscle fibres: A comprehensive experimental approach. J Muscle Res Cell Motil 35(5‐6): 279‐293, 2014.
 51.Carpentier A, Duchateau J, Hainaut K. Motor unit behaviour and contractile changes during fatigue in the human first dorsal interosseus. J Physiol 534: 903‐912, 2001.
 52.Caterini D, Gittings W, Huang J, Vandenboom R. The effect of work cycle frequency on the potentiation of dynamic function in fast mouse muscle. J Exp Biol 214: 3915‐3923, 2011.
 53.Chang AN, Battiprolu PK, Cowley PM, Chen G, Gerard RD, Pinto JR, Hill JA, Baker AJ. Constitutive phosphorylation of cardiac myosin regulatory light chain in vivo. J Biol Chem 290(17):10703‐10716, 2015.
 54.Chase PB, Martyn DA, Hannon JD. Isometric force redevelopment of skinned muscle fibers from rabbit activated with and without Ca2+. Biophys J 67: 1994‐2001, 1994.
 55.Chiarandini DJ, Stefani E. Twitch potentiation by potassium contractures in single muscle fibres of the frog. J Physiol 240(1):1‐14, 1974.
 56.Childers MK, McDonald KS. Regulatory light chain phosphorylation increases eccentric contraction‐induced injury in skinned fast‐twitch fibers. Muscle Nerv 29: 313‐317, 2004.
 57.Claflin DR, Brooks SV. Direct observation of failing fibers in muscles of dystrophic mice provides mechanistic insight into muscular dystrophy. Am J Physiol Cell Physiol 294: C651‐C658, 2008.
 58.Close RI. Dynamic properties of mammalian skeletal muscles. Physiol Rev 52(1): 129‐197, 1972.
 59.Close R, Hoh JF. Influence of temperature on isometric contractions of rat skeletal muscles. Nature 217(5134): 1179‐1180, 1968.
 60.Close R, Hoh JF. The after‐effects of repetitive stimulation on the isometric twitch contraction of rat fast skeletal muscle. J Physiol 197: 461‐477, 1968.
 61.Close R, Hoh JF. Post‐tetanic potentiation of twitch contractions of cross‐innervated rat fast and slow muscles. Nature 221(5176): 179‐181, 1969.
 62.Colomo F, Rocchi P. Eserine effects on single twitches and staircase phenomenon in frog nerve‐single muscle fibre preparations. Arch Fisiol 65(1): 24‐51, 1965.
 63.Colson BA, Gruber SJ, Thomas DD. Structural dynamics of muscle protein phosphorylation. J Muscle Res Cell Motil 33(6): 419‐429, 2012.
 64.Colson BA, Locher MR, Bekyarova T, Patel JR, Fitzsimons DP, Irving TC, Moss RL. Differential roles of regulatory light chain and myosin binding protein‐C phosphorylations in the modulation of cardiac force development. J Physiol 588(Pt 6):981‐993, 2010.
 65.Colson BA, Petersen KJ, Collins BC, Lowe DA, Thomas DD. The myosin super‐relaxed state is disrupted by estradiol deficiency. Biochem Biophys Res Commun 456(1): 151‐155, 2015.
 66.Connoly R, Gough W, Winegrad S. Characteristics of the isometric twitch of skeletal muscle immediately after a tetanus. A study of the influence of the distribution of calcium within the sarcoplasmic reticulum on the twitch. J Gen Physiol 57(6): 697‐709, 1971.
 67.Cooke R. Actomyosin interaction in striated muscle. Physiol Rev 77(3):671‐697, 1997.
 68.Cooke R. Modulation of the actomyosin interaction during fatigue of skeletal muscle. Muscle Nerve 36: 756‐777, 2007.
 69.Cooke R. The role of the myosin ATPase activity in adaptive thermogenesis by skeletal muscle. Biophys Rev 3(1):33‐45, 2011.
 70.Cooke R, Franks K, Stull JT. Myosin phosphorylation regulates the ATPase activity of permeable skeletal muscle fibers. FEBS Lett 144(1): 33‐37, 1982.
 71.Craig R, Alamo L, Padrón R. Structure of the myosin filaments of relaxed and rigor vertebrate striated muscle studied by rapid freezing electron microscopy. J Mol Biol 228(2):474‐487, 1992.
 72.Craig R, Padron R, Kendrick‐Jones J. Structural changes accompanying phosphorylation of tarantula muscle myosin filaments. J Cell Biol 105(3): 1319‐1327, 1987.
 73.Craig R, Woodhead JL. Structure and function of myosin filaments. Curr Opin Struct Biol 16(2):204‐212, 2006.
 74.Crow MT, Kushmerick MJ. Phosphorylation of myosin light chains in mouse fast‐twitch muscle associated with reduced actomyosin turnover rate. Science 217(4562): 835‐837, 1982.
 75.Crow MT, Kushmerick MJ. Myosin light chain phosphorylation is associated with a decrease in the energy cost for contraction in fast twitch mouse muscle. J Biol Chem 257(5): 2121‐2124, 1982.
 76.Crowther RA, Padrón R, Craig R. Arrangement of the heads of myosin in relaxed thick filaments from tarantula muscle. J Mol Biol 184(3): 429‐439. 1985.
 77.Dantzig JA, Goldman YE, Millar NC, Lacktis J, Homsher E. Reversal of the cross‐bridge force‐generating transition by photogeneration of phosphate in rabbit psoas muscle fibres. J Physiol 451: 247‐278, 1992.
 78.Davis JS, Hassanzadeh S, Winitsky SO, Satorius CL, Lin H, Vemuri R, Wen H, Epstein ND. The overall pattern of cardiac contraction depends on a spatial gradient of myosin regulatory light chain phosphorylation. Cell 107: 631‐641, 2001.
 79.Davis JS, Satorius CL, Epstein ND. Kinetic effects of myosin regulatory light chain phosphorylation on skeletal muscle contraction. Biophys J 83(1): 359‐370, 2002.
 80.Debold EP, Dave H, Fitts RH. Fiber type and temperature dependence of inorganic phosphate: Implications for fatigue. Am J Physiol Cell Physiol 287(3):C673‐C681. 2004.
 81.Debold EP, Romatowski J, Fitts RH. The depressive effect of Pi on the force‐pCa relationship in skinned single muscle fibers is temperature dependent. Am J Physiol Cell Physiol 290(4):C1041‐C1050, 2006.
 82.Decostre V, Gillis JM, Gailly P. Effect of adrenaline on the post‐tetanic potentiation in mouse skeletal muscle. J Musc Res Cell Motil 21: 247‐254, 2002.
 83.Desdmedt JE, Hainnaut TK. Kinetics of myofilament activation in potentiated contraction: Staircase phenomenon in human skeletal muscle. Nature 217(5128): 529‐532, 1968.
 84.Dias FA, Walker LA, Arteaga GM, Walker JS, Vijayan K, Peña JR, Ke Y, Fogaca RT, Sanbe A, Robbins J, Wolska BM. The effect of myosin regulatory light chain phosphorylation on the frequency‐dependent regulation of cardiac function. J Mol Cell Cardiol 41(2): 330‐339, 2006.
 85.Dickinson MH, Hyatt CJ, Lehmann FO, Moore JR, Reedy MC, Simcox A, Tohtong R, Vigoreaux JO, Yamashita H, Maughan DW. Phosphorylation‐dependent power output of transgenic flies: An integrated study. Biophys J 73(6): 3122‐3134, 1997.
 86.Diffee GM, Greaser ML, Reinach FC, Moss RL. Effects of a non‐divalent cation binding mutant of myosin regulatory light chain on tension generation in skinned skeletal muscle fibers. Biophys J 68(4): 1443‐1452, 1995.
 87.Diffee GM, Patel JR, Reinach FC, Greaser ML, Moss RL. Altered kinetics of contraction in skeletal muscle fibers containing a mutant myosin regulatory light chain with reduced divalent cation binding. Biophys J 71(1): 341‐350, 1996.
 88.Ding P, Huang J, Battiprolu PK, Hill JA, Kamm KE, Stull JT. Cardiac myosin light chain kinase is necessary for myosin regulatory light chain phosphorylation and cardiac performance in vivo. J Biol Chem 285(52): 40819‐40829, 2010.
 89.Ding J, Storaska JA, Binder‐Macleod SA. Effect of potentiation on the catchlike property of human skeletal muscles. Muscle Nerve 27: 312‐319, 2003.
 90.Dominguez R. Tropomyosin: The gatekeeper's view of the actin filament revealed. Biophysj, 100, 797‐798, 2011.
 91.Dominguez R, Holmes KC. Actin structure and function. Annu Rev Biophys 40: 169‐186, 2011.
 92.Dorfman LJ, Howard JE, McGill KC. Triphasic behavioral response of motor units to submaximal fatiguing exercise. Muscle Nerve 13 621‐628, 1990.
 93.Duggal D, Nagwekar J, Rich R, Huang W, Midde K, Fudala R, Das H, Gryczynski I, Szczesna‐Cordary D, Borejdo J. Effect of a myosin regulatory light chain mutation K104E on actin‐myosin interactions. Am J Physiol Heart Circ Physiol 308(10):H1248‐257, 2015.
 94.Duggal D, Nagwekar J, Rich R, Midde K, Fudala R, Gryczynski I, Borejdo J. Phosphorylation of myosin regulatory light chain has minimal effect on kinetics and distribution of orientations of cross bridges of rabbit skeletal muscle. Am J Physiol Regul Integr Comp Physiol 306(4): R222‐R233, 2014.
 95.Dulhunty AF. Excitation‐contraction coupling from the 1950s into the new millennium. Clin Exp Pharm Physiol 33(9): 763‐772, 2006.
 96.Ebashi S, Endo M. Calcium ion and muscle contraction. Prog Biophys Mol Biol 18: 123‐183, 1968.
 97.Ebashi S. Calcium ions and muscle contraction. Nature 240(5378): 217‐218, 1972.
 98.Ebashi S. Excitation‐contraction coupling and the mechanism of muscle contraction. An Rev Physiol 53(1): 1‐16, 1991.
 99.Ebben WP. A brief review of concurrent activation potentiation: Theoretical and practical constructs. J Strength Cond Res 20(4):985‐91, 2006.
 100.Edman KA. The velocity of unloaded shortening and its relation to sarcomere length and isometric force in vertebrate muscle fibres. J Physiol 291: 143‐159, 1979.
 101.Edman KA, Lou F. Changes in force and stiffness induced by fatigue and intracellular acidification in frog muscle fibres. J Physiol 424: 133‐149, 1990.
 102.Edman KA, Lou F. Myofibrillar fatigue versus failure of activation during repetitive stimulation of frog muscle fibres. J Physiol 457: 655‐673, 1992.
 103.Elting MW, Spudich JA. Future challenges in single‐molecule fluorescence and laser trap approaches to studies of molecular motors. Dev Cell 23(6):1084‐1091, 2012.
 104.Espinoza‐Fonseca LM, Alamo L, Pinto A, Thomas DD, Padrón R. Sequential myosin phosphorylation activates tarantula thick filament via a disorder‐order transition. Mol Biosyst 2015.
 105.Ferenczi MA, Homsher E, Simmons RM, Trentham DR. Reaction mechanism of the magnesium ion‐dependent adenosine triphosphatase of frog muscle myosin and subfragment 1. Biochem J 171: 165‐175, 1978.
 106.Finer JT, Simmons RM, Spudich JA. Single myosin molecule mechanics: Piconewton forces and nanometre steps. Nature 368: 113‐119, 1994.
 107.Fitts RH. The cross‐bridge cycle and skeletal muscle fatigue. J Appl Physiol (1985) 104(2):551‐558, 2008.
 108.Folland JP, Wakamatsu T, Fimland MS. The influence of maximal isometric activity on twitch and H‐reflex potentiation, and quadriceps femoris performance. Eur J Appl Physiol 104(4):739‐748, 2008.
 109.Ford LE, Huxley AF, Simmons RM. Tension responses to sudden length change in stimulated frog muscle fibres near slack length. J Physiol 269(2): 441‐515, 1977.
 110.Ford LE, Huxley AF, Simmons RM. Tension transients during steady shortening of frog muscle fibres. J Physiol 361: 131‐150, 1985.
 111.Fowles FJ, Green HJ. Coexistence of potentiation and low‐frequency fatigue during voluntary exercise in human skeletal muscle. Can J Physiol Pharmacol 81(12):1092‐100, 2003.
 112.Franks K, Cooke R, Stull JT. Myosin phosphorylation decreases the ATPase activity of cardiac myofibrils. J Mol Cell Cardiol 16(7):597‐604, 1984.
 113.Franks‐Skiba K, Lardelli R, Goh G, Cooke R. Myosin light chain phosphorylation inhibits muscle fiber shortening velocity in the presence of vanadate. Am J Physiol (Regul Integr Comp Physiol) 292: 1603‐1612, 2007.
 114.Fujita K, Ye LH, Sato M, Okagaki T, Nagamachi Y, Kohama K. Myosin light chain kinase from skeletal muscle regulates an ATP‐dependent interaction between actin and myosin by binding to actin. Mol Cell Biochem 190(1‐2): 85‐90, 1999.
 115.Fukutani A, Hirata K, Miyamoto N, Kanehisa H, Yanai T, Kawakami Y. Effect of conditioning contraction intensity on postactivation potentiation is muscle dependent. J Electromyogr Kinesiol 24(2): 240‐245, 2014.
 116.Fukutani A, Miyamoto N, Kanehisa H, Yanai T, Kawakami Y. Potentiation of isokinetic torque is velocity‐dependent following an isometric conditioning contraction. Springerplus 2: 554, 2013.
 117.Fusi L, Brunello E, Sevrieva IR, Sun YB, Irving M. Structural dynamics of troponin during activation of skekletal muscle. Proc Natl Acad Sci U S A 111(12): 4626‐4631, 2014.
 118.Galińska‐Rakoczy A, Engel P, Xu C, Jung H, Craig R, Tobacman LS, Lehman W. Structural basis for the regulation of muscle contraction by troponin and tropomyosin. J Mol Biol 379(5): 929‐935, 2008.
 119.Garland SJ, Garner SH, McComas AJ. Relationship between numbers and frequencies of stimuli in human muscle fatigue. J Appl Physiol (1985) 65(1):89‐93, 1988.
 120.Garland SJ, Garner SH, McComas AJ. Reduced voluntary electromyographic activity after fatiguing stimulation of human muscle. J Physiol 401: 547‐56, 1988.
 121.Garner SH, Hicks AL, McComas AJ. Prolongation of twitch potentiating mechanism throughout muscle fatigue and recovery. Exp Neurol 103(3):277‐281, 1989.
 122.Geeves MA. Review: The ATPase mechanism of myosin and actomyosin. Biopolymers 105: 483‐491, 2016.
 123.Geeves MA, Holmes KC. Structural mechanism of muscle contraction. Annu Rev Biochem 68, 687‐728, 1999.
 124.Geeves MA, Holmes KC. The molecular mechanism of muscle contraction. Adv Protein Chem 71: 161‐193, 2005.
 125.Gittings W, Aggarwal H, Stull JT, Vandenboom R. The force dependence of isometric and concentric potentiation in mouse muscles with and without myosin light chain kinase. Can J Physio Pharmacol 14: 1‐10, 2014.
 126.Gittings W, Huang J, Smith IC, Quadrilatero J, Vandenboom R. The effect of skeletal myosin light chain kinase gene ablation on the fatigability of mouse fast muscle. J Musc Res Cell Motil 31: 337‐348, 2011.
 127.Gittings W, Huang J, Vandenboom R. Tetanic force potentiation of mouse EDL muscle is shortening speed dependent. J Musc Res Cell Motil 33(5): 359‐368, 2012.
 128.Gittings W, Stull JT, Vandenboom R. Interactions between the catch like property and posttetanic potentiation of mouse skeletal muscle. Muscle Nerve 54(2): 308‐316, 2016.
 129.Godt RE, Nosek TM. Changes of intracellular milieu with fatigue or hypoxia depress contraction of skinned rabbit skeletal and cardiac muscle. J Physiol 412: 155‐180, 1989.
 130.Gordon DA, Enoka RM, Stuart DG. Motor‐unit force potentiation in adult cats during a standard fatigue test. J Physiol 421: 569‐582, 1990.
 131.Gordon AM, Homsher E, Regnier M. Regulation of contraction in striated muscle. Physiol Rev 80(2):853‐924, 2000.
 132.Gordon AM, Regnier M, Homsher E. Skeletal and cardiac muscle contractile activation: Tropomyosin “rocks and rolls”. News Physiol Sci 16: 49‐55, 2001.
 133.Gordon, AM, Huxley AF, Julian FJ. The variation in isometric tension with sarcomere length in vertebrate muscle fibres. J Physiol 184(1):170‐192, 1966.
 134.Gossen ER, Allingham K, Sale DG. Effect of temperature on post‐tetanic potentiation in human dorsiflexor muscles. Can J Physiol Pharmacol 79(1): 49‐58, 2001.
 135.Gossen ER, Ivanova TD, Garland SJ. 2003. The time course of the motoneuron after hyperpolarization is related to motor unit twitch speed in human skeletal muscle. J Physiol 552, 657‐664, 2003.
 136.Gossen ER, Sale DG. Effect of postactivation potentiation on dynamic knee extension performance. Eur J Appl Physiol 83(6): 524‐530, 2002.
 137.Grange RW, Cory CR, Vandenboom R, Houston ME. Myosin phosphorylation augments the force‐displacement and force velocity relationships of mouse fast muscle. Am J Physiol 269: 713‐724, 1995.
 138.Grange RW, Houston ME. Simultaneous potentiation and fatigue in quadriceps after a 60 second maximal voluntary isometric contraction. J Appl Physiol 70: 726‐731, 1991.
 139.Grange RW, Vandenboom R, Houston ME. Physiological significance of myosin phosphorylation in skeletal muscle. Can J Appl Physiol 18(3): 229‐242, 1993.
 140.Grange RW, Vandenboom R, Xeni J, Houston ME. Potentiation of in vitro concentric work in mouse fast muscle. J Appl Physiol 84: 236‐243, 1998.
 141.Greaser ML, Moss RL, Reiser PJ. Variations in contractile properties of rabbit single muscle fibres in relation to troponin T isoforms and myosin light chains. J Physiol 406: 85‐98, 1998.
 142.Green HJ, Jones SR. Does post‐tetanic potentiation compensate for low frequency fatigue? Clin Physiol 9(5): 499‐514, 1989.
 143.Greenberg MJ, Kazmierczak K, Szczesna‐Cordary D, Moore JR. Cardiomyopathy‐linked myosin regulatory light chain mutations disrupt myosin strain‐dependent biochemistry. Proc Natl Acad Sci U S A 107(40): 17403‐17408, 2010.
 144.Greenberg MJ, Mealy TR, Jones M, Szczesna‐Cordary D, Moore JR. The direct molecular effects of fatigue and myosin regulatory light chain phosphorylation on the actomyosin contractile apparatus. Am J Physiol Regul Integr Comp Physiol 298(4): 989‐996, 2010.
 145.Greenberg MJ, Mealy TR, Watt JD, Jones M, Szczesna‐Cordary D, Moore JR. The molecular effects of skeletal muscle myosin regulatory light chain phosphorylation. Am J Physiol Regul Integr Comp Physiol 297(2): 265‐274, 2009.
 146.Greenberg MJ, Watt JD, Jones M, Kazmierczak K, Szczesna‐Cordary D, Moore JR. Regulatory light chain mutations associated with cardiomyopathy affect myosin mechanics and kinetics. J Mol Cell Cardiol 46(1): 108‐115, 2009.
 147.Greising SM, Baltgalvis KA, Kosir AM, Moran AL, Warren GL, Lowe DA. Estradiol's beneficial effect on murine muscle function is independent of muscle activity. J Appl Physiol (1985) 110(1):109‐115, 2011.
 148.Greising SM, Carey RS, Blackford JE, Dalton LE, Kosir AM, Lowe DA. Estradiol treatment, physical activity, and muscle function in ovarian‐senescent mice. Exp Gerontol 46(8): 685‐693, 2011.
 149.Griffin L, Garland SJ, Ivanova T. Discharge patterns in human motor units during fatiguing arm movements. J Appl Physiol 85(5): 1684‐1692, 1998.
 150.Gruber CM. Studies in fatigue: The staircase phenomenon in mammalian skeletal muscle. Am J Physiol 63(2): 338‐349, 1923.
 151.Gulick AM, Bauer CB, Thoden JB, Pate E, Yount RG, Rayment I. X‐ray structures of the Dictyostelium discoideum myosin motor domain with six non‐nucleotide analogs. J Biol Chem 275(1): 398‐408, 2000.
 152.Gulick AM, Rayment I. Structural studies on myosin II: Communication between distant protein domains. Bioessays 19(7): 561‐569, 1997.
 153.Guttman SA, Horton RG, Wilber DT. Enhancement of muscle contraction after tetanus. Am J Physiol 119(3): 463‐473, 1937.
 154.Hamada T, Sale D, MacDougall JD, Tarnopolsky MA. Postactivation potentiation, fiber type, and twitch contraction time in human knee extensor muscles. J Appl Physiol 88: 2131‐2137, 2000.
 155.Hamada T, Sale DG, MacDougall JD, Tarnopolsky MA. Interaction of fibre type, potentiation and fatigue in human knee extensor muscles. Acta Physiol Scand 178(2): 165‐173, 2003.
 156.Hamada T, Sale DG, Macdougall JD. Postactivation potentiation in endurance‐trained male athletes. Med Sci Sports Exerc 32(2):403‐411, 2000.
 157.Hanson J. The effects of repetitive stimulation on the action potential and the twitch of rat muscle. Acta Physiol Scand 90, 387‐400, 1974.
 158.Hartman MA, Spudich JA. The myosin superfamily at a glance. J Cell Sci 125(7): 1627‐1632, 2012.
 159.Heissler SM, Sellers JR. Myosin light chains: Teaching old dogs new tricks. Bioarchitecture 4: 169‐188, 2014.
 160.Herring BP, Nunnally MH, Gallagher PJ, Stull JT. Molecular characterization of rat skeletal muscle myosin light chain kinase. Am J Physiol 256(2 Pt 1):C399‐C404, 1989.
 161.Hibberd MG, Dantzig JA, Trentham DR, Goldman YE. Phosphate release and force generation in skeletal muscle fibers. Science 228(4705): 1317‐1319, 1985.
 162.Hidalgo C, Craig R, Ikebe M, Padron R. Mechanism of phosphorylation of the regulatory light chain of myosin from tarantula striated muscle. J Musc Res Cell Motil 22(1): 51‐59, 2001.
 163.Hodgson DM, Docherty D, Robbins D. Post‐activation potentiation: Underlying physiology and implications for motor performance. Sports Med 35: 585‐595, 2005.
 164.Hofmann PA, Metzger JM, Greaser ML, Moss RL. Effects of partial extraction of light chain 2 on the Ca2+ sensitivities of isometric tension, stiffness, and velocity of shortening in skinned skeletal muscle fibers. J Gen Physiol 95(3): 477‐498, 1990.
 165.Homsher E. Muscle enthalpy production and its relationship to actomyosin ATPase. Ann Rev Physio 49(1): 673‐690, 1987.
 166.Hooijman P, Stewart MA, Cooke R. A new state of cardiac myosin with very slow ATP turnover: A potential cardioprotective mechanism in the heart. Biophys J 100(8):1969‐1976, 2011.
 167.Houston ME, Grange RW. Myosin phosphorylation, twitch potentiation, and fatigue in human skeletal muscle. Can J Physiol Pharmacol 68(7): 908‐913, 1990.
 168.Houston ME, Grange RW. Torque potentiation and myosin light‐chain phosphorylation in human muscle following a fatiguing contraction. Can J Physiol Pharmacol 69(2): 269‐273, 1991.
 169.Houston ME, Green HJ, Stull JT. Myosin light chain phosphorylation and isometric twitch potentiation in intact human muscle. Pflugers Arch 403(4): 348‐352, 1985.
 170.Houston ME, Lingley MD, Stuart DS, Grange RW. Myosin light chain phosphorylation in intact human muscle. FEBS Lett 219(2): 469‐471, 1987.
 171.Hughes J R. Post‐tetanic potentiation. Physiol Rev 38: 91‐113, 1958.
 172.Huxley AF. Muscle structure and theories of contraction. Prog Biophys Biophys Chem 7: 255‐318, 1957.
 173.Huxley AF. Muscular contraction. J Physiol 243(1):1‐43, 1974.
 174.Huxley AF. Muscle contraction. Cross‐bridge tilting confirmed. Nature 375(6533): 631‐632, 1995.
 175.Huxley AF, Niedergerke R. Structural changes in muscle during contraction; interference microscopy of living muscle fibres. Nature 173(4412):971‐973, 1954.
 176.Huxley AF, Taylor RE. Local activation of striated muscle fibres. J Physiol 144(3): 426‐441, 1958.
 177.Huxley HE, Hanson J. Changes in the cross‐striations of muscle during contraction and stretch and their structural interpretation. Nature 173(4412):973‐976, 1954.
 178.Huxley HE. The mechanism of muscular contraction. Science 164(3886):1356‐1365, 1969.
 179.Inglis JG, Howard J, McIntosh K, Gabriel DA, Vandenboom R. Decreased motor unit discharge rate in the potentiated human tibialis anterior muscle. Acta Physiol 201(4): 483‐492, 2011.
 180.Irving T, Bhattacharya S, Tesic I, Moore J, Farman G, Simcox A, Vigoreaux J, 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(8):675‐683, 2001.
 181.Irving TC, Millman BM. Changes in thick filament structure during compression of the filament lattice in relaxed frog sartorius muscle. J Muscle Res Cell Motil 10(5):385‐394, 1989.
 182.Isaacson A. Variability of twitch potentiation in frog skeletal muscle. Nature 196: 381‐382, 1962.
 183.Isaacson A. Post‐staircase potentiation, a long‐lasting twitch potentiation of muscles induced by previous activity. Life Sci 8(7): 337‐342, 1969.
 184.Jami L, Murthy KS, Petit J, Zytnicki D. After‐effects of repetitive stimulation at low frequency on fast‐contracting motor units of cat muscle. J Physiol 340: 129‐143, 1983.
 185.Jiang Y, Johnson JD, Rall JA. Parvalbumin relaxes frog skeletal muscle when sarcoplasmic reticulum Ca(2+)‐ATPase is inhibited. Am J Physiol 270(2 Pt 1):C411‐C417, 1996.
 186.Jiang Y, Julian FJ. Effects of ramp shortening during linear phase of relaxation on [Ca2+]i in intact skeletal muscle fibers. Am J Physiol 276(1 Pt 1):C152‐C160, 1999.
 187.Josephson RK. Contraction dynamics and power output of skeletal muscle. Contraction dynamics and power output of skeletal muscle. Annu Rev Physiol 55: 527‐546, 1993.
 188.Jung HS, Komatsu S, Ikebe M, Craig R. Head‐head and headtail interaction: A general mechanism for switching off myosin II activity in cells. Mol Biol Cell 19(8): 3234‐3242, 2008.
 189.Kad NM, Kim S, Warshaw DM, VanBuren P, Baker JE. Single‐myosin crossbridge interactions with actin filaments regulated by troponin‐tropomyosin. Proc Natl Acad Sci U S A 102(47):16990‐16995, 2005.
 190.Kad NM, Rovner AS, Fagnant PM, Joel PB, Kennedy GG, Patlak JB, Warshaw DM, Trybus KM. A mutant heterodimeric myosin with one inactive head generates maximal displacement. J Cell Biol 162: 481‐488, 2003.
 191.Kamm KE, Stull JT. Signaling to myosin regulatory light chain in sarcomeres. J Biol Chem 286(12):9941‐9947, 2011.
 192.Kandel E, Schwartz J, Jessel T. Principles of Neural Science (4th ed). McGraw‐Hill: New York, 2000.
 193.Karatzaferi C, Franks‐Skiba K, Cooke R. Inhibition of shortening velocity of skinned skeletal muscle fibers in conditions that mimic fatigue. Am J Physiol (Regul Integr Comp Physiol) 294: 948‐955, 2008.
 194.Kensler R, Peterson S, Norberg M. The effect of changes in temperature or ionic strength on isolated rabbit and fish skeletal muscle thick filaments. J Muscle Res Cell Motil 15: 69‐79, 1994.
 195.Kensler R, Stewart M. The relaxed crossbridge pattern in isolated rabbit psoas muscle thick filaments. J Cell Sci 105: 841‐848, 1993.
 196.Kensler RW, Levine RJ. An electron microscopic and optical diffraction analysis of the structure of Limulus telson muscle thick filaments. J Cell Biol 92(2):443‐451, 1982.
 197.Kitamura K, Tokunaga M, Esaki S, Iwane AH, Yanagida T. Mechanism of muscle contraction based on stochastic properties of single actomyosin motors observed in vitro. Biophysics 1: 1‐19, 2005.
 198.Klein CS, Ivanova TD, Rice CL, Garland SJ. Motor unit discharge rate following twitch potentiation in human triceps brachii muscle. Neurosci Lett 316(3): 153‐156, 2001.
 199.Klein CS, Rice CL, Ivanova TD, Garland SJ. Changes in motor unit discharge rate are not associated with the amount of twitch force potentiation in old men. J Appl Physiol 96: 1616‐1621, 2002.
 200.Klug GA, Biedermann M, Houston ME, Stuart D, Mumby M, Stull JT. Chronic low frequency stimulation reduces myosin phosphorylation in rabbit fast twitch muscle. Can J Physiol Pharmacol 70(6): 859‐865, 1992.
 201.Klug GA, Botterman BR, Stull JT. The effect of low frequency stimulation on myosin light chain phosphorylation in skeletal muscle. J Biol Chem 257: 4670‐4688, 1982.
 202.Klug GA, Houston ME, Stull JT, Pette D. Decrease in myosin light chain kinase activity of rabbit fast muscle by chronic stimulation. FEBS Lett 200(2): 352‐354, 1986.
 203.Krarup C. Electrical and mechanical responses in the platysma and in the adductor pollicis muscle: In normal subjects. J Neurol Neurosurg Psychiat 40: 234‐240, 1977.
 204.Krarup C. Enhancement and diminution of mechanical tension evoked by staircase and by tetanus in rat muscle. J Physiol 311: 355‐372, 1981.
 205.Krarup C. Temperature dependence of enhancement and diminution of tension evoked by staircase and by tetanus in rat muscle. J Physiol 311: 373‐387, 1981.
 206.Krarup C. The effect of dantrolene on the enhancement and diminution of tension evoked by staircase and by tetanus in rat muscle. J Physiol 311: 389‐400, 1981.
 207.Krarup C, Horowitz SH. Evoked responses of the elbow flexors in control subjects and in myopathy patients. Muscle Nerve 2: 465‐477, 1979.
 208.Krendel M, Mooseker MS. Myosins: Tails (and heads) of functional diversity. Physiology 20: 239‐251, 2005
 209.Kushmerick MJ, Paul RJ. Aerobic recovery metabolism following a single isometric tetanus in frog sartorius muscle at 0°C. J Physiol 254: 693‐709, 1976.
 210.Lai S, Collins BC Colson BA, Kararigas G, Lowe DA. Estradiol modulates myosin regulatory light chain phosphorylation and contractility in skeletal muscle of female mice. Am J Physiol Endocrinol Metab 310: E724‐E733, 2016.
 211.Lawson JD, Pate E, Rayment I, Yount RG. Molecular dynamics analysis of structural factors influencing back door Pi release in myosin. Biophys J 86(6): 3794‐3803, 2004.
 212.Lee FS. The cause of treppe. Am J Physiol 18: 267‐282, 1907.
 213.Lehman W, Galinska‐Rakoczy A, Hatch V, Tobacman LS, Craig R. Structural basis for the activation of muscle contraction by troponin and tropomyosin. J Mol Biol 388(4): 673‐681, 2009.
 214.Levine RJ. Evidence of overlapping myosin heads on relaxed thick filaments of fish, frog and scallop striated muscles. J Struct Biol 110(2): 99‐110, 1993.
 215.Levine RJ, Chantler PD, Kensler RW, Woodhead JL. Effects of phosphorylation by myosin light chain kinase on the structure of Limulus thick filaments. J Cell Biol 113(3): 563‐572, 1991.
 216.Levine R, Kensler R, Yang Z, Stull J, Sweeney H. Myosin light chain phosphorylation affects the structure of rabbit skeletal muscle thick filaments. Biophys J 71: 898‐907, 1996.
 217.Levine RJ, Kensler RW, Yang Z, Sweeney HL. Myosin regulatory light chain phosphorylation and the production of functionally significant changes in myosin head arrangement on striated muscle thick filaments. Biophys J 68(4 Suppl):224S, 1995.
 218.Levine RJ, Yang Z, Epstein ND, Fananapazir L, Stull JT, Sweeney HL. Structural and functional responses of mammalian thick filaments to alterations in myosin regulatory light chains. J Struct Biol 122: 149‐161, 1998.
 219.Li XE, Orzechowski M, Lehman W, Fischer S. Structure and flexibility of the tropomyosin overlap junction. Biochem Biophys Res Commun 446(1):304‐308, 2014.
 220.Li XE, Tobacman LS, Mun JY, Craig R, Fischer S, Lehman W. Tropomyosin position on F‐actin revealed by EM reconstruction and computational chemistry. Biophys J 100(4): 1005‐1013, 2011.
 221.Liang B, Chen Y, Wang CK, Luo Z, Regnier M, Gordon AM, Chase PB. Ca2+ regulation of rabbit skeletal muscle thin filament sliding: Role of cross‐bridge number. Biophys J 85(3):1775‐1786, 2003.
 222.Lorenz D. Postactivation potentiation: An introduction. Int J Sports Phys Ther 6(3):234‐240, 2011.
 223.Lowe DA, Baltgalvis KA, Greising SM. Mechanisms behind estrogen's beneficial effect on muscle strength in females. Exerc Sport Sci Rev 38(2): 61‐67, 2010.
 224.Lowey S, Trybus KM. Role of skeletal and smooth muscle myosin light chains. Biophys J 68(4 Suppl):120S‐126S; 1995.
 225.Lowey S, Waller GS, Trybus KM. Skeletal muscle myosin light chains are essential for physiological speeds of shortening. Nature 365(6445):454‐456, 1993.
 226.Lowey S, Waller GS, Trybus KM. Function of skeletal muscle myosin heavy and light chain isoforms by an in vitro motility assay. J Biol Chem 268(27): 20414‐20418, 1993.
 227.Luo Y, Davis JP, Smillie LB, Rall JA. Determinants of relaxation rate in rabbit skinned skeletal muscle fibres. J Physiol 545(Pt 3): 887‐901, 2002.
 228.MacIntosh BR. Role of calcium sensitivity modulation in skeletal muscle performance. News Physiol Sci: 18: 222‐225, 2003.
 229.MacIntosh BR. Review: Cellular and whole muscle studies of activity dependent potentiation. Adv Exp Med Biol 682: 315‐342, 2010.
 230.MacIntosh BR, Bryan SN. Potentiation of shortening and velocity of shortening during repeated isotonic tetanic contractions in mammalian skeletal muscle. Pflugers Arch 443: 804‐812, 2002.
 231.MacIntosh BR, Gardiner PF. Posttetanic potentiation and skeletal muscle fatigue: Interactions with caffeine. Can J Physiol Pharmacol 65(2): 260‐268, 1987.
 232.MacIntosh BR, Grange RW, Cory CR, Houston ME. Myosin light chain phosphorylation during staircase in fatigued skeletal muscle. Pflugers Arch 425: 9‐15, 1993.
 233.MacIntosh BR, Kupsh CC. Staircase, fatigue, and caffeine in skeletal muscle in situ. Muscle Nerve 10(8): 717‐722. 1987.
 234.MacIntosh BR, Roberge MC, Gardiner PF. Absence of staircase following disuse in rat gastrocnemius muscle. Can J Physiol Pharmacol 66(6): 707‐713, 1988.
 235.Macintosh BR, Robillard ME, Tomaras EK. Should postactivation potentiation be the goal of your warmup? Appl Physiol Nutr Metab 37(3):546‐550, 2012.
 236.MacIntosh BR, Smith MJ, Rassier DE. Staircase but not posttetanic potentiation in rat muscle after spinal cord hemisection. Muscle Nerve 38(5): 1455‐1465, 2008.
 237.MacIntosh BR, Taub EC, Dormer GN, Tomaras EK. Potentiation of isometric and isotonic contractions during highfrequency stimulation. Pflugers Arch 456: 449‐458, 2008.
 238.MacIntosh BR, Willis JC. Force‐frequency relationship and potentiation in ammalian skeletal muscle. J Appl Physiol 88(6): 2088‐2096, 2000.
 239.Manning DR, Stull JT. Myosin light chain phosphorylation and phosphorylase A activity in rat extensor digitorum longus muscle. Biochem Biophys Res Commun 90(1): 164‐170, 1979.
 240.Manning DR, Stull JT. Myosin light chain phosphorylation—dephosphorylation in mammalian skeletal muscle. Am J Physiol 242: C234‐C241, 1982.
 241.Månsson A, Mörner J, Edman KA. Effects of amrinone on twitch, tetanus and shortening kinetics in mammalian skeletal muscle. Acta Physiol Scand 136: 37‐45, 1989.
 242.Martyn DA, Chase PB, Regnier M, Gordon AM. A simple model with myofilament compliance predicts activation‐dependent crossbridge kinetics in skinned skeletal fibers. Biophys J 83(6):3425‐3434, 2002.
 243.Matsubara I, Goldman YE, Simmons RM. Changes in the lateral filament spacing of skinned muscle fibres when cross‐bridges attach. J Mol Biol 173(1): 15‐33, 1984.
 244.Maytum R, Lehrer SS, Geeves MA. Cooperativity and switching within the three‐state model of muscle regulation. Biochemistry 38(3): 1102‐1110, 1999.
 245.McKillop DF, Geeves MA. Regulation of the interaction between actin and myosin subfragment 1: Evidence for three states of the thin filament. Biophys J 65(2):693‐701, 1993.
 246.McNamara JW, Li A, dos Remedios CG, Cooke R. The role of super‐relaxed myosin in skeletal and cardiac muscle. Biophys Rev 7(1): 5‐14. 2015.
 247.Mettler JA, Griffin L. Postactivation potentiation and muscular endurance training. Muscle Nerve 45(3): 416‐425, 2012.
 248.Metzger JM, Greaser ML, Moss RL. Variations in cross‐bridge attachment rate and tension with phosphorylation of myosin in mammalian skinned skeletal muscle fibres. J Gen Physiol 93: 855‐883, 1989.
 249.Metzger JM, Moss RL. Calcium‐sensitive cross‐bridge transitions in mammalian fast and slow skeletal muscle fibers. Science 247(4946): 1088‐1090, 1990.
 250.Metzger JM, Moss RL. Kinetics of a Ca(2+)‐sensitive cross‐bridge state transition in skeletal muscle fibers. Effects due to variations in thin filament activation by extraction of troponin C. J Gen Physiol 98(2): 233‐248, 1991.
 251.Metzger JM, Moss RL. Myosin light chain 2 modulates calcium‐sensitive cross‐bridge transitions in vertebrate skeletal muscle. Biophys J 63(2):460‐468, 1992.
 252.Midde K, Rich R, Marandos P, Fudala R, Li A, Gryczynski I, Borejdo J. Comparison of orientation and rotational motion of skeletal muscle cross‐bridges containing phosphorylated and dephosphorylated myosin regulatory light chain. J Biol Chem 288(10):7012‐7023, 2013.
 253.Mijailovich SM, Kayser‐Herold O, Li X, Griffiths H, Geeves MA. Cooperative regulation of myosin‐S1 binding to actin filaments by a continuous flexible Tm‐Tn chain. Eur Biophys J 41(12): 1015‐1032, 2012.
 254.Miledi R, Parker I, Zhu PH. Calcium transients evoked by action potentials in frog twitch muscle fibres. J Physiol 333: 655‐679, 1982.
 255.Miledi R, Parker I, Zhu PH. Calcium transients in frog skeletal muscle fibres following conditioning stimuli. J Physiol 339: 223‐242, 1983.
 256.Millar NC, Homsher E. The effect of phosphate and calcium on force generation in glycerinated rabbit skeletal muscle fibers. A steady‐state and transient kinetic study. J Biol Chem 265(33): 20234‐20240, 1990.
 257.Millar NC, Homsher E. Kinetics of force generation and phosphate release in skinned rabbit soleus muscle fibers. Am J Physiol 262(5 Pt 1):C1239‐C1245, 1992.
 258.Miller MS, Bedrin NG, Callahan DM, Previs MJ, Jennings ME, Ades PA, Maughan DW, Palmer BM, Toth MJ. Age‐related slowing of myosin‐actin cross‐bridge kinetics is sex‐specific and predicts decrements in whole skeletal muscle performance in humans. J Appl Physiol 115: 1004‐1014, 2013.
 259.Miller MS, Farman GP, Braddock JM, Soto‐Adames FN, Irving TC, Vigoreaux JO, Maughan DW. Regulatory light chain phosphorylation and N‐terminal extension increase cross‐bridge binding and power output in Drosophila at in vivo myofilament lattice spacing. Biophys J 100(7): 1737‐1746, 2011.
 260.Miyamoto N, Fukutani A, Yanai T, Kawakami Y. Twitch potentiation after voluntary contraction and neuromuscular electrical stimulation at various frequencies in human quadriceps femoris. Muscle Nerve 45(1): 110‐115, 2012.
 261.Miyamoto N, Kanehisa H, Kawakami Y. Potentiation of maximal voluntary concentric torque in human quadriceps femoris. Med Sci Sports Exerc 44(9): 1738‐1746, 2012.
 262.Miyamoto N, Yanai T, Kawakami Y. Twitch potentiation induced by stimulated and voluntary isometric contractions at various torque levels in human knee extensor muscles. Muscle Nerve 43(3): 360‐366, 2011.
 263.Molloy JE, Burns JE, Kendrick‐Jones J, Tregear RT, White DC. Movement and force produced by a single myosin head. Nature 378: 209‐212, 1995.
 264.Moore JR, Dickinson MH, Vigoreaux JO, 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 indirect flight muscle. Biophys J 78(3):1431‐1440, 2000.
 265.Moore RL, Houston ME, Iwamoto GA, Stull JT. Phosphorylation of rabbit skeletal muscle myosin in situ. J Cell Physiol 125: 301‐305, 1985.
 266.Moore RL, Palmer BL, Williams SL, Tanabe H, Grange RW, Houston ME. Effect of temperature on myosin phosphorylation in mouse skeletal muscle. Am J Physiol 259: C432‐C438, 1990.
 267.Moore RL, Persechini A. Length‐dependence of isometric twitch tension potentiation and myosin phosphorylation in mouse skeletal muscle. J Cell Physiol 143(2): 257‐262, 1990.
 268.Moore RL, Stull JT. Myosin light chain phosphorylation in fast and slow skeletal muscles in situ. Am J Physiol Cell Physiol 247(5): C462‐C471, 1984.
 269.Moran AL, Nelson SA, Landisch RM, Warren GL, Lowe DA. Estradiol replacement reverses ovariectomy‐induced muscle contractile and myosin dysfunction in mature female mice. J Appl Physiol (1985). 102(4):1387‐1393, 2007.
 270.Morgan MS, Perry SV, Ottaway J. Myosin light chain phosphatase. Biochem J 157: 687‐697, 1976.
 271.Mrakovcic‐Zenic A, Reisler E. Light‐chain phosphorylation and cross‐bridge conformation in myosin from vertebrate skeletal muscle. Biochemistry 22: 525‐529, 1983.
 272.Myburgh KH, Franks‐Skiba K, Cooke R. Nucleotide turnover rate measured in fully relaxed rabbit skeletal muscle myofibrils. J Gen Physiol 106: 957‐973, 1995.
 273.Naber N, Cooke R, Pate E. Slow myosin ATP turnover in the super‐relaxed state in tarantula muscle. J Mol Biol 411(5):943‐950, 2011.
 274.Nelson CR, Debold EP, Fitts RH. Phosphate and acidosis act synergistically to depress peak power in rat muscle fibers. Am J Physiol Cell Physiol 307(10):C939‐C950, 2014.
 275.Nelson CR, Fitts RH. Effects of low cell pH and elevated inorganic phosphate on the pCa‐force relationship in single muscle fibers at near‐physiological temperatures. Am J Physiol Cell Physiol 306(7):C670‐C678, 2014.
 276.Nunnally MH, Stull JT. Mammalian skeletal muscle myosin light chain kinases. A comparison by antiserum cross‐reactivity. J Biol Chem 259(3):1776‐1780, 1984.
 277.Nyitrai M, Rossi R, Adamek N, Pellegrino MA, Bottinelli R, Geeves MA. What limits the velocity of fast‐skeletal muscle contraction in mammals? J Mol Biol 355(3): 432‐442, 2006.
 278.O'Leary DD, Hope K, Sale DG. Posttetanic potentiation of human dorsiflexors. J Appl Physiol 83(6): 2131‐2138, 1997.
 279.O'Leary DD, Hope K, Sale DG. Influence of gender on post‐tetanic potentiation in human dorsiflexors. Can J Physiol Pharmacol 76(7‐8): 772‐779, 1998.
 280.Orzechowski M, Li XE, Fischer S, Lehman W. An atomic model of the tropomyosin cable on F‐actin. Biophys J 107(3): 694‐699, 2014.
 281.Padrón R, Alamo L, Murgich J, Craig R. Towards an atomic model of the thick filaments of muscle. J Mol Biol 275(1):35‐41, 1998.
 282.Padrón R, Granados M, Alamo L, Guerrero JR, Craig R. Visualization of myosin helices in sections of rapidly frozen relaxed tarantula muscle. J Struct Biol 108(3):269‐276, 1992.
 283.Padron R, Panté N, Sosa H, Kendrick‐Jones J. X‐ray diffraction study of the structural changes accompanying phosphorylation of tarantula muscle. J Muscle Res Cell Motil 12(3): 235‐241, 1991.
 284.Palmer BM, Moore RL. Myosin light chain phosphorylation and tension potentiation in mouse skeletal muscle. Am J Physiol 257: C1012‐C1019, 1989.
 285.Pant K, Watt J, Greenberg M, Jones M, Szczesna‐Cordary D, Moore JR. Removal of the cardiac myosin regulatory light chain increases isometric force production. FASEB J 23(10): 3571‐3580, 2009.
 286.Pastoret C, Sebille A. Time course study of the isometric contractile properties of mdx mouse striated muscles. J Muscle Res Cell Motil 14(4): 423‐431, 1993.
 287.Patel JR, Diffee GM, Huang XP, Moss RL. Phosphorylation of myosin regulatory light chain eliminates force‐dependent changes in relaxation rates in skeletal muscle. Biophys J 74: 360‐368, 1998.
 288.Patel JR, Diffee GM, Moss RL. Myosin regulatory light chain phosphorylation modulates the Ca2+ dependence of the kinetics of tension development in skeletal muscle. Biophys J 70(5): 2333‐2340, 1996.
 289.Pemrick S. The phosphorylated L2 light chain of skeletal myosin is a modifier of the actomyosin ATPase. J Biol Chem 255(18): 8836‐8841, 1980.
 290.Perrie WT, Smillie LB, Perry SB. A phosphorylated light‐chain component of myosin from skeletal muscle. Biochem J 135(1): 151‐164, 1973.
 291.Persechini A, Stull JT, Cooke R. The effect of myosin phosphorylation on the contractile properties of skinned rabbit skeletal muscle fibers. J Biol Chem 260: 7951‐7954, 1985.
 292.Petrella RJ, Cunningham DA, Vandervoort AA, Paterson DH. Comparison of twitch potentiation in the gastrocnemius of young and elderly men. Eur J Appl Physiol Occup Physiol 58(4): 395‐399, 1989.
 293.Piazzesi G, Dolfi M, Brunello E, Fusi L, Reconditi M, Bianco P, Linari M, Lombardi V. The myofilament elasticity and its effect on kinetics of force generation by the myosin motor. Arch Biochem Biophys 553: 108‐116, 2014.
 294.Piazzesi G, Reconditi M, Linari M, Lucii L, Bianco P, Brunello E, Decostre V, Stewart A, Gore DB, Irving TC, Irving M, Lombardi V. Skeletal muscle performance determined by modulation of number of myosin motors rather than motor force or stroke size. Cell 131(4): 784‐795, 2007.
 295.Pinto A, Sánchez F, Alamo L, Padrón R. The myosin interacting‐heads motif is present in the relaxed thick filament of the striated muscle of scorpion. J Struct Biol 180(3): 469‐478, 2012.
 296.Pirani A, Vinogradova MV, Curmi PM, King WA, Fletterick RJ, Craig R, Tobacman LS, Xu C, Hatch V, Lehman W. An atomic model of the thin filament in the relaxed and Ca2+‐activated states. J Mol Biol 357(3):707‐717, 2006.
 297.Pires E, Perry SV, Thomas MA. Myosin light‐chain kinase, a new enzyme from striated muscle. FEBS Lett 41: 292‐296, 1974.
 298.Pires EM, Perry SV. Purification and properties of myosin light‐chain kinase from fast skeletal muscle. Biochem J 167(1): 137‐146, 1977.
 299.Preller M, Holmes KC. The myosin start‐of‐power stroke state and how actin binding drives the power stroke. Cytoskeleton (Hoboken) 70(10): 651‐660, 2013.
 300.Ramsey RW, Street SF. Muscle function as studied in single muscle fibres. Biol Symp 3: 9‐34, 1941.
 301.Rankin LL, Enoka RM, Volz KA, Stuart DG. Coexistence of twitch potentiation and tetanic force decline in rat hindlimb muscle. J Appl Physiol 65(6): 2687‐2699, 1988.
 302.Rassier DE, Herzog W. Effect of pH on the length dependent twitch potentiation in skeletal muscle. J Appl Physiol (1985) 92(3): 1293‐1299, 2002.
 303.Rassier DE, MacIntosh BR. Length dependence of staircase potentiation: Interactions with caffeine and dantrolene sodium. Can J Physiol Pharmacol 78(4): 350‐357, 2000.
 304.Rassier DE, Macintosh BR. Coexistence of potentiation and fatigue in skeletal muscle. Braz J Med Biol Res 33(5):499‐508, 2000.
 305.Rassier DE, MacIntosh BR. Sarcomere length‐dependence of activity‐dependent twitch potentiation in mouse skeletal muscle. BMC Physiol 2: 19, 2002.
 306.Rassier DE, MacIntosh BR, Herzog W. Length dependence of active force production in skeletal muscle. J Appl Physiol (1985) 86(5):1445‐1457, 1999.
 307.Rassier DE, Tubman LA, MacIntosh BR. Length‐dependent potentiation and myosin light chain phosphorylation in rat gastrocnemius muscle. Am J Physiol 273(1 Pt 1):C198‐204, 1997.
 308.Rassier DE, Tubman LA, MacIntosh BR. Caffeine and length dependence of staircase potentiation in skeletal muscle. Can J Physiol Pharmacol 76(10‐11): 975‐982, 1998.
 309.Rassier DE, Tubman LA, MacIntosh BR. Staircase in mammalian muscle without light chain phosphorylation. Braz J Med Biol Res 32(1): 121‐129, 1999.
 310.Rayment I, Holden HM. The three‐dimensional structure of a molecular motor. Trends Biochem Sci 19(3):129‐134, 1994.
 311.Rayment I, Holden HM, Whittaker M, Yohn CB, Lorenz M, Holmes KC, Milligan RA. Structure of the actin‐myosin complex and its implications for muscle contraction. Science 261: 58‐65, 1993.
 312.Rayment I, Rypniewski WR, Schmidt‐Base K, Smith R, Tomchick DR, Benning MM, Winkleman DA, Wesenberg G, Holden HM. Three‐dimensional structure of myosin subfragment‐1: A molecular motor. Science 261: 50‐58, 1993.
 313.Rayment I, Smith C, Yount RG. The active site of myosin. Annu Rev Physiol 58: 671‐702, 1996.
 314.Reconditi M, Linari M, Lucii L, Stewart A, Sun YB, Boesecke P, Narayanan T, Fischetti RF, Irving T, Piazzesi G, Irving M, Lombardi V. The myosin motor in muscle generates a smaller and slower working stroke at higher load. Nature 428: 578‐581, 2004.
 315.Regnier M, Rivera AJ, Wang CK, Bates MA, Chase PB, Gordon AM. Thin filament near‐neighbour regulatory unit interactions affect rabbit skeletal muscle steady‐state force‐Ca(2+) relations. J Physiol 540(Pt 2):485‐497, 2002.
 316.Rijkelijkhuizen JM, de Ruiter CJ, Huijing PA, de Haan A. Low‐frequency fatigue, post‐tetanic potentiation and their interaction at different muscle lengths following eccentric exercise. J Exp Biol 208(Pt 1):55‐63, 2005.
 317.Ritchie JM, Wilkie DR. The effect of previous stimulation on the active state of muscle. J Physiol 130: 488‐496, 1955.
 318.Ritz‐Gold CJ, Cooke R, Blumenthal DK, Stull JT. Light chain phosphorylation alters the conformation of skeletal muscle myosin. Biochem Biophys Res Commun 93(1): 209‐214, 1980.
 319.Robbins DW. Postactivation potentiation and its practical applicability: A brief review. J Strength Cond Res 19(2): 453‐458, 2005.
 320.Roche SM, Gumucio JP, Brooks SV, Mendias CL, Claflin DR. Measurement of maximum isometric force generated by permeabilized skeletal muscle fibers. J Vis Exp 2015(100), e52695.
 321.Ruff C, Furch M, Brenner B, Manstein DJ, Meyhöfer E. Single‐molecule tracking of myosins with genetically engineered amplifier domains. Nat Struct Biol 8(3):226‐229, 2001.
 322.Ryder JW, Lau KS, Kamm KE, Stull JT. Enhanced skeletal muscle contraction with myosin light chain phosphorylation by a calmodulin‐sensing kinase. J Biol Chem 282: 20447‐20454, 2007.
 323.Sale D. Postactivation potentiation: Role in performance. Br J Sports Med 38: 386‐387, 2004.
 324.Sale DG. Postactivation potentiation: Role in human performance. Exerc Sport Sci Rev 30(3):138‐143, 2002.
 325.Sandow, A. Excitation‐contraction coupling in muscular response. Yale J Biol Med 25(3): 176‐201, 1952.
 326.Schiaffino S, Reggiani C. Molecular diversity of myofibrillar proteins: Gene regulation and functional significance. Physiol Rev 76(2): 371‐423, 1996.
 327.Schiaffino S, Reggiani C. Fiber types in mammalian skeletal muscles. Physiol Rev 91(4): 1447‐1531, 2011.
 328.Scruggs SB, Hinken AC, Thawornkaiwong A, Robbins J, Walker LA, de Tombe PP, Geenen DL, Buttrick PM, Solaro RJ. Ablation of ventricular myosin regulatory light chain phosphorylation in mice causes cardiac dysfunction in situ and affects neighboring myofilament protein phosphorylation. J Biol Chem 284(8):5097‐5106, 2009.
 329.Scruggs SB, Solaro RJ. The significance of regulatory light chain phosphorylation in cardiac physiology. Arch Biochem Biophys 510(2):129‐134, 2011.
 330.Sherwood JJ, Waller GS, Warshaw DM, Lowey S. A point mutation in the regulatory light chain reduces the step size of skeletal muscle myosin. Proc Natl Acad Sci U S A 101(30): 10973‐10978, 2004.
 331.Siegman MJ, Mooers SU, Warren TB, Warshaw DM, Ikebe M, Butler TM. Comparison of the effects of 2,3‐butanedione monoxime on force production, myosin light chain phosphorylation and chemical energy usage in intact and permeabilized smooth and skeletal muscles. J Muscle Res Cell Motil 15(4): 457‐472, 1994.
 332.Siemankowski RF, Wiseman MO, White HD. ADP dissociation from actomyosin subfragment 1 is sufficiently slow to limit the unloaded shortening velocity in vertebrate muscle. Proc Natl Acad Sci U S A 82(3): 658‐662, 1985.
 333.Smith CB, Allen MD, Rice CL. Voluntary rate of torque development is impaired after a voluntary versus tetanic conditioning contraction. Muscle Nerve 49(2): 218‐224, 2014.
 334.Smith CB, Cheng AJ, Rice CL. Potentiation of the triceps brachii during voluntary submaximal contractions. Muscle Nerve 43(6): 859‐865, 2011.
 335.Smith JC, Fry AC. Effects of a ten‐second maximum voluntary contraction on regulatory myosin light‐chain phosphorylation and dynamic performance measures. J Strength Cond Res 21(1): 73‐76, 2007.
 336.Smith IC, Gittings W, Bloemberg D, Huang J, Quadrialtero J, Tupling AR, Vandenboom R. Potentiation in mouse lumbrical muscle without myosin light chain phosphorylation: Is resting calcium responsible? J Gen Physiol 141(3): 297‐308, 2013.
 337.Smith IC, Vandenboom R, Tupling R. Juxtaposition of the changes in intracellular calcium and force during staircase potentiation at 30 and 37°C. J Gen Physiol 144(6): 561‐570, 2014.
 338.Spangenburg EE, Geiger PC, Leinwand LA, Lowe DA. Regulation of physiological and metabolic function of muscle by female sex steroids. Med Sci Sports Exerc 44(9):1653‐1662, 2012.
 339.Spudich JA. Biochemistry. Molecular motors, beauty in complexity. Science 331(6021): 1143‐1144, 2011.
 340.Spudich JA. Molecular motors: Forty years of interdisciplinary research. Mol Biol Cell 22(21): 3936‐3939, 2011.
 341.Standaert FG. The mechanisms of post‐tetanic potentiation in cat soleus and gastrocnemius muscles. J Gen Physiol 47: 987‐1001, 1964.
 342.Stephenson GM, Stephenson DG. Endogenous MLC2 phosphorylation and Ca(2+)‐activated force in mechanically skinned skeletal muscle fibres of the rat. Pflugers Arch 424(1): 30‐38, 1993.
 343.Stevens L, Bastide B, Bozzo C, Mounier Y. Hybrid fibres under slow‐to‐fast transformations: Expression is of myosin heavy and light chains in rat soleus muscle. Pflugers Arch 448(5): 507‐514, 2004.
 344.Stevens L, Firinga C, Gohlsch B, Bastide B, Mounier Y, Pette D. Effects of unweighting and clenbuterol on myosin light and heavy chains in fast and slow muscles of rat. Am J Physiol (Cell Physiol) 279(5): 1558‐1563, 2000.
 345.Stewart M, Franks‐Skiba K, Cooke R. Myosin regulatory light chain phosphorylation inhibits shortening velocities of skeletal muscle fibers in the presence of the myosin inhibitor blebbistatin. J Musc Res Cell Motil 30: 17‐27, 2009.
 346.Stewart MA, Franks‐Skiba K, Chen S, Cooke R. Myosin ATP turnover rate is a mechanism involved in thermogenesis in resting skeletal muscle fibers. Proc Natl Acad Sci U S A 107(1):430‐435, 2010.
 347.Stuart DS, Lingley MD, Grange RW, Houston ME. Myosin light chain phosphorylation and contractile performance of human skeletal muscle. Can J Physiol Pharmacol 66(1): 49‐54, 1988.
 348.Stull JT, Kamm C, Vandenboom R. Myosin light chain kinase and the role of myosin light chain phosphorylation in skeletal muscle. Arch Biochem Biophys 510: 120‐128, 2011.
 349.Sulbarán G, Biasutto A, Alamo L, Riggs C, Pinto A, Méndez F, Craig R, Padrón R. Different head environments in tarantula thick filaments support a cooperative activation process. Biophys J 105(9): 2114‐2122, 2013.
 350.Sun YB, Lou F, Edman KA. The relationship between the intracellular Ca2+ transient and the isometric twitch force in frog muscle fibres. Exp Physiol 81(5):711‐724, 1996.
 351.Sweeney HL, Bowman BF, Stull JT. Myosin light chain phosphorylation in vertebrate striated muscle: Regulation and function. Am J Physiol 264: 1085‐1095, 1993.
 352.Sweeney HL, Houdusse A. Structural and functional insights into the myosin motor mechanism. Annu Rev Biophys 39: 539‐557, 2010.
 353.Sweeney HL, Kushmerick MJ. Myosin phosphorylation in permeabilized rabbit psoas fibers. Am J Physiol 249(3 Pt 1): C362‐C365. 1985.
 354.Sweeney HL, Stull JT. Phosphorylation of myosin in permeabilized mammalian cardiac and skeletal muscle cells. Am J Physiol 250(4 Pt 1):657‐660, 1986.
 355.Sweeney HL, Stull JT. Alteration of cross‐bridge kinetics by myosin light chain phosphorylation in rabbit skeletal muscle: Implications for regulation of actin‐myosin interaction. Proc Natl Acad Sci U S A 87: 414‐418, 1990.
 356.Sweeney HL, Yang Z, Zhi G, Stull JT, Trybus KM. Charge replacement near the phosphorylatable serine of the myosin regulatory light chain mimics aspects of phosphorylation. Proc Natl Acad Sci U S A 91: 1490‐1494, 1994.
 357.Szczesna D. Regulatory light chains of striated muscle myosin. Structure, function and malfunction. Curr Drug Targets Cardiovasc Haematol Disord 3(2):187‐197, 2003.
 358.Szczesna D, Zhao J, Jones M, Zhi G, Stull J, Potter JD. Phosphorylation of the regulatory light chains of myosin affects Ca2+ sensitivity of skeletal muscle contraction. J Appl Physiol 92(4): 1661‐1670, 2002.
 359.Szczesna D, Zhao J, Potter JD. The regulatory light chains of myosin modulate cross‐bridge cycling in skeletal muscle. J Biol Chem 271(9):5246‐5250, 1996.
 360.Szczesna‐Cordary D, Jones M, Moore JR, Watt J, Kerrick WG, Xu Y, Wang Y, Wagg C, Lopaschuk GD. Myosin regulatory light chain E22K mutation results in decreased cardiac intracellular calcium and force transients. FASEB J 21(14): 3974‐3985, 2007.
 361.Szent‐Gyorki A. Free‐energy relations and contraction of actomyosin. Biol Bull 96(2): 140‐161, 1949.
 362.Tanner BC, Farman GP, Irving TC, Maughan DW, Palmer BM, Miller MS. Thick‐to‐thin filament surface distance modulates cross‐bridge kinetics in Drosophila flight muscle. Biophys J 103(6):1275‐1284, 2012.
 363.Thomas CK, Johansson RS, Bigland‐Ritchie B. EMG changes in human thenar motor units with force potentiation and fatigue. J Neurophysiol 95(3):1518‐1526, 2006.
 364.Tiidus PM, Lowe DA, Brown M. Estrogen replacement and skeletal muscle: Mechanisms and population health. J Appl Physiol (1985) 115(5):569‐578, 2013.
 365.Tillin NA, Bishop D. Factors modulating post‐activation potentiation and its effect on performance of subsequent explosive activities. Sports Med 39(2):147‐166.
 366.Tohtong R, Yamashita H, Graham M, Haeberle J, Simcox A, Maughan D. Impairment of muscle function caused by mutations of phosphorylation sites in myosin regulatory light chain. Nature 374(6523): 650‐653, 1995.
 367.Trybus KM. Role of myosin light chains. J Musc Res Cell Motil 15(6): 587‐594, 1994.
 368.Tsianos GA, Rustin C, Loeb GE. Mammalian muscle model for predicting force and energetics during physiological behaviors. IEEE Trans Neural Syst Rehabil Eng, 20(2), 117‐133, 2012.
 369.Tubman LA, MacIntosh BR, Maki WA. Myosin light chain phosphorylation and posttetanic potentiation in fatigued skeletal muscle. Pflugers Arch 431(6): 882‐887, 1996.
 370.Tubman LA, Rassier DE, MacIntosh BR. Absence of myosin light chain phosphorylation and twitch potentiation in atrophied skeletal muscle. Can J Physiol Pharmacol 74(6): 723‐728, 1996.
 371.Tubman LA, Rassier DE, MacIntosh BR. Attenuation of myosin light chain phosphorylation and posttetanic potentiation in atrophied skeletal muscle. Pflugers Arch 434(6): 848‐851, 1997.
 372.Tupling AR. The sarcoplasmic reticulum in muscle fatigue and disease: Role of the sarco(endo)plasmic reticulum Ca2+‐ATPase. Can J Appl Physiol 29(3):308‐329, 2004.
 373.Tupling AR. Excitation‐contraction coupling. Encyclopedia Neurosci 2009.
 374.Tyska MJ, Dupuis DE, Guilford WH, Patlak JB, Waller GS, Trybus KM, Warshaw DM, Lowey S. Two heads of myosin are better than one for generating force and motion. Proc Natl Acad Sci U S A 96: 4402‐4407, 1999.
 375.Uyeda TQ, Abramson PD, Spudich JA. The neck region of the myosin motor domain acts as a lever to generate movement. Proc Natl Acad Sci U S A 93(9): 4459‐4464, 1996.
 376.Vale RD, Milligan RA. The way things move: Looking under the hood of molecular motor proteins. Science 288(5463): 88‐95, 2000.
 377.VanBuren P, Waller GS, Harris DE, Trybus KM, Warshaw DM, Lowey S. The essential light chain is required for full force production by skeletal muscle myosin. Proc Natl Acad Sci U S A 91(26): 12403‐12407, 1994.
 378.Vandenboom R. The myofibrillar complex and fatigue: A review. Can J Appl Physiol 29(3): 330‐356, 2004.
 379.Vandenboom R, Claflin DR, Julian FJ. The effects of rapid shortening on rate of force regeneration and myoplasmic [Ca2+] in intact frog skeletal muscle fibers. J Physiol (London) 511: 171‐180, 1998.
 380.Vandenboom R, Gittings W, Smith IC, Grange RW, Stull JT. Myosin phosphorylation and force potentiation in skeletal muscle: Evidence from animal models. J Muscle Res Cell Motil 34(5‐6): 317‐332, 2013.
 381.Vandenboom R, Grange RW, Houston ME. Threshold for force potentiation associated with skeletal myosin phosphorylation. Am J Physiol 265: 1456‐1462, 1993.
 382.Vandenboom R, Grange RW, Houston ME. Myosin phosphorylation enhances rate of force development in fast‐twitch skeletal muscle. Am J Physiol 268: 596‐603, 1995.
 383.Vandenboom R, Hannon JD, Sieck GC. Isotonic force modulates force redevelopment rate of intact frog skeletal muscle fibers: Evidence for cross‐bridge induced thin filament activation. J Physiol 543: 555‐566, 2002.
 384.Vandenboom R, Houston ME. Phosphorylation of myosin and twitch potentiation in fatigued skeletal muscle. Can J Physiol Pharmacol 74(12): 1315‐1321, 1996.
 385.Vandenboom R, Xeni J, Bestic M, Houston ME. Increased force development rates of fatigued skeletal muscle are graded to myosin light chain phosphate content. Am J Physiol 272: 1980‐1984, 1997.
 386.Vandervoort AA, Quinlan J, McComas AJ. Twitch potentiation after voluntary contraction. Exp Neurol 81(1): 141‐152, 1983.
 387.Vergara JL, Rapoprot SI, Nassar‐Gentina V. Fatigue and posttetanic potentiation in single muscle fibers of the frog. Am J Physiol 232(5): 185‐190, 1977.
 388.Vibert P, Craig R. Structural changes that occur in scallop myosin filaments upon activation. J Cell Biol 101(3):830‐837, 1985.
 389.von der Ecken J, Müller M, Lehman W, Manstein DJ, Penczek PA, Raunser S. Structure of the F‐actin tropomyosin complex. Nature 519(7541):114‐117, 2015.
 390.Wahr PA, Johnson JD, Rall JA. Determinants of relaxation rate in skinned frog skeletal muscle fibers. Am J Physiol 274(6 Pt 1):C1608‐C1615, 1998.
 391.Walcott S, Warshaw DM, Debold EP. Mechanical coupling between myosin molecules causes differences between ensemble and single‐molecule measurements. Biophys J 103(3): 501‐510, 2012.
 392.Walker SM. Action potentials in rat muscle with twitch tension potentiated by KCI treatment, adrenalectomy, tetanus and treppe. Am J Physiol 154(1): 63‐72, 1948.
 393.Walklate J, Ujfalusi Z, Geeves MA. Myosin isoforms and the mechanochemical cross‐bridge cycle. J Exp Biol 219(2): 168‐174, 2016.
 394.Waller GS, Ouyang G, Swafford J, Vibert P, Lowey S. A minimal motor domain from chicken skeletal muscle myosin. J Biol Chem 270: 15348‐15352, 1995.
 395.Wang Y, Szczesna‐Cordary D, Craig R, Diaz‐Perez Z, Guzman G, Miller T, Potter JD. Fast skeletal muscle regulatory light chain is required for fast and slow skeletal muscle development. FASEB J 21(9): 2205‐2214, 2007.
 396.Warshaw DM. Lever arms and necks: A common mechanistic theme across the myosin superfamily. J Muscle Res Cell Motil 25(6): 467‐474, 2004.
 397.Weiss S, Rossi R, Pellegrini MA, Geeves MA. Differing ADP release rates from myosin heavy chain isoforms define the shortening velocity of skeletal muscle fibers. J Biol Chem 276: 45902‐45908, 2001.
 398.Westerblad H, Allen DG. Changes of myoplasmic calcium concentration during fatigue in single mouse muscle fibers. J Gen Physiol 98(3):615‐635, 1991.
 399.Westerblad H, Allen DG. Relaxation, [Ca2+] and [Mg2+] during prolonged tetanic stimulation of intact, single fibres from mouse skeletal muscle. J Physiol 480: 31‐43, 1994.
 400.Westwood SA, Hudlicka O, Perry SV. Phosphorylation in vivo of the P light chain of myosin in rabbit fast and slow skeletal muscles. Biochem J 218(3): 841‐847, 1984.
 401.Wilson JM, Duncan NM, Marin PJ, Brown LE, Loenneke JP, Wilson SM, Jo E, Lowery RP, Ugrinowitsch C. Meta‐analysis of postactivation potentiation and power: Effects of conditioning activity, volume, gender, rest periods, and training status. J Strength Cond Res 27(3):854‐859, 2013.
 402.Wilson C, Naber N, Pate E, Cooke R. The myosin inhibitor blebbistatin stabilizes the super‐relaxed state in skeletal muscle. Biophys J 107(7):1637‐1646, 2014.
 403.Woodhead JL, Zhao FQ, Craig R, Egelman EH, Alamo L, Padron R. Atomic model of a myosin filament in the relaxed state. Nature 436(7054): 1195‐1199, 2005.
 404.Woodhead JL, Zhao FQ, Craig R. Structural basis of the relaxed state of a Ca2+‐regulated myosin filament and its evolutionary implications. Proc Natl Acad Sci U S A 110(21): 8561‐8566, 2013.
 405.Woods JJ, Furbush F, Bigland‐Ritchie B. Evidence for a fatigue‐induced reflex inhibition of motoneuron firing rates. J Neurophysiol 58(1): 125‐137, 1987.
 406.Xeni J, Gittings W, Caterini D, Huang J, Houston ME, Grange RW, Vandenboom R. Myosin light chain phosphorylation and potentiation of dynamic function in mouse fast muscle. Pflugers Archiv 362: 349‐358, 2011.
 407.Yanagida T, Iwaki M, Ishii Y. Single molecule measurements and molecular motors. Philos Trans R Soc Lond B Biol Sci 363: 2123‐2134, 2008.
 408.Yanagida T, Iwane AH. A large step for myosin. Proc Natl Acad Sci U S A 97(17): 9357‐9359, 2000.
 409.Yang Z, Stull JT, Levine RJ, Sweeney HL. Changes in interfilament spacing mimic the effects of myosin regulatory light chain phosphorylation in rabbit psoas fibers. J Struct Biol 122: 139‐148, 1998.
 410.Yang Z, Sweeney HL. Restoration of phosphorylation‐dependent regulation to the skeletal muscle myosin regulatory light chain. J Biol Chem 270(42):24646‐24649, 1995.
 411.Zhao FQ, Craig R, Woodhead JL. Head‐head interaction characterizes the relaxed state of Limulus muscle myosin filaments. J Mol Biol 385(2): 423‐431, 2009.
 412.Zhao FQ, Padrón R, Craig R. Blebbistatin stabilizes the helical order of myosin filaments by promoting the switch 2 closed state. Biophys J 95(7):3322‐3329, 2008.
 413.Zhi G, Ryder JW, Huang J, Ding P, Chen Y, Zhao Y, Kamm KE, Stull JT. Myosin light chain kinase and myosin phosphorylation effect frequency‐dependent potentiation of skeletal muscle contraction. Proc Natl Acad Sci U S A 102: 17519‐17524, 2005.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Rene Vandenboom. Modulation of Skeletal Muscle Contraction by Myosin Phosphorylation. Compr Physiol 2016, 7: 171-212. doi: 10.1002/cphy.c150044