Abstract
In the present work, we investigate whether changes in excitation–contraction (EC) coupling mode occur in skeletal muscles from ageing mammals by examining the dependence of EC coupling on extracellular Ca2+. Single intact muscle fibres from flexor digitorum brevis muscles from young (2–6 months) and old (23–30 months) mice were subjected to tetanic contractile protocols in the presence and absence of external Ca2+. Contractile experiments in the absence of external Ca2+ show that about half of muscle fibres from old mice are dependent upon external Ca2+ for maintaining maximal tetanic force output, while young fibres are not. Decreased force in the absence of external Ca2+ was not due to changes in charge movement as revealed by whole-cell patch-clamp experiments. Ca2+ transients, measured by fluo-4 fluorescence, declined in voltage-clamped fibres from old mice in the absence of external Ca2+. Similarly, Ca2+ transients declined in parallel with tetanic contractile force in single intact fibres. Examination of inward Ca2+ current and of mRNA and protein assays suggest that these changes in EC coupling mode are not due to shifts in dihydropyridine receptor (DHPR) and/or ryanodine receptor (RyR) isoforms. These results indicate that a change in EC coupling mode occurs in a population of fibres in ageing skeletal muscle, and is responsible for the age-related dependence on extracellular Ca2+.
The role of external Ca2+ ions in mammalian skeletal muscle contraction is not completely understood. Controversy exists over whether external Ca2+ is necessary for adult mammalian skeletal muscle excitation–contraction (EC) coupling. EC coupling is the series of ionic and molecular events whereby membrane depolarization is transmitted to contraction in striated muscles. Two protein complexes particularly important in EC coupling are the dihydropyridine receptor (DHPR) and the ryanodine receptor (RyR). These two proteins align during development at the junction of the t-tubule with the terminal cisternae of the sarcoplasmic reticulum (SR) (Flucher & Franzini-Armstrong, 1996) – a collection of structures termed the triad. The α1 subunit of the DHPR is a voltage-gated L-type Ca2+ channel in the t-tubule membrane that also serves as the voltage sensor (Rios & Brum, 1987). The DHPR is responsible for activating Ca2+ release from the SR via the RyR (Schneider & Chandler, 1973; Marty et al. 1994) into the cytoplasm to trigger muscular contraction. The DHPR and RyR isoforms in cardiac and skeletal muscle are different, with cardiac muscle expressing the DHPR α1C subunit (Mikami et al. 1989) and RyR2 (Nakai et al. 1990), and skeletal muscle expressing the DHPR α1S subunit (Tanabe et al. 1988, 1990a) and RyR1 (Takeshima et al. 1989). In cardiac muscle, RyR2 is activated to release Ca2+ from the SR by Ca2+ influx through the DHPR α1C (Fabiato, 1985; Nabauer et al. 1989). This Ca2+ influx is necessary for cardiac EC coupling, as indicated by the complete elimination of contraction and Ca2+ release transients in cells where external Ca2+ has been removed or Ca2+ entry has been blocked (Tanabe et al. 1990b; García et al. 1994). By contrast, RyR1 in skeletal muscle does not depend on Ca2+ influx through the DHPR α1S to activate SR Ca2+ release and skeletal muscle contraction. Experiments in which external Ca2+ has been removed from (Armstrong et al. 1972; Dulhunty & Gage, 1988; Tanabe et al. 1990b) or calcium channel blockers are added to (González-Serratos et al. 1982; Dulhunty & Gage, 1988) the bathing medium show that EC coupling and contraction persist in skeletal muscle cells. These experiments suggest a mechanical coupling between the α1S and RyR1 (Rios & Brum, 1987; Tanabe et al. 1990a).
While adult skeletal muscle does not seem to depend on external Ca2+ for EC coupling, developing skeletal muscle does (Dangain & Neering, 1991; Cognard et al. 1992; Pereon et al. 1993), exhibiting some aspects of cardiac-like EC coupling. This is presumably due to the temporary expression of the cardiac DHPR α1C during development (Chaudhari & Beam, 1993). As muscle cells develop, the expression of α1S increases while α1C decreases (Chaudhari & Beam, 1993), shifting to a more skeletal type EC coupling (Cognard et al. 1992). Ca2+-dependent EC coupling in developing muscle could also be due to transient expression of RyR3 (Flucher et al. 1999; Chun et al. 2003) or RyR1 splice variants (Futatsugi et al. 1995) or a combination of all these different isoforms, together allowing the RyR to be responsive to Ca2+ influx through the DHPR at normal cytoplasmic Mg2+ concentrations. Regenerating skeletal muscle, which goes through many stages similar to developing skeletal muscle, is at least partially dependent on external Ca2+ for effective EC coupling and contraction (Louboutin et al. 1995, 1996; Pereon et al. 1997a). This is also presumably due to the transient expression of α1C during the regeneration period (Pereon et al. 1997a). Over time, expression of α1C diminishes while α1S increases, decreasing the dependence of regenerating muscle on external Ca2+ for effective EC coupling (Pereon et al. 1997a).
Prompted by the changes in EC coupling mode during development and muscle injury in adult rodents described above, we tested the hypothesis that skeletal muscle fibres from senescent mice undergo a ‘shift’ towards cardiac-like EC coupling (dependence on external Ca2+). Therefore, in the present study, we examined the effects of Ca2+-free solution on the contractile properties of fibres from flexor digitorum brevis (FDB) muscle in young and old mice, and whether differences were due to shifts in DHPR and/or RyR isoforms. To that end, we employed a combination of contraction and intracellular Ca2+ recordings in single intact muscle fibres, and measurements of Ca2+ current, charge movement and intracellular Ca2+ in patch-clamped fibres, together with molecular techniques to address this issue.
Methods
Animals
Flexor digitorum brevis (FDB) muscles were dissected from 3- to 6-month-old (young; n = 30) and 22- to 24-month-old FVB (our colony) or 21 month-old DBA or 26–27 month-old CB6F1 (Harlan-NIA colonies) (old; n = 23) mice. FVB and DBA strains have been successfully used in the study of ageing muscle in previous work from our laboratory (Renganathan et al. 1998; González et al. 2000, 2003) and the CB6F1 strain has been used by other investigators (Miller et al. 1997). The availability of the oldest animals determined the inclusion of three mouse strains in the present work. The findings reported in this study are independent of the mouse strain (see below). Animals were housed at Wake Forest University School of Medicine (WFUSM). Mice were killed by cervical dislocation. Animal handling and procedures were approved by the Animal Care and Use Committee of WFUSM.
Single intact fibre contraction experiments
The technique for dissecting single intact fibres followed procedures previously described (Lannergren & Westerblad, 1987; González et al. 2000). Two physiological buffering solutions were used for contraction experiments: Ca2+-containing (recording) and Ca2+-free solutions. The recording solution consisted of (mm): NaCl 121, KCl 5, CaCl2 1.8, MgCl2 0.5, NaH2PO4 0.4, NaHCO3 24, and glucose 5.5. The Ca2+-free solution was identical to the recording solution, except that the MgCl2 concentration was increased to 2.3 mm in place of CaCl2. Both solutions also contained 10−5 g ml−1 of d-tubocurarine chloride and were bubbled continuously with a mixture of 5% CO2–95% O2 to achieve a pH of 7.4. Fibres were stimulated by an electrical field generated between two parallel silver electrodes connected to a Grass S48 stimulator (Astro-Medical, Inc., West Warwick, RI, USA). Fibre length was adjusted until maximum force was elicited by a single twitch contraction (LO) under isometric conditions. Suprathreshold square wave pulses of 0.5 ms duration were delivered to elicit twitch contractions. Tetanic contractions were elicited with 0.5 ms square wave pulses delivered in 300-ms trains. Frequency was increased until maximum force was attained. All subsequent tetanic contractions were elicited with the frequency that elicited maximal force, as described (González et al. 2003). All experiments were carried out at room temperature (20–21°C).
The first set of contraction experiments in single intact FDB fibres consisted of two prolonged contractile sequences: the first (reference trial) to assess any degree of force decline for that fibre, the second (test trial) to assess the effects of the Ca2+-free solution on tetanic force. The experimental sequence was as follows. (1) reference trial consisting of repeated maximal tetanic contractions, set at a 10-s interval, persisting for 25 min, in recording solution (total of 150 contractions). (2) Ten min of rest in recording solution. (3) Test trial repeating the reference trial contraction procedure, replacing recording solution with Ca2+-free solution from min 5 until min 15 of the 25-min procedure (contractions 31–90 of 150). (4) Up to 15 min of rest in recording solution, with single tetanic contractions at 5-min intervals. Force was normalized to baseline values for each fibre. Baseline was defined as an average of the five contractions immediately preceding the start of Ca2+-free solution flow. Force decline was assessed during min 5–15 of the reference trial, the same time frame during which Ca2+-free solution was perfused during the test trial. Only those experiments in which the fibre force recovered to at least 90% of baseline force were included for analysis.
The second set of contraction experiments in single intact FDB fibres also consisted of two contractile sequences: the first to assess the effects of the Ca2+-free solution on tetanic force, the second to record simultaneous force and Ca2+ fluorescence (see below). The experimental sequence was as follows. (1) Single tetanic contractions elicited at 2-min intervals for 20 min. From 0 to 10 min, Ca2+-free solution was perfused; from 10 min, recording solution was perfused. (2) Incubation for 40–50 min with the acetoxymethyl ester of fluo-4 (fluo-4 AM; 5 μm), added as a 1 mm stock in DMSO, followed by washout. (3) Readjustment of LO, if necessary, and focusing of the laser scanning confocal microscope (Noran, OZ, Middleton, WI, USA). (4) A repeat of step (1) with simultaneous recording of force and fluorescence. Only those experiments in which the fibre force recovered to at least 90% of baseline force were included for analysis.
Data were acquired with a personal computer, an A–D converter (Digidata 1200, Axon Instruments, Union City, CA, USA) and pCLAMP software (Axon Instruments). The pulse waveform and the contraction signal were acquired and digitized together and stored for later analysis.
Charge movement, Ca2+ current and intracellular Ca2+ recordings
Single dissociated fibres were obtained from FDB muscles. FDB muscles were dissected in a solution containing (mm): caesium aspartate 155, magnesium aspartate 5 and Hepes 10; pH adjusted to 7.4 with CsOH (Beam & Franzini Armstrong, 1997). Muscles were treated for 3 h with 2 mg ml−1 collagenase in a shaking bath at 37°C. Then, fibres were dissociated with Pasteur pipettes of different tip sizes. Fibres were transferred to a small flow-through chamber set on an inverted microscope stage (Axiovert S100 2TV, Zeiss, Göttingen, Germany). Fibres were continuously perfused with external solution (see below) using a push–pull pump (WPI, Sarasota, FL, USA). Muscle fibres were voltage-clamped using an Axopatch 200B amplifier (Axon Instruments, Union City, CA, USA) in the whole-cell configuration of the patch-clamp technique, following published procedures (Wang et al. 2000, 2002). The voltage clamp of short FDB fibres utilizing the whole-cell configuration of the patch-clamp technique together with low resistance pipette tips allows for a higher seal resistance, better control of the space clamp, and more prolonged and stable recordings of membrane currents and intracellular Ca2+ compared to previous procedures applied to longer muscle fibres (for discussion see Wang et al. 1999). Patch pipettes were pulled from borosilicate glass using a Flaming Brown micropipette puller (P97, Sutter Instrument Co., Novato, CA, USA) and then fire-polished to obtain an electrode resistance ranging from 450 to 650 kΩ. The pipette was filled with the following solution (mm): caesium aspartate 145, EGTA 10, MgCl2 5 and Hepes 10; pH adjusted to 7.4 with CsOH.
Inward Ca2+ currents were evoked with 350-ms depolarizing pulses from the holding potential (−80 mV) to command potentials ranging from −70 to 50 mV with 10 mV intervals. The Ca2+ current–voltage relationship was fitted to the following equation:
where Gmax is the maximum conductance, V is the membrane potential, Vr is the reversal potential, V½ is the half-activation potential, z is the valence of the mobile charge, F is the Faraday constant, R is the gas constant and T is the absolute temperature (Wang et al. 1999). The rising phase of the calcium current, elicited in response to command pulses of −50 to 50 mV, was adequately fitted to a single-exponential function as described (Delbono, 1992). The activation time constant (τa) was determined by fitting the current trace to an exponential function from the beginning of the pulse to the point where the current reached a steady-state level (∼30 ms). The data were fitted to a single barrier Eyring model assuming both first- and second-order terms to be present in the rate constant–voltage function:
where τmax is the activation time constant at the voltage of equal charge distribution, V½ is the half activation potential, and K = kT/aez (where k is the steepness of the curve, T is the temperature, a and b are constants that express the dependence of the rates on the first and second powers of the electric potential, e is the elemental charge, and z is valence of the mobile charge) (Brum & Rios, 1987). Double-pulse experiments were also performed to examine Ca2+-dependent inactivation of Ca2+ current. Inward Ca2+ and Ba2+ currents were evoked with 300-ms depolarizing pulses from the holding potential (−80 mV) to command potentials ranging from −70 to 60 mV with 10 mV intervals (first pulse). The second depolarizing pulse (also 300-ms duration) was held at a constant 20 mV potential, and followed the first pulse by a 100-ms interval. Relative current inactivation was calculated as current in the second pulse relative to maximum current attained in the second pulse (ICa/ICa,max). Ba2+ currents were measured in an external solution identical to the Ca2+-containing external solution, except that the 2 mm CaCl2 was replaced by 2 mm BaCl2. Data were fitted according to a Boltzmann distribution using the following equation:
where I is current, Imax is the maximum current amplitude, V½ is the half-inactivation potential, Vm is membrane voltage, k is the steepness of the curve, and A is the amplitude factor.
Intramembrane charge movements were elicited with 25-ms depolarizing pulses from the holding potential (−80 mV) to command potentials ranging from −70 to 50 mV with 10 mV intervals. Intramembrane charge movement was calculated as the integral of the current in response to depolarizing pulses (charge on, Qon) and is expressed per membrane capacitance (coulombs per farad). Charge movement was recorded in two different external solutions to assess the influence of external Ca2+ ions. Fibres were patch-clamped using a Ca2+-containing external solution containing (mm): tetraethlyammonium hydroxide (TEA-OH) 150, CaSO4 2, MgSO4 2, 3,4-diaminopyridine (DAP) 2, Hepes 5 and tetrodotoxin 0.001; pH adjusted to 7.4 with CH4SO3. Ca2+ current was blocked with the addition of 0.5 mm CdCl2 and 0.3 mm LaCl3 to record charge movement in the presence of external Ca2+ ions. Ca2+-free external solution containing (mm): TEA-OH 150, MgSO4 4, DAP 2, Hepes 5 and tetrodotoxin 0.001 (pH adjusted to 7.4 with CH4SO3) was perfused to record charge movement in the absence of external Ca2+ ions. Mean data points were fitted to a Boltzmann distribution of the form:
where Qmax is the maximum charge, Vm is the membrane potential, VQ½ is the charge movement half-activation potential, and k is the steepness of the curve (Zheng et al. 2002a).
For intracellular Ca2+ recordings, FDB fibres were loaded with 5 μm fluo-4 AM (Molecular Probes, Eugene, OR, USA) for 15–20 min. After washout with the Ca2+-containing solution, fibres were patch- and voltage-clamped (see above). In these experiments, the pipette solution was identical to that used for charge movement recordings, except that the EGTA concentration was reduced to 0.2 mm. Ca2+ transients were elicited with 20-ms depolarizing pulses from the holding potential (−80 mV) to command potentials ranging from −40 to 30 mV. Ca2+ transients were recorded in both the Ca2+-containing and the Ca2+-free solutions used for charge movement recordings as described above.
Fluorescent image recording
For Ca2+ fluorescence recordings in contraction and patch-clamp experiments, the fibres were illuminated with a laser beam at 488 nm wavelength. The beam passed through an OZ Scan module (Noran Instruments) and through a 20 X Fluar objective (Zeiss) before reaching the fibre. Emitted fluorescence was collected by the objective and directed to the OZ scan module, in a non-slit mode, through the emission filter at 525 nm wavelength before being collected by a photomultiplier tube and digitized. Hardware control, image acquisition and processing were performed with Intervision Software (Noran Instruments) run on a Silicon Graphics O2 Workstation (Mountain View, CA, USA). Sequences of 150 images at 8 ms intervals were collected for each contraction. For data analysis, several regions of interest (ROIs) were selected for each cell and the maximum fluorescence deflection was used. Fluorescence data are reported as a percentage change in fluorescence normalized to basal fluorescence (%ΔF/F) (Finch & Augustine, 1998).
Ribonuclease protection assay
Ribonuclease protection assay (RPA) technique followed procedures previously described (Zheng et al. 2001). Briefly, total RNA was extracted from mouse skeletal muscle and heart by using TRI reagent (Molecular Research Center, Inc., Cincinnati, OH, USA). The probe for mouse DHPR α1C is generated by RT-PCR using sense primer 5′TACGGACTTCTCTTCCACCC-3′, anti-sense primer 5′TCCCTCCTAGAGCATTGGCC-3′, corresponding to mRNA sequence (Accession number NM_009781) from 1707 to 1845. Its PCR fragment is cloned in TA-easy vector (Promega, WI, USA) and confirmed by DNA sequencing, the construct is linearized by Sal I digestion and purified. In vitro transcription was performed using the Maxiscript kit from Ambion (Austin, TX, USA). Briefly, the in vitro transcription reaction was performed by mixing linearized plasmid, 10 × transcription buffer, 10 mm ATP/CTP/GTP/UTP, [32P]UTP, T7 RNA polymerase plus ribonuclease inhibitor. The transcription was followed by removal of template DNA with DNase and gel purification of the probe. The cRNA probe is 229 bp (139 plus 90-bp vector fragment), and protected fragment after RNase digestion in RPA is 139 bp. RPA was performed using RPA II reagents (Ambion). Total RNA 25 μg was hybridized with labelled probes at 56°C overnight. This was followed by RNase digestion of non-hybridized probes and sample RNA, separation and detection of the protected fragments in a 10% urea denaturing polyacrylamide gel and gel exposure to an X-ray film.
DHPR protein detection
Immunoprecipitation
Mouse skeletal and heart muscles were prepared for immunoprecipitation as previously described (Zheng et al. 2002b) with modification. Briefly, muscles were homogenized with a blender homogenizer (Kinematica, Switzerland) in 1% digitonin buffer (1% digitonin, 185 mm KCl, 1.5 mm CaCl2 and 10 mm Hepes; pH 7.4) on ice. Cellular debris was pelleted by centrifugation at 10 000 g for 10 min at 4°C. Protein concentration was measured by bicinchoninic acid (BCA) protein assay (Pierce Biotechnology, Rockford, IL, USA). The lysate (500 μg total cellular protein) was pre-cleared by adding 0.5 μg of the appropriate control IgG (normal rabbit IgG), together with 20 μl of resuspended volume of the appropriate agarose conjugate (protein G-agarose). Samples were then incubated at 4°C for 30 min. After centrifugation at 500 g for 5 min at 4°C, supernatant was transferred to a fresh tube on ice. Rabbit anti-DHPR α1C primary antibody (Sigma, St Louis, MO, USA; 1 μg) was added and incubated overnight at 4°C on a rotating device. The control tube received only rabbit IgG. Then 20 μl of resuspended volume of the Protein G-Agarose was added to each tube and incubated at 4°C for 2 h. After centrifugation at 500 g for 5 min at 4°C, the pellets were washed in PBS three times and resuspended in 20 μl of 1 × electrophoresis sample buffer. All samples were boiled for 2–3 min and separated by 10% denaturing SDS-PAGE at 100 V for 4–5 h. Rainbow Molecular Weight Marker (Amersham Pharmacia Biotech Inc., Piscataway, NJ, USA) was loaded for reference. Proteins were transferred from the gel to nitrocellulose membrane (Amersham Pharmacia Biotech Inc.).
Microsome preparation
Microsomes from mouse heart and skeletal muscle were prepared as described (Saito et al. 1984; Inui et al. 1988) with modification. Mouse skeletal (5 g) and heart (3 g) muscles were finely cut, and were homogenized with a blender homogenizer (Kinematica) in a homogenization buffer containing 5 mm imidazol (pH 7.4) and 300 mm sucrose with complete protease inhibitor cocktail (Roche Diagnostics, Indianapolis, IN, USA). The homogenate was centrifuged at 5000 g for 15 min and the supernatants were filtered through four layers of cheesecloth and centrifuged at 15 000 g 15 min, then the second supernatants were centrifuged at 100 000 g for 90 min. Microsome pellets were resuspended in 1% digitonin buffer. After the measurement of protein concentration by BCA protein assay, samples were mixed with 1 × sample buffer. All samples were boiled for 2–3 min, separated on 10% denaturing SDS-PAGE gels, and transferred to nitrocellulose membranes.
Immunoblot
Nitrocellulose membranes for microsome preparations and immunoprecipitation were blocked by incubating membranes in 150 mM NaCl, 10 mM Tris-HCl, pH 7.4 (TBS), 0.1% Tween-20, 5% milk for 30–60 min at room temperature. Membranes were then incubated with rabbit anti-DHPR α1C primary antibody (Sigma), diluted 1: 100 in TBS, 5% milk, 0.1% Tween-20, for 1 h at room temperature and washed three times for 5 min each with TBS, 0.05% Tween-20. After incubation for 30 min at room temperature with peroxidase-conjugated anti-rabbit lgG (diluted 1: 3000 in TBS, 0.1% Tween-20, 5% milk), membranes were washed three times for 5 min each with TBS, 0.05% Tween-20, and once for 5 min with TBS. Finally, the membrane was incubated in ECL Reagent (Pierce Biotechnology) and visualized by X-ray film.
RyR protein detection
Crude membrane preparation
Crude membrane fractions from pooled hindlimb muscles from adult frog and young and old mice were prepared as previously described (Chun et al. 2003). Briefly, frozen muscles were chopped into small pieces with scissors and homogenized in 500 μl ice-cold homogenization buffer (20 mm Hepes, pH 7.4, 250 mm sucrose, 0.2% sodium azide, with complete protease inhibitor cocktail) with a blender homogenizer (Kinematica, Switzerland). Crude membrane fraction was extracted by adding 500 μl extraction buffer (5 mm NaPO3, pH 7.4, 75 mm NaCl, 1% Triton X-100, 1% deoxycholate, 0.1% SDS, with complete protease inhibitor cocktail) and centrifuging at 10 000 g for 15 min. Proteins in the supernatant were separated on 4% SDS-PAGE gels, and transferred to polyvinylidene difluoride membranes (Amersham Biosciences, UK) in transfer buffer (12 mm Tris-base, 92 mm glycine, 0.02% SDS and 20% methanol) at 25 V overnight at 4°C.
Immunoblot
Membranes were blocked for 3 h in 5% milk TBST at room temperature. Membranes were then incubated in monoclonal anti-RyR (Affinity Bioreagents, Golden, CO, USA) diluted at 1:5000 in 5% milk TBST at room temperature for 60 min, and washed three times for 5 min each in TBS and 0.05% Tween-20. This antibody reacts with both RyR1 and RyR3 (Chun et al. 2003). After incubation for 60 min in peroxidase-conjugated anti-mouse IgG (Amersham Biosciences) diluted 1:10 000 in 5% milk TBST, membranes were washed again three times for 5 min each in TBS and 0.05% Tween 20, and once in TBS. Finally the membranes were incubated in ECL reagent (Pierce Biotechnology) and visualized on X-ray film.
Statistics
Values reported are mean ± s.e.m. Statistical analysis was performed using analysis of variance (ANOVA) followed by multiple comparisons tests (Tukey's HSD). An alpha value of P < 0.05 was considered significant.
Results
FDB single intact fibre contractility in Ca2+-free solution
The first set of experiments was performed in order to determine whether the absence of external Ca2+ ions affects young or old muscle contractility. To assess the effects of external Ca2+ on young and old muscle fibre contractility, force was measured in single intact FDB fibres from young (n = 11 fibres) and old (n = 14 fibres) mice in the presence of both normal recording solution and Ca2+-free solution. The single intact fibre preparation allows for several advantages over a multifibre preparation. First, the contracting fibre is very rapidly exposed to the perfusion solution. Second, whereas fibres can heterogeneously contribute to whole muscle force production due to differences in pennation and fibre length (Sugi & Tsuchiya, 1998), one can be assured the single intact fibre is functioning at optimal length (LO) (González et al. 2000). Force was normalized to baseline values for each fibre. Figure 1A shows that the reference trial contraction protocol induced no decline in peak force in single intact fibres from young and old mice (P > 0.05). The same contraction protocol in the absence of external Ca2+ (the test trial) produced no significant force decline in fibres from young mice (P > 0.05), while force declined significantly in the fibres from old mice (35% decline, P < 0.05), compared to the reference trial. The effects of Ca2+-free solution on old FDB fibres was greater than on young fibres (P < 0.05).
Figure 1. Absence of external Ca2+ ions causes peak tetanic force to decline in single intact FDB fibres from old mice.
Single intact FDB fibres from young (n = 11) and old (n = 14) mice underwent the same two contractile protocols (see Methods). A, peak force in fibres from young mice did not significantly decline during the test trial compared to the reference trial (P > 0.05), while force declined significantly in fibres from old mice during the test trial compared to the reference trial (*P < 0.05). Force decline was significantly different in fibres from old and young mice (#P < 0.05). B, two populations of muscle fibres were distinguishable in old mice: fibres in which peak force did not decline in Ca2+-free solution (n = 8, Old Non-Affected, P > 0.05) similar to fibres from young mice, and fibres in which force greatly declined in Ca2+-free solution (n = 6, Old Affected, **P < 0.001 versus reference trial, ##P < 0.001 versus Young and Old Non-Affected test trial).
Upon further examination of the old FDB fibres, two clearly delineated populations of fibres from old mice were distinguished. Within the group there exists a population of fibres (Old Non-Affected, n = 8 of 14) that were not affected by Ca2+-free solution (P > 0.05), similar to young fibres. There is also a population of fibres from old mice (Old Affected, n = 6 of 14) that showed dramatic force decline in Ca2+-free solution compared to the reference trial (P < 0.001) and compared to both young (P < 0.001) and Old Non-Affected (P < 0.001) fibres in the absence of external Ca2+ (Fig. 1B). Of the fibres from young mice, only one fibre declined substantially in Ca2+-free solution (∼35% decline). Even though this fibre exists more than two standard deviations outside the mean for this group, it was included in the data because force recovered completely by the end of the experiment. The presence of this fibre, combined with some depression in force recorded in fibre bundles (n = 5–10 fibres) from young mice (data not shown), indicates that some fibres from young, although rare, may also be partially dependent on external Ca2+ ions for EC coupling.
Figure 2A shows the time course of the reference (left) and test (right) trials and a recovery tetanus (5 min after the end of test trial, far right) in one fibre from an old mouse that was unaffected by Ca2+-free solution. Peak force is maintained throughout the reference trial, indicating that the protocol induced no decline in peak tetanic force in this fibre population. Similarly, peak force does not significantly decline during the test trial when Ca2+-free solution is perfused. Individual force traces from the same fibre during each contraction protocol are shown in Fig. 2B (a–f). Figure 2C shows the two contraction protocols in one fibre from an old mouse that was significantly affected by Ca2+-free solution. Peak force was maintained during the reference trial (left), indicating that the protocol also induced no fatigue in this fibre population. However, during the test trial (right), peak force significantly declined (approximately 60% for this fibre) during perfusion of Ca2+-free solution. Upon return to normal 1.8 mm Ca2+ solution, force recovered to near baseline. Given more time (5 min after the end of test trial; far right), force recovered to initial levels. This, taken together with the lack of force decline in the reference trial, indicates that force decline during the test trial is due to the absence of external Ca2+ ions. Individual force traces from the same fibre during each contraction protocol are shown in Fig. 2D (a–f).
Figure 2. Two populations of single intact FDB fibres from old mice.
A, time course of reference (left traces) and test (right traces) trials for one representative ‘unaffected’ single intact FDB fibre from an old mouse. Top bar shows solution changes during the trials. Peak force in this fibre did not decline during the reference trial, and peak force in this fibre minimally declined in the absence of external Ca2+ ions during the test trial. C, time course of reference (left traces) and test (right traces) trials for one representative ‘affected’ single intact FDB fibre from an old mouse. Top bar shows solution changes during the trials. Force declined by approximately 60% for this fibre in Ca2+-free solution, then recovered to ∼90% of baseline once normal physiological solution was returned. Given enough recovery time, force recovered to initial levels (far right trace). B and D, tetanic force traces corresponding to selected traces in A and C, repectively: contraction 30 of 150 (a and d, immediately preceding change to Ca2+-free solution in test trial), contraction 90 of 150 (b and e, immediately preceding return to normal recording solution in test trial), contraction 150 of 150 (c, end of the reference trial), and the recovery tetanus (f, 5 min after the end of test trial).
DHPR function
A key step in EC coupling is the voltage sensing function of the DHPR resulting in intramembrane charge movements. Decreases in charge movements would probably decrease force generation in a muscle fibre. Therefore, to determine whether the absence of external Ca2+ ions affected the function of the DHPR, single dissociated FDB fibres were voltage-clamped in the whole cell configuration of the patch-clamp technique (Wang et al. 2000). Intramembrane charge movements were measured in FDB fibres from young (n = 16) and old (n = 11) mice after blocking the inward Ca2+ current (see Methods). Maximum charge movement (Qmax) was lower in old compared to young fibres (P < 0.05), which agrees with previously published data (Wang et al. 2000, 2002). However, no charge movement difference was found between 2 mm and 0 mm Ca2+ external solutions in either young (P > 0.05, Fig. 3A and B) or old (P > 0.05, Fig. 3C and D) FDB fibres. Mean data points were fitted to a Boltzmann distribution (eqn (4), see Methods). The best-fitting parameters for Qmax, VQ½, and k for fibres from young mice are: 54.9 nC nF−1, 16.9 mV, and 18.5, respectively, in Ca2+-free conditions and 47.1 nC nF−1, 18.6 mV, and 17.7, respectively, in 2 mm Ca2+solution. The best-fitting parameters for Qmax, VQ1/2, and k for fibres from old mice are: 34.1 nC nF−1, 19.2 mV, and 18.5, respectively, in Ca2+-free conditions, and 31.1 nC nF−1, 15.4 mV, and 18.8, respectively, in 2 mm Ca2+ solution. These data suggest that fibre force decline in Ca2+-free solution is not due to any inhibition of DHPR function in Ca2+-free solution. These results are supported by the fact that Ca2+-free solution did not have a general effect on contraction in all fibres. Indeed, a general deleterious effect of Ca2+-free solution on charge movements in all fibres would probably lead to force decline in all fibres. Additionally, two populations of fibres were not found in fibres from old mice (Fig. 3E), further suggesting that alteration in DHPR function in Ca2+-free solution does not contribute to force decline in Ca2+-free solution.
Figure 3. Intramembrane charge movement (Qon) measured in the presence and the absence of external Ca2+.
Qon measured over a range of stimulation voltages in FDB fibres from both young (A and B, n = 16) and old (C and D, n = 11) mice. Single dissociated FDB fibres were voltage-clamped in the whole-cell configuration of the patch-clamp technique. Charge movements were evoked with 25-ms depolarizing voltage steps from the holding potential (−80 mV) to command potentials ranging from −70 to 60 mV. Qon–Vm relationship in fibres from young (A) and old (C) mice in 2 mm Ca2+ external solution (□) and in Ca2+-free external solution (•). Data points were fitted to a Boltzmann distribution (eqn (4)) (continuous line, 0 mm Ca2+ solution; dashed line, 2 mm Ca2+ solution). Sample charge movement recordings in fibres from a young mouse (B) and an old mouse (D). Numbers on the right indicate the membrane potential. Dotted lines represent the baseline. E, individual maximum Qon data points for fibres from young and old mice in 2 mm and 0 mm Ca2+ solution show single populations.
Intracellular Ca2+ -transients
Following DHPR activation, the next key step in EC coupling is SR Ca2+ release via RyR1. Impairment of SR Ca2+ release from RyR1 results in a decreased intracellular Ca2+ transient and decreased force. Therefore, in order to determine whether intracellular Ca2+ was reduced in muscle fibres while in Ca2+-free solution, intracellular Ca2+ transients were measured in FDB fibres under two conditions: in voltage-clamped fibres under whole cell patch-clamp and in single intact fibres undergoing electrically elicited contractions. Intracellular Ca2+ transients, across a range of depolarization potentials in voltage-clamped fibres, are not affected by the absence of external Ca2+ ions in fibres from young mice (n = 15). Similar to the contractile data, the results in Fig. 4 show that two populations of fibres exist in old mice: fibres in which Ca2+ transients were not reduced while in Ca2+-free solution similar to fibres from young (P > 0.05, Old Non-Affected, n = 8 of 11) and fibres in which Ca2+ transients were greatly reduced in Ca2+-free solution compared to fibres from both young and old (P < 0.01, Old Affected, n = 3 of 11). These results confirm, with a different technique, the results in contracting fibres and suggest that a population of fibres from old mice are at least partially dependent upon external Ca2+ for efficient excitation–Ca2+ release coupling.
Figure 4. Intracellular Ca2+ measured in voltage-clamped FDB fibres in whole-cell patch clamp.
Single dissociated FDB fibres (young, n = 12; old n = 11) were voltage-clamped in the whole-cell configuration of the patch-clamp technique. Cells were loaded with the fluorescent Ca2+ indicator fluo-4. Intracellular Ca2+ release was evoked with 20-ms depolarizing voltage steps from the holding potential (−80 mV) to command potentials ranging from −40 to 30 mV in both 2 mm and 0 mm external Ca2+. The ratio of fluorescence measured in 0 mm external Ca2+ to fluorescence measured in 2 mm external Ca2+ was calculated. Two populations of fibres were identified from old mice: those in which Ca2+ transients were unaffected by the absence of external Ca2+ (n = 8 of 11, Old Non-Affected) similar to fibres from young mice and those in which Ca2+ transients were reduced in Ca2+-free solution (n = 3 of 11, Old Affected). *P < 0.05, Old Affected versus Old Non-Affected; **P < 0.01 Old Affected versus Young and Old Non-Affected.
In order to determine whether the lower intracellular Ca2+ transient recorded in Ca2+-free external solution is the cause of the force decline in fibres from old mice, intracellular Ca2+ transients were recorded simultaneously with force in fibres from young (n = 3) and old (n = 3) mice in normal recording solution and in Ca2+-free solution. These experiments also contained two contraction sequences: the first to assess the effects of the Ca2+-free solution on elicited force for that fibre, the second to record simultaneous force and Ca2+ transients (see Methods).
Ca2+-free solution did not cause force or intracellular Ca2+ transients to decline in fibres from young mice (Fig. 5A). However, in fibres from old mice, Ca2+-free solution caused both intracellular Ca2+ transients and force to decline in parallel (Fig. 5B). Within these three fibres from old mice, there was a large range of force decline (fibre 1021 A showed ∼92% force decline; fibre 1105 B ∼51% force decline; and fibre 1124 A ∼15% force decline). The range of force decline in these three fibres was within the range reported for the original experiments shown in Fig. 1, and covers the range of force decline of the two populations of fibres: fibres 1031 A and 1105 B fall into the ‘Old Affected’ group while fibre 1124 A falls into the ‘Old Non-Affected’ group. The data illustrated in Fig. 5B are from fibre 1105 B (∼51% force decline). Individual data points for force and intracellular Ca2+ from all time points during experiments in all fibres from young and old mice were analysed with regression analysis. These data indicate that the decreased force and decreased intracellular Ca2+ transient induced by Ca2+-free solution are directly related (r2 = 0.829, P < 0.001). These data support the concept that some muscle fibres from old mice are partially dependent upon external Ca2+ for efficient EC coupling and maintenance of tetanic force.
Figure 5. Simultaneous force and Ca2+ traces from single FDB fibres.
Fibres were loaded with fluorescent Ca2+ indicator fluo-4. Maximal tetanic contractions were elicited every 2 min for 20 min; Ca2+-free solution was perfused from start time to min 10, and normal 1.8 mm Ca2+ solution was perfused from min 10–20. ‘Baseline’ and ‘Recovery’ traces were recorded in the presence of normal 1.8 mm external Ca2+, while all other traces were recorded in the absence of external Ca2+. A, compared to baseline, neither force nor Ca2+ fluorescence declined in the absence of external Ca2+ in a fibre from a young mouse. B, compared to baseline, both force and Ca2+ fluorescence declined in parallel in the absence of external Ca2+ in a fibre from an old mouse. Vertical scale bars represent both 100% force and 100% ΔF/F normalized to baseline values.
DHPR isoforms in skeletal muscle from young and old mice
Ca2+ current recordings
The current–voltage relationship, the activation time constant and the Ca2+ -dependence of inactivation of inward Ca2+ current (ICa) were examined to explore the possibility of an age-related shift from DHPR α1S to DHPR α1C. Figure 6A shows the whole cell Ca2+ current–voltage (I–V) relationship for fibres from young (n = 11) and old (n = 10) mice. The data were fitted to eqn (1) (see Methods). The best fitting parameters for Gmax, Vr, V½ and z determined from mean data points are 73 nS nF−1, 66 ± 4.2 mV, − 1.9 ± 0.02 mV and 6.5 ± 2.1, respectively, for fibres from young mice; and 66 nS nF−1, 63 ± 3.7 mV, 0.5 ± 0.01 mV and 4.8 ± 0.02, respectively, for fibres from old mice. The I–V relationship for DHPR α1C is shifted leftwards compared to α1S (García et al. 1994). Here, we show no differences in I–V curve orientation in fibres from young compared to old mice. Individual Ca2+ current data at 0 and + 10 mV for fibres from both young and old mice are evenly distributed, suggesting that two populations do not exist in this measure (Fig. 6B). The activation time constant (τa) was determined by fitting the current trace to an exponential function from the beginning of the pulse to the point where the current reached a steady-state level. As DHPR α1C calcium current has a faster τa than that of α1S (Tanabe et al. 1991; Nakai et al. 1994), an age-related subunit shift should be evident in the τa values. Figure 6C shows τa values corresponding to a range of test potentials. At 10 mV, τa is significantly faster in fibres from old (n = 7) than from young (n = 8) FDB fibres, suggesting the possibility of the presence of DHPR α1C subunits in old muscle. The one barrier Eyring rate model was fitted to the mean time constant of ICa current activation. The best fitting parameters for τmax, V½, and K (see eqn (2), Methods) were: 16.4 ms, 7.7 mV and 10.7 mV, respectively, for fibres from young mice; and 11.2 ms, 5.5 mV and 15.4 mV, respectively, for fibres from old mice. Despite the difference in τa between young and old, individual data for τa at +10 mV also do not display two populations in fibres from young of old mice, suggesting that this may not be a mechanism to explain force decline in Ca2+-free solution. Representative Ca2+ current traces from young and old at a 10 mV command potential are shown in Fig. 6E. The raw Ca2+ current recording is shown (dashed line) with a single exponential fitting to the rising phase (Delbono, 1992) overlaid onto the traces. The rising phase, and hence τa, is faster in old compared to young at 10 mV.
Figure 6. Current–voltage relationship and activation time constant of Ca2+ current measured in patch-clamped fibres from young and old mice.
Fibres were voltage-clamped, and inward Ca2+ currents (ICa) were evoked with 350-ms depolarizing voltage steps from the holding potential (−80 mV) to command potentials ranging from −70 to 50 mV. A, Ca2+ current–voltage (I–V) relationship. Data from young (n = 11) and old (n = 10) were fitted to eqn (1) (continuous lines, see Methods). I–V relationship shows no age-related shift in the voltage-dependence of the Ca2+ current. B, individual data points for fibres from young and old mice at 0 and + 10 mV show single populations of data. C, the rising phase of the inward Ca2+ current was fitted to a single exponential function (see Methods). The activation time constant (τa) was determined by fitting the current trace to an exponential function from the beginning of the pulse to steady state (∼30 ms). The data were fitted to a single barrier Eyring model (continuous lines, eqn (2)). τa–Vm relationship in fibres from young (•) and old (○) animals. At 10 mV, τa is significantly slower in fibres from young than from old mice. D, individual τa data points for fibres from young and old mice at the + 10 mV command potential show single populations of data. E, sample ICa traces from representative fibres from young and old mice at 10 mV.
ICa inactivation of DHPR α1S is voltage-dependent (Cota et al. 1984), whereas ICa inactivation of DHPR α1C is voltage and Ca2+ dependent (De Leon et al. 1995; Zuhlke et al. 1999). Double-pulse experiments were performed in both Ca2+-and Ba2+-containing solutions (see Methods) to examine the Ca2+ dependence of inactivation of ICa. Relative current inactivation (ICa/ICa,max) was determined by normalizing the current during the test pulse to the maximum current measured during the set of test pulses. Current inactivation was measured with Ba2+ because Ba2+ current (IBa) inactivation is sensitive only to voltage inactivation. An ICa inactivation curve that is shifted leftwards in relation to IBa inactivation indicates Ca2+ dependence of ICa current inactivation. Figure 7 shows no Ca2+ dependence of ICa inactivation in either young (Figs 7A and B; n = 6) or old (Figs 7C and D; n = 7) fibres (P > 0.05). The best fitting parameters for the mean values of Ca2+ current (see eqn (3), Methods) for A, V½ and k were: 0.8, −6.9 mV and 13.1 mV, respectively, for fibres from young mice; and 0.73, −7.4 mV and 8.5 mV, respectively, for fibres from old mice. The best fitting parameters for the mean values of Ba2+ current for A, V½ and k were: 0.93, −17 mV and 13.6 mV for fibres from young mice; and 0.75, −12.2 mV and 11.0 mV, respectively, for fibres from old mice. These data disagree with the τa data and the suggestion of the presence of DHPR α1C in old muscle fibres.
Figure 7. Ca2+ dependence of inactivation of Ca2+ current (ICa) measured in patch-clamped fibres from young and old mice.
Fibres were voltage-clamped, and ICa was evoked with two 300-ms depolarizing voltage steps from the holding potential (−80 mV) in double-pulse experiments. The first pulse ranged from −70 to 60 mV. The second (test) pulse was held constant at 20 mV, with a 100-ms interpulse interval (B, top illustration). Relative current inactivation (ICa/ICa,max) was determined by subtracting the current during a given test pulse from the highest current measured during the set of test pulses for that fibre. Experiments were repeated replacing Ca2+ in solution with Ba2+. Data were fitted to a Boltzmann equation (eqn (3)). ICa and IBa inactivation curves in fibres from young (A) and old (C) mice show no Ca2+ dependence of ICa inactivation, as no leftward shift was found for the ICa inactivation curve compared to the IBa curve. Sample ICa traces from double-pulse experiments from single FDB fibres from young (B) and old (D) mice display time course and magnitude of ICa responses over a range of command potentials.
DHPR α1C detection
To clarify whether the dependence of EC coupling on external Ca2+ ions in old fibres is due to an age-related shift in DHPR isoforms from the skeletal α1S subunit to the cardiac α1C subunit, we performed ribonuclease protection assay (RPA) and immunoblot analysis. RPA could not detect the presence of DHPR α1C mRNA in either young or old skeletal muscle (Fig. 8A). Similarly, immunoblot analysis following immunoprecipitation (Fig. 8B) and following isolation of muscle membrane fraction (Fig. 8C) did not detect the presence of DHPR α1C protein in young or old skeletal muscle, suggesting that a shift between these two specific DHPR subunit isoforms does not occur with ageing. Samples from heart muscle were used as positive control for DHPR α1C.
Figure 8. Detection of DHPR α1C mRNA and protein and RyR protein isoforms.
A, ribonuclease protection assay did not reveal the presence of DHPR α1C mRNA in hindlimb skeletal muscle from young (Y) or old (O) mice. Size marker (M) and undigested probe (P) are shown. Mouse heart muscle was used as a positive control (H). B, immunoblot following IgG immunopreciptiation (see Methods) failed to show the presence of DHPR α1C protein in hindlimb muscle from young (Y) or old (O) mice. Mouse heart muscle was used as a positive control (H). C, immunoblot following membrane isolation failed to reveal the presence of DHPR α1C protein in hindlimb muscle from either young (Y) or old (O) mice. Mouse heart muscle was used as a positive control (H). D, immunoblot following membrane fraction isolation revealed the presence of RyR1, but failed to reveal the presence of RyR3 in hindlimb skeletal muscle from young or old mice. Frog hindlimb muscle was used as a positive control for both RyR1 (RyRα) and RyR3 (RyRβ).
RyR isoforms in skeletal muscle from young and old
As no shift from DHPR α1S to α1C is evident, we examined a potential change in RyR isoforms expressed in ageing skeletal muscle. Embryonic skeletal muscle co-expresses RyR1 and RyR3 (Flucher et al. 1999; Chun et al. 2003), whereas adult fast-twitch skeletal muscle does not express RyR3 (Flucher et al. 1999). A shift in expression of RyR isoforms in ageing mammalian skeletal muscle towards RyR3 expression may also induce EC coupling dependence on external Ca2+ ions, due to the fact that DHPR and RyR3 do not physically couple (Fessenden et al. 2000; Protasi et al. 2000). In this scenario, RyR3 would only be able to contribute to the SR Ca2+ release transient via Ca2+ -induced Ca2+ release (CICR) activated by Ca2+ influx through the DHPR and Ca2+ release from neighbouring RyR1s. Figure 8D shows immunoblot results from young and old mouse pooled hindlimb skeletal muscle. No RyR3 protein could be detected in either young or old skeletal muscle, suggesting no age-induced shift in RyR isoforms. Adult frog muscle was used as a positive control for both RyR1 and RyR3, as frog muscle has been previously shown to express both proteins in relatively equal amounts (Chun et al. 2003).
Discussion
This work supports the conclusion that ageing leads to dependence on external Ca2+ ions for maintenance of intracellular Ca2+ release and force generation in old skeletal muscle fibres, and that this age-related dependence is not due to a switch in DHPR or RyR subtype.
Necessity of external Ca2+ ions to maintain force
In this work, we investigated the effects of the absence of external Ca2+ ions on single intact FDB muscle fibre contractile force. We hypothesized that EC coupling in muscle fibres from old animals becomes dependent on external Ca2+. Controversy exists over whether external Ca2+ is necessary for adult mammalian skeletal muscle EC coupling. Some previous studies have shown a need for external Ca2+ to maintain tetanic force in adult skeletal muscle (Anwyl et al. 1984; Kotsias et al. 1986; Oz & Frank, 1991), while others have shown no need (Dulhunty & Gage, 1988; Tanabe et al. 1990b; García et al. 1994). These differences may exist due to the inclusion of metal ion chelators, such as EGTA, in experiments that indicate the necessity of external Ca2+ for maintenance of tetanic force (Anwyl et al. 1984; Kotsias et al. 1986; Oz & Frank, 1991). No examination of EC coupling dependence on extracellular Ca2+ ions has been conducted in skeletal muscle from ageing animals. This study is the first to examine the importance of external Ca2+ ions for EC coupling in ageing skeletal muscle fibres. This work examined the question without EGTA, and shows that a population of muscle fibres from old mice is, indeed, dependent on the presence of external Ca2+. Force decline in Ca2+-free solution was not due to fatigue, as the reference contraction protocol in normal 1.8 mm Ca2+ solution did not cause force decline in fibres from either young or old mice.
Mechanisms of decreased force in the absence of extracellular Ca2+
To ascertain the mechanism(s) of decreased force in fibres from old mice induced by the removal of external Ca2+, we examined the activation of the DHPR. Intramembrane charge movement in fibres from young or old mice showed no difference in Ca2+-free compared to Ca2+-containing solution, and showed a single population in Ca2+-free solution. This indicates that the activation of DHPR is unaffected by a reduction in external Ca2+ ions, and is still capable of coupling with RyR1 and inducing SR Ca2+ release.
Intracellular Ca2+ release was examined in two experimental preparations: in voltage-clamped fibres and in electrically stimulated contracting fibres. Voltage-clamp experiments revealed reduced intracellular Ca2+ transients in a population of fibres from old mice, while contraction experiments showed parallel reduction in force and intracellular Ca2+ in the absence of external Ca2+ ions. The direct relationship between Ca2+ transient amplitude and force generation shown here supports the conclusion that impaired force generation in fibres from old mice upon removal of external Ca2+ is due to impaired SR Ca2+ release.
Mechanisms of decreased SR Ca2+ release
If ageing causes a subunit switch from α1S to α1C similar to regenerating skeletal muscle (Pereon et al. 1997a), DHPR–RyR coupling would become cardiac-like (dependent upon influx of external Ca2+ ions) similar to regenerating skeletal muscle (Louboutin et al. 1995, 1996; Pereon et al. 1997b). In this case, the absence of external Ca2+ would prevent Ca2+ influx through the α1C to trigger CICR from the RyR1. The analysis of the activation phase of the Ca2+ current suggests a possible age-related DHPR subunit shift. However, the I–V curves for fibres from young and old show that activation does not occur at more negative potentials for old versus young as expected for α1C expression. The inward Ca2+ current becomes apparent around −20 mV and the half-activation potential for young and old is similar. Additionally, Ca2+ current inactivation in response to double-pulse experiments in voltage-clamped cells showed no Ca2+ dependence in fibres from either young or old mice. Also, none of these measures displayed two populations in fibres from old mice. These data suggest no shift in DHPR isoform from α1S to α1C with age. Therefore, RPA and two immunoblot techniques were performed to clarify this. These assays failed to detect DHPR α1C mRNA or protein in muscle from either young or old mice, again suggesting no age-related shift in DHPR isoforms from α1S to α1C.
Faster τa in old muscle fibres compared to young cannot be explained here by the presence of DHPR α1C in aged skeletal muscle. The faster current activation may be due to a number of factors. Repeat I of DHPR is very important for L-type Ca2+ current activation kinetics (Tanabe et al. 1991; Nakai et al. 1994). Age-related alterations (splice variations, post-translational processing and phosphorylation) in this portion of the α1 subunit of the DHPR could, conceivably, alter the Ca2+ current kinetics. Similarly, alterations in expression of accessory subunits may play a role in age-related alterations of Ca2+ current activation kinetics. For example, the β subunit increases τa when co-expressed with the α1 subunit in a cell transfection system (Varadi et al. 1991).
The second major protein complex involved in EC coupling at the triad is the RyR. Adult skeletal muscle expresses RyR1 almost exclusively (Marks et al. 1989; Takeshima et al. 1989). Given that a shift from RyR1/RyR3 co-expression during development to exclusive expression of RyR1 in adult skeletal muscle occurs (Flucher et al. 1999; Chun et al. 2003), it is conceivable that a shift in the opposite direction may occur with ageing. A shift in expression from RyR1 to RyR1/RyR3 co-expression would require Ca2+ influx through the DHPR to activate RyR3, via CICR, to release Ca2+ from the SR, since RyR3 is not directly coupled to DHPR (Fessenden et al. 2000; Protasi et al. 2000). However, no expression of RyR3 protein was found in either young or old mouse hindlimb skeletal muscle, suggesting that no shift occurs in RyR isoform expression with ageing.
Another potential mechanism for impaired SR Ca2+ release in the absence of external Ca2+ ions is SR Ca2+ depletion. If the SR becomes depleted of Ca2+ during the contractile protocol in the absence of external Ca2+ ions, the SR Ca2+ release and force will decline. Skeletal muscle from MG29 knockout mice has been shown to contain two apparently different Ca2+ storage pools – one voltage-sensitive, the other non-voltage-sensitive – as shown by caffeine contracture after a SR Ca2+ depletion protocol. This secondary, voltage-insensitive, caffeine-sensitive Ca2+ pool has been hypothesized to result from fragmentation of the SR, leaving fragments of Ca2+-containing SR disconnected from the t-tubule membranes (Kurebayashi et al. 2003). A smaller voltage-sensitive Ca2+ storage pool could conceivably lead to depletion of available Ca2+ during contraction in Ca2+-free solution. Although alterations in MG29 have not been explored in ageing skeletal muscle, similar changes in the conformation of the intracellular membranes may explain the decreased tetanic force in muscle fibres from old mice in Ca2+-free solution.
Physical changes to the muscle fibre membrane structure with age could also possibly lead to age-related Ca2+-dependent EC coupling. Certain types of heart failure show expansion in the space between t-tubules and SR terminal cisternae (Gomez et al. 1997), interfering with the efficiency of normal CICR necessary for cardiac EC coupling. A similar spacing increase in skeletal muscle could disrupt the mechanical coupling of DHPR α1S and RyR1, changing the coupling mechanism to Ca2+ dependent in muscle fibres from old mice. Altered membrane structure/spacing can result from decreased expression of putative triad membrane anchoring proteins, such as MG29 (Takeshima et al. 1998), which has been shown to alter t-tubule structure and arrangement and cause EC coupling to be Ca2+ dependent (Nishi et al. 1999). Recent evidence shows that these membrane structures in ageing human skeletal muscle may, indeed, be altered. The number of contact points between t-tubule and SR membranes is decreased with age, and the t-tubule network shows signs of disarrangement similar to developing skeletal muscle – greater numbers of dyads and longitudinal t-tubules than in normal skeletal muscle (F. Protasi, personal communication). Similar changes to intracellular membrane structures have been shown in junctophilin-1 knockout mice – reduced number of triads in favour of dyad formation and vacuolated terminal cisternae of SR (Komazaki et al. 2002). Whether ageing leads to alterations in MG29 and junctophilin-1 protein expression, as well as other triad proteins, such as JP-45 (Anderson et al. 2003), and whether these functional effects are due to changes in expression of these or other triad proteins with age is unknown at this time. However, age-related decreased expression and/or alterations in structure, function or trafficking of MG29 and similar triad proteins may be candidate mechanisms to explain the results reported here.
In order for uncoupled DHPRs and RyR1s in aged skeletal muscle fibres to initiate Ca2+ release from the SR, the RyR1 must be activated, via CICR, by the Ca2+ influx through the DHPR. Normally, Ca2+ and Mg2+ inactivation of the RyR1 in skeletal muscle is about 20-fold more sensitive than for RyR2 in cardiac muscle (Laver et al. 1995). This allows the Ca2+ influx through the DHPR α1C to activate Ca2+ release from RyR2 in cardiac muscle. In skeletal muscle, the cytoplasmic [Mg2+] may inhibit Ca2+ release from a normal RyR1 that is not undergoing physical interaction with (and activation by) DHPR α1S, undermining EC coupling in these fibres. However, EC coupling does occur in these fibres in the presence of external Ca2+. Splice variants of the RyR1 lacking one or both of 5- and 6-amino acid segments in the modulatory region of the RyR1 responsible for Ca2+ or Mg2+ and calmodulin binding are expressed developmentally (Futatsugi et al. 1995). If the DHPR α1S and RyR1 are indeed physically uncoupled in some aged skeletal muscle fibres, expression of one or more of these splice variants may possibly reduce the sensitivity of RyR1 Mg2+ inactivation allowing EC coupling to occur in a cardiac-like manner.
Age-related alterations to the DHPR α1S molecule may also explain the Ca2+ dependence of EC coupling and possibly changes in Ca2+ current activation. Recent evidence shows that there is a splice variant of the DHPR α1S isoform which has a 19 amino acid deletion in the repeat IV S3–S4 extracellular loop (Jurkat-Rott & Lehmann-Horn, 2004). This splice variant has been found in adult human skeletal muscle at levels at or below 10% of total DHPR transcripts. However, in myotubes regenerating from human satellite cells – a condition known to induce Ca2+-dependent EC coupling (Louboutin et al. 1995, 1996; Pereon et al. 1997a) – this 19 amino acid deletion transcript has been found to make up more than 66% of total DHPR transcript (K. Jurkat-Rott and F. Lehmann-Horn, personal communication). The functional significance of this splice variant and its expression level in ageing muscle have not yet been examined. If ageing muscle fibres display a shift in DHPR expression such that the deletion variant becomes dominant in some fibres, and the deleted sequence plays a role in overall DHPR function, this could help explain our findings. Previous work has shown that the addition of divalent cations such as Mg2+, Co2+ and Cd2+ can alter EC coupling (Dulhunty & Gage, 1989; Mould & Dulhunty, 1999, 2000), all of which may affect the binding of external Ca2+ by sites on the sarcolemmal or t-tubule membrane (Mould & Dulhunty, 2000). Whether this site or some other potential age-related DHPR alteration plays a role in ‘sensing’ the presence or absence of external divalent cations remains unknown at this time.
Other DHPR subunits may play a role in inducing Ca2+-dependent EC coupling in skeletal muscle too. Expression of the cardiac DHPR β2 subunit in skeletal DHPR β1 knockout myotubes conferred Ca2+-dependent EC coupling to those fibres (Sheridan et al. 2003a). Similarly, truncation of the C-terminus of the skeletal DHPR β1 subunit also caused Ca2+-dependent EC coupling (Sheridan et al. 2003b). Whether ageing causes a shift in the expression of DHPR β subunit isoforms or a splice variant of the DHPR β subunit exists is unknown at this time.
The question remains: what is the consequence for old muscle fibres of becoming dependent on external Ca2+ ions for EC coupling? Cardiac myocytes are dependent on the influx of external Ca2+ for EC coupling (Fabiato, 1985; Nabauer et al. 1989) and undergo billions of contractions throughout a lifetime. As the t-tubules of skeletal muscle fibres are much smaller in diameter (approximately 20 nm; Franzini-Armstrong et al. 1975) than cardiac muscle fibres (approximately 250 nm; Soeller & Cannell, 1999), the Ca2+ buffering ability of the skeletal muscle t-tubule lumen is very low (Almers et al. 1981), and the coefficient for diffusion of Ca2+ in the t-tubule is very low (Almers et al. 1981), the possibility exists that Ca2+ is depleted from the t-tubule during a series of tetanic contractions. A muscle fibre that is dependent upon the influx of these Ca2+ ions for EC coupling would thus be at a force-generating disadvantage compared to normal skeletal muscle fibres that are not dependent upon the influx of Ca2+ from the t-tubule. Therefore, we propose that the Ca2+ dependence of EC coupling in a population of skeletal muscle fibres from ageing mice described herein contributes to a decline in force production in fibres undergoing repetitive contraction.
Acknowledgments
The present study was supported by grants from the National Institutes of Health/National Institute on Ageing (AG18755, AG13934 and AG15820) and Muscular Dystrophy Association of America to O.D. We thank Chris Huang for initial studies on DHPR α1C mRNA detection.
References
- Almers W, Fink R, Palade P. Calcium depletion in frog muscle tubules: the decline of calcium current under maintained depolarization. J Physiol. 1981;312:177–207. doi: 10.1113/jphysiol.1981.sp013623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anderson A, Treves S, Biral D, Betto R, Sandona D, Ronjat M, Zorzato F. The novel skeletal muscle sarcoplasmic reticulum JP-45 protein. Molecular cloning, tissue distribution, developmental expression, and interaction with alpha 1.1 subunit of the voltage-gated calcium channel. J Biol Chem. 2003;278:39987–39992. doi: 10.1074/jbc.M305016200. [DOI] [PubMed] [Google Scholar]
- Anwyl R, Bruton J, McLoughlin J. Potassium contractures and mechanical activation in rat skeletal muscle: effects of multivalent cations, temperature and tetracaine. Br J Pharmacol. 1984;82:615–621. doi: 10.1111/j.1476-5381.1984.tb10800.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Armstrong C, Benzanilla F, Horowicz P. Twitches in the presence of ethylene glycol bis (-aminoethyl ether) -N,N-9-tetraacetic acid. Biochim Biophys Acta. 1972;267:605–608. doi: 10.1016/0005-2728(72)90194-6. [DOI] [PubMed] [Google Scholar]
- Beam K, Franzini Armstrong C. Functional and structural approaches to the study of excitation-contraction coupling. Methods Cell Biol. 1997;52:283–306. doi: 10.1016/s0091-679x(08)60384-2. [DOI] [PubMed] [Google Scholar]
- Brum G, Rios E. Intramembrane charge movement in frog skeletal muscle fibres. Properties of charge 2. J Physiol. 1987;387:489–517. doi: 10.1113/jphysiol.1987.sp016586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chaudhari N, Beam K. mRNA for cardiac calcium channel is expressed during development of skeletal muscle. Dev Biol. 1993;155:507–515. doi: 10.1006/dbio.1993.1048. [DOI] [PubMed] [Google Scholar]
- Chun L, Ward C, Schneider M. Ca2+ sparks are initiated by Ca2+ entry in embryonic mouse skeletal muscle and decrease in frequency postnatally. Am J Physiol Cell Physiol. 2003;285:C686–C697. doi: 10.1152/ajpcell.00072.2003. [DOI] [PubMed] [Google Scholar]
- Cognard C, Rivet-Bastide M, Constantin B, Raymond G. Progressive predominance of ‘skeletal’ versus ‘cardiac’ types of excitation-contraction coupling during in vitro myogenesis. Pflugers Arch. 1992;42:207–209. doi: 10.1007/BF00370424. [DOI] [PubMed] [Google Scholar]
- Cota G, Siri L, Stefani E. Calcium channel inactivation in frog (Rana pipiens and Rana moctezuma) skeletal muscle fibers. J Physiol. 1984;354:99–108. doi: 10.1113/jphysiol.1984.sp015365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dangain J, Neering I. Effect of low extracellular calcium and ryanodine on muscle contraction of the mouse during postnatal development. Can J Physiol Pharmacol. 1991;69:1294–1300. doi: 10.1139/y91-190. [DOI] [PubMed] [Google Scholar]
- De Leon M, Wang Y, Jones L, Perez-Reyes E, Wei X, Soong T, Snutch T, Yue D. Essential Ca2+ binidng motif for Ca2+ -sensitive inactivation of L-type Ca2+ channels. Science. 1995;270:1502–1506. doi: 10.1126/science.270.5241.1502. [DOI] [PubMed] [Google Scholar]
- Delbono O. Calcium current activation and charge movement in denervated mammalian skeletal muscle fibres. J Physiol. 1992;451:187–203. doi: 10.1113/jphysiol.1992.sp019160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dulhunty A, Gage P. Effects of extracellular calcium concentration and dihydropyridines on contraction mammalian skeletal muscle. J Physiol. 1988;399:63–80. doi: 10.1113/jphysiol.1988.sp017068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dulhunty A, Gage P. Effects of cobalt, magnesium, and cadmium on contraction of rat soleus muscle. Biophys J. 1989;56:1–14. doi: 10.1016/S0006-3495(89)82647-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fabiato A. Time and calcium dependence of activation and inactivation of calcium-induced release of calcium from the sarcoplasmic reticulum of a skinned canine cardiac Purkinje cell. J Gen Physiol. 1985;85:247–289. doi: 10.1085/jgp.85.2.247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fessenden J, Wang Y, Moore R, Chen S, Allen P, Pessah I. Divergent functional properties of ryanodine receptor types 1 and 3 expressed in a myogenic cell line. Biophys J. 2000;79:2509–2525. doi: 10.1016/S0006-3495(00)76492-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Finch E, Augustine G. Local calcium signaling by inositol-1,4,5,triphosphate in Purkinje cell dendrites. Nature. 1998;396:753–756. doi: 10.1038/25541. [DOI] [PubMed] [Google Scholar]
- Flucher B, Conti A, Takeshima H, Sorrentino V. Type 3 and type 1 ryanodine receptors are localized in triads of the same mammalian skeletal muscle fibers. J Cell Biol. 1999;146:621–630. doi: 10.1083/jcb.146.3.621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flucher B, Franzini-Armstrong C. Formation of junctions involved in excitation-contraction coupling in skeletal and cardiac muscle. Proc Natl Acad Sci U S A. 1996;93:8101–8106. doi: 10.1073/pnas.93.15.8101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Franzini-Armstrong C, Landmesser L, Pilar G. Size and shape of transverse tubule openings in frog twitch muscle fibers. J Cell Biol. 1975;64:493–497. doi: 10.1083/jcb.64.2.493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Futatsugi A, Kuwajima G, Mikoshiba K. Tissue specific and developmentally regulated alternative splicing in mouse skeletal muscle ryanodine receptor mRNA. Biochem J. 1995;305:373–378. doi: 10.1042/bj3050373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- García J, Tanabe T, Beam K. Relationship of calcium transients to calcium currents and charge movements in myotubes expressing skeletal and cardiac dihydropyridine receptors. J Gen Physiol. 1994;103:125–147. doi: 10.1085/jgp.103.1.125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gomez A, Valdivia H, Cheng H, Lederer M, Santana L, Cannell M, McCune S, Altshuld R, Lederer W. Defective excitation-contraction coupling in experimental cardiac hypertrophy and heart failure. Science. 1997;276:800–806. doi: 10.1126/science.276.5313.800. [DOI] [PubMed] [Google Scholar]
- González E, Messi M, Delbono O. The specific force of single intact extensor digitorum longus and soleus mouse muscle fibers declines with aging. J Membr Biol. 2000;178:175–183. doi: 10.1007/s002320010025. [DOI] [PubMed] [Google Scholar]
- González E, Messi L, Zheng Z, Delbono O. IGF-1 prevents age-related decrease in specific-force and intracellular Ca2+ in single intact muscle fibres from transgenic mice. J Physiol. 2003;552:833–844. doi: 10.1113/jphysiol.2003.048165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- González-Serratos H, Valle-Aguilera R, Lathrop D, Garcia M. Slow inward calcium currents have no obvious role in muscle excitation-contraction coupling. Nature. 1982;298:292–294. doi: 10.1038/298292a0. [DOI] [PubMed] [Google Scholar]
- Inui M, Wang S, Saito A, Fleischer S. Characterization of junctional and longitudinal sarcoplasmic reticulum from heart muscle. J Biol Chem. 1988;263:10843–10850. [PubMed] [Google Scholar]
- Jurkat-Rott K, Lehmann-Horn F. The impact of splice isoforms on voltage-gated calcium channel α1 subunits. J Physiol. 2004;554:609–619. doi: 10.1113/jphysiol.2003.052712. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Komazaki S, Ito K, Takeshima H, Nakamura H. Deficiency of triad formation in developing skeletal muscle cells lacking junctophilin type 1. FEBS Lett. 2002;524:225–229. doi: 10.1016/s0014-5793(02)03042-9. [DOI] [PubMed] [Google Scholar]
- Kotsias B, Muchnik S, Obejero Paz C. Co2+, low Ca2+, and verapamil reduce mechanical activity in rat skeletal muscles. Am J Physiol. 1986;250:C40–C46. doi: 10.1152/ajpcell.1986.250.1.C40. [DOI] [PubMed] [Google Scholar]
- Kurebayashi N, Takeshima H, Nishi M, Murayama T, Suzuki E, Ogawa Y. Changes in Ca2+ handling in adult MG29-deficient skeletal muscle. Biochem Biophys Res Commun. 2003;310:1266–1272. doi: 10.1016/j.bbrc.2003.09.146. [DOI] [PubMed] [Google Scholar]
- Lannergren J, Westerblad H. The temperature dependence of isometric contractions of single, intact fibres dissected from a mouse foot muscle. J Physiol. 1987;390:285–293. doi: 10.1113/jphysiol.1987.sp016700. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laver D, Baynes T, Dulhunty A. Magnesium inhibition of ryanodine-receptor calcium channels: evidence for two independent mechanisms. J Membr Biol. 1995;156:213–229. doi: 10.1007/s002329900202. [DOI] [PubMed] [Google Scholar]
- Louboutin J-P, Fitcher Gagnepain V, Noireaud J. Comparison of contractile properties between developing and regenerating soleus muscle – influence of external calcium concentration upon the contractility. Muscle Nerve. 1995;18:1292–1299. doi: 10.1002/mus.880181112. [DOI] [PubMed] [Google Scholar]
- Louboutin J-P, Fitcher Gagnepain V, Noireaud J. External calcium dependence of extensor digitorum longus muscle contractility during bupivicaine induced regeneration. Muscle Nerve. 1996;19:994–1002. doi: 10.1002/(SICI)1097-4598(199608)19:8<994::AID-MUS7>3.0.CO;2-3. [DOI] [PubMed] [Google Scholar]
- Marks A, Tempst P, Hwang K, Taubman M, Inui M, Chadwick C, Fleischer S, Nadal-Ginard B. Molecular cloning and characterization of the ryanodine receptor/junctional channel complex cDNA from skeletal muscle sarcoplasmic reticulum. Proc Natl Acad Sci U S A. 1989;86:8683–8687. doi: 10.1073/pnas.86.22.8683. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marty I, Robert M, Villaz M, De Jongh K, Lai Y, Caterall W, Ronjat M. Biochemical evience for a complex involving dihydropyridine receptor and ryanodine receptor in triad junctions of skeletal muscle. Proc Natl Acad Sci U S A. 1994;91:2270–2274. doi: 10.1073/pnas.91.6.2270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mikami A, Imoto K, Tanabe T, Niidone T, Mori Y, Takeshima H, Narumiya S, Numa S. Primary structure and functional expression of the cardiac dihydropyridine-sensitive calcium channel. Nature. 1989;340:230–233. doi: 10.1038/340230a0. [DOI] [PubMed] [Google Scholar]
- Miller R, Bookstein F, Van Der Meulen J, Engle S, Kim J, Mullins L, Faulkner J. Candidate biomarkers of aging: age-sensitive indices of immune and muscle function covary in genetically heterogeneous mice. J Gerontol A Biol Sci Med Sci. 1997;52:B39–B47. doi: 10.1093/gerona/52A.1.B39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mould J, Dulhunty A. Effects of external cadmium ions on excitation-contraction coupling in rat soleus fibres. Pflugers Arch. 1999;437:197–203. doi: 10.1007/s004240050769. [DOI] [PubMed] [Google Scholar]
- Mould J, Dulhunty A. Cadmium withdrawal contractures in rat soleus muscle fibres. Pflugers Arch. 2000;440:68–74. doi: 10.1007/s004240000266. [DOI] [PubMed] [Google Scholar]
- Nabauer M, Callawaert G, Cleeman L, Morad M. Regulation of calcium release is gated by calcium current, not gating charge, in cardiac myocytes. Science. 1989;244:800–803. doi: 10.1126/science.2543067. [DOI] [PubMed] [Google Scholar]
- Nakai J, Adams B, Imoto K, Beam K. Critical roles of the S3 segment and S3–S4 linker of repeat I in activation of L-type calcium channels. Proc Natl Acad Sci U S A. 1994;91:1014–1018. doi: 10.1073/pnas.91.3.1014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakai J, Imagawa T, Hakamata Y, Shigekawa M, Takeshima H, Numa S. Primary structure and functional expression from cDNA of the cardiac ryanodine receptor/calcium release channel. FEBS Lett. 1990;271:169–177. doi: 10.1016/0014-5793(90)80399-4. [DOI] [PubMed] [Google Scholar]
- Nishi M, Komazaki S, Kurebayashi N, Ogawa Y, Noda T, Iino M, Takeshima H. Abnormal features in skeletal muscle from mice lacking mitsugumin29. J Cell Biol. 1999;147:1473–1480. doi: 10.1083/jcb.147.7.1473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oz M, Frank G. Decrease in the size of tetanic responses produced by nitrendipine or by extracellular calcium ion removal without blocking twitches or action potentials in skeletal muscle. J Pharmacol Exp Ther. 1991;257:575–581. [PubMed] [Google Scholar]
- Pereon Y, Louboutin J-P, Noireaud J. Contractile responses in rat extensor digitorum longus muscles at different times of postnatal development. J Comp Physiol [B] 1993;163:203–211. doi: 10.1007/BF00261666. [DOI] [PubMed] [Google Scholar]
- Pereon Y, Navarro J, Sorrentino V, Louboutin J-P, Noireaud J, Palade P. Regulation of dihydropyridine receptor and ryanodine receptor gene expression in regenerating skeletal muscle. Pflugers Arch. 1997a;433:221–229. doi: 10.1007/s004240050271. [DOI] [PubMed] [Google Scholar]
- Pereon Y, Sorrentino V, Dettbarn C, Noireaud J, Palade P. Dihydropyridine receptor and ryanodine receptor gene expression in long-term denervated rat muscles. Biochem Biophys Res Commun. 1997b;240:612–617. doi: 10.1006/bbrc.1997.7712. [DOI] [PubMed] [Google Scholar]
- Protasi F, Takekura H, Wang Y, Chen S, Meissner G, Allen P, Franzini-Armstrong C. RYR1 and RYR3 have different roles in the assembly of calcium release units of skeletal muscle. Biophys J. 2000;79:2494–3508. doi: 10.1016/S0006-3495(00)76491-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Renganathan M, Messi M, Delbono O. Overexpression of IGF-1 exclusively in skeletal muscle prevents age-related decline in the number of dihydropyridine receptors. J Biol Chem. 1998;273:28845–28851. doi: 10.1074/jbc.273.44.28845. [DOI] [PubMed] [Google Scholar]
- Rios E, Brum G. Involvement of dihydropyridine receptors in excitation-contraction coupling in skeletal muscle. Nature. 1987;325:717–720. doi: 10.1038/325717a0. [DOI] [PubMed] [Google Scholar]
- Saito A, Seiler S, Chu A, Fleischer S. Preparation and morphology of sarcoplasmic reticulum terminal cisternae from rabbit skeletal muscle. J Cell Biol. 1984;99:875–885. doi: 10.1083/jcb.99.3.875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schneider M, Chandler W. Voltage dependent charge movement of skeletal muscle: a possible step in excitation-contraction coupling. Nature. 1973;242:244–246. doi: 10.1038/242244a0. [DOI] [PubMed] [Google Scholar]
- Sheridan D, Carbonneau L, Ahern C, Nataraj P, Coronado R. Ca2+ dependent excitation-contraction coupling triggered by the heterologous cardiac/brain DHPR β2a -subunit in skeletal muscle. Biophys J. 2003a;85:3739–3757. doi: 10.1016/S0006-3495(03)74790-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheridan D, Cheng W, Ahern C, Mortenson L, Alsammarae D, Vallejo P, Coronado R. Truncation of the carboxyl terminus of the dihydropyridine receptor β1a subunit promotes Ca2+ dependent excitation-contraction coupling in skeletal myotubes. Biophys J. 2003b;84:220–237. doi: 10.1016/S0006-3495(03)74844-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Soeller C, Cannell M. Examination of the transverse tubular system in living cardiac rat myocytes by 2-photon microscopy and digital image-processing techniques. Circ Res. 1999;84:266–275. doi: 10.1161/01.res.84.3.266. [DOI] [PubMed] [Google Scholar]
- Sugi H, Tsuchiya T. Muscle mechanics I: Intact single muscle fibers. In: Sugi H, editor. Current Methods in Muscle Physiology. Advantages, Problems and Limitations. New York: Oxford University Press; 1998. pp. 3–31. [Google Scholar]
- Takeshima H, Nishimura S, Matsumoto T, Ishida H, Kangawa K, Minamino N, Matsuo H, Ueda M, Hanaoka M, Hirose T, Numa S. Primary structure and expression from complementary DNA of skeletal muscle ryanodine receptor. Nature. 1989;369:439–445. doi: 10.1038/339439a0. [DOI] [PubMed] [Google Scholar]
- Takeshima H, Shimuta M, Komazaki S, Ohmi K, Nishi M, Iino M, Miyata A, Kangawa K. Mitsugumin29, a novel synaptophysin family member from the triad junction in skeletal muscle. Biochem J. 1998;331:317–322. doi: 10.1042/bj3310317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tanabe T, Adams B, Numa S, Beam K. Repeat I of the dihydropyridine receptor is critical in determining calcium channel activation kinetics. Nature. 1991;352:800–803. doi: 10.1038/352800a0. [DOI] [PubMed] [Google Scholar]
- Tanabe T, Beam K, Adams B, Niidone T, Numa S. Regions of the skeletal muscle dihydropyridine receptor critical for excitation-contraction coupling. Nature. 1990a;346:567–569. doi: 10.1038/346567a0. [DOI] [PubMed] [Google Scholar]
- Tanabe T, Beam K, Powell J, Numa S. Restoration of excitation-contraction coupling and slow calcium current in dysgenic muscle by dihydropyridine receptor complementary DNA. Nature. 1988;336:134–139. doi: 10.1038/336134a0. [DOI] [PubMed] [Google Scholar]
- Tanabe T, Mikami A, Numa S, Beam K. Cardiac-type excitation-contraction coupling in dysgenic skeletal muscle injected with cardiac dihydropyridine receptor cDNA. Nature. 1990b;344:451–453. doi: 10.1038/344451a0. [DOI] [PubMed] [Google Scholar]
- Varadi G, Lory P, Schultz D, Varadi M, Schwartz A. Acceleration of activation and inactivation by the β subunit of the skeletal muscle calcium channel. Nature. 1991;352:159–162. doi: 10.1038/352159a0. [DOI] [PubMed] [Google Scholar]
- Wang Z, Messi L, Delbono O. Patch-clamp recording of charge movement, Ca2+ current, and Ca2+ transients in adult skeletal muscle fibers. Biophys J. 1999;77:2709–2716. doi: 10.1016/s0006-3495(99)77104-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Z, Messi M, Delbono O. L-type Ca(2+) channel charge movement and intracellular Ca(2+) in skeletal muscle fibers from aging mice. Biophys J. 2000;78:1947–1954. doi: 10.1016/S0006-3495(00)76742-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Z, Messi M, Delbono O. Sustained overexpression of IGF-1 prevents age-dependent decrease in charge movement and intracellular Ca(2+) in mouse skeletal muscle. Biophys J. 2002;82:1338–1344. doi: 10.1016/S0006-3495(02)75489-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng Z, Messi M, Delbono O. Age-dependent IGF-1 regulation of gene transcription of Ca2+ channels in skeletal muscle. Mech Ageing Dev. 2001;122:373–384. doi: 10.1016/s0047-6374(00)00236-0. [DOI] [PubMed] [Google Scholar]
- Zheng Z, Wang Z, Delbono O. Charge movement and transcription regulation of L-type calcium channel α1s in skeletal muscle cells. J Physiol. 2002a;540:397–409. doi: 10.1113/jphysiol.2001.013464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng Z, Wang Z, Delbono O. Insulin-like growth factor-1 increases skeletal muscle dihydropyridine receptor α1s transcriptional activity by acting on the cAMP-response element-binding protein element of the promoter region. J Biol Chem. 2002b;277:50535–50542. doi: 10.1074/jbc.M210526200. [DOI] [PubMed] [Google Scholar]
- Zuhlke R, Pitt G, Deisseroth K, Tsien R, Reuter H. Calmodulin supports both inactivation and facilitation of L-type calcium channels. Nature. 1999;399:159–162. doi: 10.1038/20200. [DOI] [PubMed] [Google Scholar]








