首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The nuclear magnetic resonance (NMR) spectrum of Na+ is suitable for qualitative and quantitative analysis of Na+ in tissues. The width of the NMR spectrum is dependent upon the environment surrounding the individual Na+ ion. NMR spectra of fresh muscle compared with spectra of the same samples after ashing show that approximately 70% of total muscle Na+ gives no detectable NMR spectrum. This is probably due to complexation of Na+ with macromolecules, which causes the NMR spectrum to be broadened beyond detection. A similar effect has been observed when Na+ interacts with ion exchange resin. NMR also indicates that about 60% of Na+ of kidney and brain is complexed. Destruction of cell structure of muscle by homogenization little alters the per cent complexing of Na+. NMR studies show that Na+ is complexed by actomyosin, which may be the molecular site of complexation of some Na+ in muscle. The same studies indicate that the solubility of Na+ in the interstitial water of actomyosin gel is markedly reduced compared with its solubility in liquid water, which suggests that the water in the gel is organized into an icelike state by the nearby actomyosin molecules. If a major fraction of intracellular Na+ exists in a complexed state, then major revisions in most theoretical treatments of equilibria, diffusion, and transport of cellular Na+ become appropriate.  相似文献   

2.
Summary A theory for Na+, K+ and Ca2+ competitive adsorption to a charged membrane is used to explain a number of experimental observations in smooth muscle. Adsorption is described by Langmuir isotherms for mono- and divalent cations which in turn are coupled in a self-consistent way to the bulk solution through the diffuse double layer theory and the Boltzman equations. We found that the dissociation constants for binding of Na+, K+ and Ca2+ in guinea pig taenia coli areca. 0.009, 1.0, and 4×10–8 m, respectively. Furthermore, the effect of a Ca2+ pump that maintains free surface Ca2+ concentration constant is investigated. A decrease in intracellular Na+ content results in an increased Ca2+ uptake; part of this uptake is due to an increase in surface-bound Ca2+ in an intracellular compartment which is in contact with the myofilaments. Variations in the amount of charge available to bind Ca2+ and the surface charge density are studied and their effect interpreted in terms of different pharmacological agents.  相似文献   

3.
Total energy production in rabbit reticulocytes amounted to 136·52 ± 6·50μmol ATP h?1ml?1 of reticulocytes: 88·3 per cent was provided by oxidative phosphorylation, whereas only 11·7 per cent by aerobic glycolysis. Na+K+-ATPase accounted for 23 per cent, i.e. 27·65 ± 2·55μmol ATP h?1ml?1 of reticulocytes, in the overall energy consumption in reticulocytes of rabbits. Under basal conditions ATP for Na+K+-ATPase activity was derived exclusively from oxidative phosphorylation. However, when the activity of Na+K+-ATPase was increased due to the stimulation of adenylate cyclase by (?)-isoprenaline, the additional energy required was provided by aerobic glycolysis. These results indicate that two different compartments, one cytosolic and the other mitochondrial, provide energy for Na+K+-ATPase activity in reticulocytes.  相似文献   

4.
Summary A Na+-sensitive uptake of 3-O-methylglucose (3-O-MG), a nonmetabolized sugar, was characterized in frog skeletal muscle. A removal of Na+ from the bathing solution reduced 3-O-MG uptake, depending on the amount of Na+ removed. At a 3-O-MG concentration of 2mm, the Na+-sensitive component of uptake in Ringer's solution was estimated to be about 26% of the total uptake. The magnitude of Na+-sensitive component sigmoidally increased with an increase of 3-O-MG in bathing solution, whereas in Na+-free Ringer's solution the uptake was proportional to the concentration. The half saturation of the Na+-sensitive component was at a 3-O-MG concentration of about 13mm, and the Hill coefficient was 1.4 to 1.6. Phlorizin (5mm), a potent inhibitor specific for Na+-coupled glucose transport, reduced the uptake in a solution containing Na+ to the level in Na+-free Ringer's solution. Glucose of concentrations higher than 20mm suppressed 3-O-MG uptake to a level slightly lower than that in Na+-free Ringer's solution. These observations indicate that there are Na+-coupled sugar transport systems in frog skeletal muscle which are shared by both glucose and 3-O-MG.  相似文献   

5.
A. A. Rubashkin 《Biophysics》2013,58(5):660-663
A theory of change of the ionic fluxes in the lymphoid cells in their transition from normal to apoptosis we have developed previously is applied to the analysis of Na+/Na+ exchange fluxes in human lymphoid cells U937 exposed to ouabain. We solve a system of equations describing changes in the intracellular concentrations of Na+, K+ and Cl?, membrane potential and cell volume. It is shown that the Na+ influx (I Na/Na) and output flux through the Na+/Na+ tract increased 4 times in 8 h after disconnecting Na+/K+-ATPase for normal cell U937. These fluxes increased 2.6 times for apoptotic cells. The value of I Na/Na after 8 h off pump by ouabain is 97% of the total Na+ input for both cell types. It is concluded that ouabain not only inhibits the Na+/K+-ATPase, but also increases Na+ exchange fluxes through the Na+/Na+ tract, thereby switching sodium transport across the membrane of lymphoid cells to Na+/Na+ equivalent exchange.  相似文献   

6.
The existence of a subsarcolemmal space with restricted diffusion for Na+ in cardiac myocytes has been inferred from a transient peak electrogenic Na+-K+ pump current beyond steady state on reexposure of myocytes to K+ after a period of exposure to K+-free extracellular solution. The transient peak current is attributed to enhanced electrogenic pumping of Na+ that accumulated in the diffusion-restricted space during pump inhibition in K+-free extracellular solution. However, there are no known physical barriers that account for such restricted Na+ diffusion, and we examined if changes of activity of the Na+-K+ pump itself cause the transient peak current. Reexposure to K+ reproduced a transient current beyond steady state in voltage-clamped ventricular myocytes as reported by others. Persistence of it when the Na+ concentration in patch pipette solutions perfusing the intracellular compartment was high and elimination of it with K+-free pipette solution could not be reconciled with restricted subsarcolemmal Na+ diffusion. The pattern of the transient current early after pump activation was dependent on transmembrane Na+- and K+ concentration gradients suggesting the currents were related to the conformational poise imposed on the pump. We examined if the currents might be accounted for by changes in glutathionylation of the β1 Na+-K+ pump subunit, a reversible oxidative modification that inhibits the pump. Susceptibility of the β1 subunit to glutathionylation depends on the conformational poise of the Na+-K+ pump, and glutathionylation with the pump stabilized in conformations equivalent to those expected to be imposed on voltage-clamped myocytes supported this hypothesis. So did elimination of the transient K+-induced peak Na+-K+ pump current when we included glutaredoxin 1 in patch pipette solutions to reverse glutathionylation. We conclude that transient K+-induced peak Na+-K+ pump current reflects the effect of conformation-dependent β1 pump subunit glutathionylation, not restricted subsarcolemmal diffusion of Na+.  相似文献   

7.
Epithelial Na+ channel (ENaC) function is regulated by the intracellular Na+ concentration ([Na+]i) through a process known as Na+ feedback inhibition. Although this process is known to decrease the expression of proteolytically processed active channels on the cell surface, it is unknown how [Na+]i alters ENaC cleavage. We show here that [Na+]i regulates the posttranslational processing of ENaC subunits during channel biogenesis. At times when [Na+]i is low, ENaC subunits develop mature N-glycans and are processed by proteases. Conversely, glycan maturation and sensitivity to proteolysis are reduced when [Na+]i is relatively high. Surface channels with immature N-glycans were not processed by endogenous channel activating proteases, nor were they sensitive to cleavage by exogenous trypsin. Biotin chase experiments revealed that the immature surface channels were not converted into mature cleaved channels following a reduction in [Na+]i. The hypothesis that [Na+]i regulates ENaC maturation within the biosynthetic pathways is further supported by the finding that Brefeldin A prevented the accumulation of processed surface channels following a reduction in [Na+]i. Therefore, increased [Na+]i interferes with ENaC N-glycan maturation and prevents the channel from entering a state that allows proteolytic processing.  相似文献   

8.
Double quantum and triple quantum filtered 23Na nuclear magnetic resonance techniques were used to characterise in detail the isotropic and anisotropic binding and dynamics of intra- and extracellular Na+ in different cellular systems, in the absence and presence of Li+. The kinetics of Li+ influx by different cell types was evaluated. At steady state, astrocytes accumulated more Li+ than red blood cells (RBCs), while a higher intracellular Li+ concentration was found in chromaffin than in SH-SY5Y cells. Anisotropic and isotropic motions were detected for extracellular Na+ in all cellular systems studied. Isotropic intracellular Na+ motions were observed in all types of cells, while anisotropic Na+ motions in the intracellular compartment were only detected in RBCs. 23Na triple quantum signal efficiency for intracellular Na+ was SH-SY5Y > chromaffin > RBCs, while the reverse order was observed for the extracellular ions. 23Na double quantum signal efficiency for intracellular Na+ was non-zero only in RBCs, and for extracellular Na+ the order RBCs > chromaffin > SH-SY5Y cells was observed. Li+ loading generally decreased intracellular Na+ isotropic movements in the cells, except for astrocytes incubated with a low Li+ concentration and increased anisotropic intracellular Na+ movements in RBCs. Li+ effects on the extracellular signals were more complex, reflecting Li+/Na+ competition for isotropic and anisotropic binding sites at the extracellular surface of cell membranes and also at the surface of the gel used for cell immobilisation. These results are relevant and contribute to the interpretation of the in vivo pharmacokinetics and sites of Li+ action.  相似文献   

9.
Summary The purpose of this work was to determine if hypotonicity, in addition to the stimulation of active Na+ transport (Venosa, R.A., 1978,Biochim. Biophys. Acta 510:378–383), promoted changes in (i) active K+ influx, (ii) passive Na and K+ fluxes, and (iii) the number of3H-ouabain binding sites.The results indicate that a reduction of external osmotic pressure () to one-half of its normal value (=0.5) produced the following effects: (i) an increase in active K+ influx on the order of 160%, (ii) a 20% reduction in Na+ influx and K+ permeability (P K), and (iii) a 40% increase in the apparent density of ouabain binding sites. These data suggest that the hypotonic stimulation of the Na+ pump is not caused by an increased leak of either Na+ (inward) or K+ (outward). It is unlikely that the stimulation of active Na+ extrusion and the rise in the apparent number of pump sites produced by hypotonicity were due to a reduction of the intracellular ionic strength. It appears that, at least in part, the stimulation of active Na+ transport takes place whenever muscles are transferred from one medium to another of lower tonicity even if neither one was hypotonic (for instance =2 to =1 transfer). Comparison of the present results with those previously reported indicate that in addition to the number of pump sites, the cycling rate of the pump is increased by hypotonicity. Active Na+ and K+ fluxes were not significantly altered by hypertonicity (=2).  相似文献   

10.
Because intracellular [Na+] is kept low by Na+/K+-ATPase, Na+ dependence is generally considered a property of extracellular enzymes. However, we found that p94/calpain 3, a skeletal-muscle-specific member of the Ca2+-activated intracellular “modulator proteases” that is responsible for a limb-girdle muscular dystrophy (“calpainopathy”), underwent Na+-dependent, but not Cs+-dependent, autolysis in the absence of Ca2+. Furthermore, Na+ and Ca2+ complementarily activated autolysis of p94 at physiological concentrations. By blocking Na+/K+-ATPase, we confirmed intracellular autolysis of p94 in cultured cells. This was further confirmed using inactive p94:C129S knock-in (p94CS-KI) mice as negative controls. Mutagenesis studies showed that much of the p94 molecule contributed to its Na+/Ca2+-dependent autolysis, which is consistent with the scattered location of calpainopathy-associated mutations, and that a conserved Ca2+-binding sequence in the protease acted as a Na+ sensor. Proteomic analyses using Cs+/Mg2+ and p94CS-KI mice as negative controls revealed that Na+ and Ca2+ direct p94 to proteolyze different substrates. We propose three roles for Na+ dependence of p94; 1) to increase sensitivity of p94 to changes in physiological [Ca2+], 2) to regulate substrate specificity of p94, and 3) to regulate contribution of p94 as a structural component in muscle cells. Finally, this is the first example of an intracellular Na+-dependent enzyme.  相似文献   

11.
During excitation, muscle cells gain Na+ and lose K+, leading to a rise in extracellular K+ ([K+]o), depolarization, and loss of excitability. Recent studies support the idea that these events are important causes of muscle fatigue and that full use of the Na+,K+-ATPase (also known as the Na+,K+ pump) is often essential for adequate clearance of extracellular K+. As a result of their electrogenic action, Na+,K+ pumps also help reverse depolarization arising during excitation, hyperkalemia, and anoxia, or from cell damage resulting from exercise, rhabdomyolysis, or muscle diseases. The ability to evaluate Na+,K+-pump function and the capacity of the Na+,K+ pumps to fill these needs require quantification of the total content of Na+,K+ pumps in skeletal muscle. Inhibition of Na+,K+-pump activity, or a decrease in their content, reduces muscle contractility. Conversely, stimulation of the Na+,K+-pump transport rate or increasing the content of Na+,K+ pumps enhances muscle excitability and contractility. Measurements of [3H]ouabain binding to skeletal muscle in vivo or in vitro have enabled the reproducible quantification of the total content of Na+,K+ pumps in molar units in various animal species, and in both healthy people and individuals with various diseases. In contrast, measurements of 3-O-methylfluorescein phosphatase activity associated with the Na+,K+-ATPase may show inconsistent results. Measurements of Na+ and K+ fluxes in intact isolated muscles show that, after Na+ loading or intense excitation, all the Na+,K+ pumps are functional, allowing calculation of the maximum Na+,K+-pumping capacity, expressed in molar units/g muscle/min. The activity and content of Na+,K+ pumps are regulated by exercise, inactivity, K+ deficiency, fasting, age, and several hormones and pharmaceuticals. Studies on the α-subunit isoforms of the Na+,K+-ATPase have detected a relative increase in their number in response to exercise and the glucocorticoid dexamethasone but have not involved their quantification in molar units. Determination of ATPase activity in homogenates and plasma membranes obtained from muscle has shown ouabain-suppressible stimulatory effects of Na+ and K+.

Introduction: Transport and content of Na+ and K+ in skeletal muscle

The Na+,K+-ATPase (also known as the Na+,K+ pump) is the major translator of metabolic energy in the form of ATP to electrical and chemical gradients for the two most common ions in the body. These gradients enable the generation of action potentials, which are essential for muscle cell function. Evaluation of the physiological and clinical significance of the Na+,K+ pumps requires measuring the transmembrane fluxes of Na+ and K+ in intact muscles or cultured muscle cells. The simplest approach involves incubating intact muscles isolated from small animals in temperature-controlled and oxygenated buffers with electrolyte and glucose concentration comparable to that normally present in blood plasma. Initial studies used cut hemi- or quarter-diaphragm muscles from rats, mice, or guinea pigs for incubation because these muscles were considered thin enough to allow adequate oxygenation under these conditions (Gemmill, 1940). However, such preparations have numerous cut muscle ends, allowing large passive movements of Na+ and K+ and free access of Ca2+ to the cell interior. This unavoidably boosts the energy required for active transport of Na+, K+, and Ca2+, and leads to impaired cell survival. Thus, in cut muscles, the components of O2 consumption and 42K uptake attributable to the Na+,K+ pump (i.e., the fraction suppressible by the cardiac glycoside ouabain, which binds to and inhibits the Na+,K+-ATPase) have been severely overestimated. (The ouabain-suppressible components of O2 consumption or 42K uptake are measured in isolated muscles incubated without or with ouabain and calculated as the difference.) Such overestimation led to the assumption that in skeletal muscle, the Na+,K+ pumps mediate a large fraction of total energy turnover, suggesting that a major part of the thermogenic action of thyroid hormone is caused by an increased rate of active Na+,K+ transport (Asano et al., 1976). In contrast, in intact resting muscle preparations, only 2–10% of the total energy turnover is used for active Na+,K+ transport (Creese, 1968; for details, see Clausen et al., 1991). Even during maximum contractile work in human muscles, only a small fraction (2%) (Medbø and Sejersted, 1990) of total energy release is used for the Na+,K+ pumps. Thus, in skeletal muscle, the thermogenic action of the Na+,K+ pumps is modest.For the analysis of Na+,K+ transport in skeletal muscle, isolated intact limb muscles are used, primarily mammalian soleus, extensor digitorum longus (EDL), extensor digitorum brevis, or epitrochlearis muscles. These preparations can survive during incubation for many hours and can undergo repeated excitation. More recently, the isolated rat sternohyoid muscle, which also offers thin dimensions, cellular integrity, and simple handling, has been introduced (Mu et al., 2011).Measurement of muscle Na+ and K+ content requires extraction. In the past, this was done by digesting the muscle preparation in nitric acid, a sometimes risky procedure. More recently, this approach has been superseded by homogenization of the tissue in 0.3 M trichloroacetic acid, followed by centrifugation to sediment the proteins (Kohn and Clausen, 1971). The clear supernatant may then be diluted for flame photometric determination of Na+ and K+ or counting of the isotopes 22Na or 42K. Because 42K has a short half-life (12.5 h) and is of limited availability, the K+ analogue 86Rb is often used as a tracer for K+. For the quantification of Na+,K+ pump–mediated (i.e., ouabain-suppressible) transport of K+, 86Rb gives the same results as 42K (Clausen et al., 1987; Dørup and Clausen, 1994). For other flux measurements, however, the results obtained with 86Rb differ appreciably from those obtained with 42K. For example, the fractional loss of 86Rb from intact resting rat soleus muscles is only 45% of that measured using 42K. Moreover, two agents shown to stimulate the Na+,K+ pumps in isolated rat soleus muscle, salbutamol (Clausen and Flatman, 1977) and rat calcitonin gene–related peptide (CGRP) (Andersen and Clausen, 1993), both induce a highly significant stimulation of 86Rb efflux from the same muscle (Dørup and Clausen, 1994). In contrast, these same agents induced a rapid but transient (20-min duration) inhibition of the fractional loss of 42K, indicating that a large fraction of the 42K lost from the cells is reaccumulated as a result of stimulation of the Na+,K+ pumps (Andersen and Clausen, 1993; Dørup and Clausen, 1994). Finally, in skeletal muscle, bumetanide, an inhibitor of the NaK2Cl cotransporter, produces no inhibition of 42K uptake but clear-cut inhibition of 86Rb uptake (see and22 in Dørup and Clausen, 1994). Thus, results obtained with 86Rb must be verified with 42K or flame photometric measurements of changes in intracellular Na+ and K+ content. The activity of the Na+,K+ pump depends on ATP supplied by glycolysis or oxidative phosphorylation. Excitability of isolated rat soleus or EDL muscles can be maintained in the presence of the electron transport inhibitor cyanide or during anoxia (Murphy and Clausen, 2007; Fredsted et al., 2012), whereas contractions are markedly suppressed by 2-deoxyglucose, which interferes with the production of glycolytic ATP. Thus, in keeping with the studies of Glitsch (2001), these data suggest that glycolytic ATP is a primary source of energy for the Na+,K+ pumps (see also Okamoto et al., 2001). More focused studies showed that Na+,K+-pump function is not adequately supported when cytoplasmic ATP is at the normal resting concentration of ∼8 mM, but that the addition of as little as 1 mM phosphoenol pyruvate produces a marked increase in Na+,K+-pump function that is supported by endogenous pyruvate kinase bound within the t-tubular triad (Dutka and Lamb, 2007). In anoxic rat EDL muscles, contractility could be restored by stimulating the Na+,K+ pumps with the β2 agonists salbutamol or terbutaline, effects that were abolished by ouabain or 2-deoxyglucose (Fredsted et al., 2012). Glycolytic ATP furnishes energy for contractile activity under the critical condition of anoxia. Because the Na+,K+ pumps are essential for the maintenance of excitability and contractions, it would not be surprising if they were also kept going on glycolytic ATP.

Table 1.

Acute stimulation of the Na+,K+ pumps in skeletal muscle
Stimulating hormones, agents, and conditionsMechanisms of action
Epinephrine, norepinephrineStimulate generation of cAMP, which in turn activates the Na+,K+ pumps via PKA, increasing the affinity of the Na+K+ pumps for Na+
Isoproterenol, salbutamol, salmeterol
Other β2 agonists
β3 agonists
CalcitoninsStimulate generation of cAMP
CGRP
Amylin
cAMP, dibutyryl cAMPActivates Na+,K+ pumps via PKA
TheophyllineInhibits phosphodiesterase A, which degrades cAMP. This leads to intracellular accumulation of cAMP. Theophylline is a degradation product of caffeine.
Insulin, insulin-like growth factor IBoth act via the insulin receptors, increasing the affinity of Na+,K+ pumps for Na+
MonensinA Na+ ionophore that increases [Na+]i, which directly stimulates the Na+,K+ pumps
VeratridineAugments the Na+ influx per action potential, thereby increasing [Na+]i
ExcitationAugments [Na+] influx, thereby increasing [Na+]i
Increasing temperatureWhen temperature increases 10°C, the rate of Na+,K+ pumping increases 2.3-fold
ATP, ADPStimulate the Na+,K+ pumps via purinergic receptors
Open in a separate window

Table 2.

Relative changes in [3H]ouabain binding and the α2 subunit of Na+,K+-ATPase in skeletal muscle
References, muscle preparation, treatment[3H]Ouabain bindingα2-subunit abundance
Thompson et al., 2001, rats treated for 14 d with dexamethasone+22–48% (P < 0.05)+53% (P < 0.05)
Juel et al, 2001, 1 h treadmill running 150-200 g rats, giant vesicles+29% (P < 0.05)+32% (P < 0.05)
Green et al., 2004, 6 d submaximal cycling+13% (P < 0.05)+9% (P < 0.05)
Sandiford et al., 2005, electrical stimulation of rat soleus+16–21% (P < 0.05)+38% (P < 0.05)
Green et al., 2007, 16 h heavy intermittent cycling+8% (P < 0.05)+26% (P < 0.05)
Green et al., 2008, 3 d submaximal cycling+12% (P < 0.05)+42% (P < 0.05)
Green et al., 2009, chronic obstructive lung disease−6% (NS)+12% (P < 0.05)
McKenna et al., 2012, young compared to aged subjects−0.7% (NS)−24% (P < 0.05)
Boon et al., 2012, spinal cord injury in human subject−47% (P < 0.05)−52% (P < 0.05)
Chibalin et al., 2012, nicotine pretreatment in 190 g rats0% (NS)−25% (P < 0.05)
Open in a separate windowData were obtained from 10 different publications listed in the references of this paper. Each of them was selected for reporting the results of measurements of the content of [3H]ouabain-binding sites as well as α2-subunit abundance in the same muscle. Experimental details are given in the cited articles. The relative changes induced by the listed factors are given in percentages. Biopsies from human subjects were taken from the vastus lateralis muscle. The rat muscles were obtained from the hind limbs.

The Na+,K+-ATPase and why it should be quantified

The Na+,K+ pump, which is identical to the membrane-bound Na+,K+-ATPase discovered by Skou (1965), has been found in most skeletal muscle types from many species. As illustrated in Fig. 1, the Na+,K+-pump molecule comprises a catalytic α subunit, a β subunit involved in its translocation to the plasma membrane, and a regulatory subunit (FXYD; in skeletal muscle named FXYD1 or phospholemman). The Na+,K+ pumps in skeletal muscle are subject to acute activation or inhibition of their transport rate as well as long-term regulation of their content (expressed in molar units, usually as pmol per gram wet weight) (Clausen, 2003). In human skeletal muscle, the content of Na+,K+ pumps measured using [3H]ouabain binding is ∼300 pmol/g wet wt (Clausen, 2003). In an adult human subject weighing 70 kg, the muscles will weigh around 28 kg and therefore contain a total of 28,000 × 300 pmol = 8.4 µmoles of Na+,K+ pumps. Muscles represent the largest single pool of cells in the human body, containing the largest pool of K+ (2,600 mmol; Clausen, 2010) and one of the largest pools of Na+,K+ pumps (Clausen, 1998). Many diseases are associated with anomalies in the content of Na+,K+ pumps in skeletal muscle (Clausen, 1998, 2003), and during the last decades, the number of published articles on Na+,K+ pumps in skeletal muscle has increased appreciably. The capacity of the Na+,K+ pumps in muscles is crucial for the clearance of extracellular K+ at rest or during exercise, and details of the transport capacity are required for clinical evaluation of the risk of developing hypo- or hyperkalemia (for clinical examples and details, see Clausen, 1998, 2003, 2010; Sejersted and Sjøgaard, 2000). Moreover, this information is key to understanding the functional significance of physiological changes or pathological anomalies in the content of Na+,K+ pumps in skeletal muscle. The Na+,K+-pump capacity may be calculated by multiplying the number of Na+,K+ pumps by their turnover number (determined with the Na+,K+-ATPase from ox brain at 37°C as 8,000 molecules of ATP split per Na+,K+-ATPase molecule per min; Plesner and Plesner, 1981) and by the number of K+ ions transported per cycle, which is normally 2. From this information it can be calculated that the total transport capacity for K+ of the pool of Na+,K+ pumps in human skeletal muscle amounts to 2 × 8,000 × 8.4 µmoles/min = 134 mmoles/min. Provided that all the Na+,K+ pumps in all muscles are operating at maximum speed (which rarely happens), they could clear all extracellular K+ in the human body (56 mmoles, calculated by multiplying the extracellular water space [14 liters] by the content of K+ [4 mmol/liter] in 25 s; Clausen, 2010). The present Review was prompted by the ongoing problems of Na+,K+-pump quantification and recovery, how they can be solved, and what insight may be gained by using more specific methods and adequate quantitative analysis. The early development of the field has been discussed previously (Ewart and Klip, 1995; Sejersted and Sjøgaard, 2000; Clausen, 2003). This Review focuses on the most relevant publications appearing during the last 10 years; to cover the background, it includes references back to the discovery of the Na,K pump in 1957 (Skou, 1957).Open in a separate windowFigure 1.The molecular structure of Na+,K+-ATPase. The figure is based upon the crystal structure of the homologous Na+,K+-ATPase in the E2P2K conformation (Shinoda et al., 2009) and drawn by Flemming Cornelius (Aarhus University, Aarhus, Denmark). The Na+,K+ pump comprises an α and a β subunit, a glycoprotein that participates in the translocation of the molecule from the cell interior to its correct position in the lipid bilayer of the plasma membrane. A regulatory subunit (FXYD) is also shown. During each transport cycle of the Na+,K+ pump, one ATP molecule is bound to the cytoplasmic site of the α subunit; its hydrolysis provides energy for the active transport of Na+ and K+. The transmembrane domain consists of 10 transmembrane helices and contains the binding sites for three Na+ or two K+ ions, respectively, which pass sequentially through the same cavity in the molecule during each transport cycle.

Problems in assessing Na+,K+-ATPase activity and recovery.

When the classical assay for membrane-bound ATPase activity stimulated by Na+ and K+ (Skou, 1965) and blocked by ouabain is used with crude muscle homogenates, there is an overwhelming background of other ATPases. Therefore, during the early phase of Na+,K+-pump investigation (1957–1977), many research groups were primarily interested in the detection and purification of the Na+,K+-ATPase. The differential centrifugation procedures developed during that period often discarded a major part of the enzyme activity present in homogenates of the intact tissue to eliminate other ATP-splitting cellular components, such as myosin, Ca2+-activated ATPase, and unspecified Mg2+-activated ATPase, and obtain a pure enzyme. Therefore, accurate information about Na+,K+-ATPase recovery was not always available, and these methods were not suitable for quantification of the total content of Na+,K+ pumps in the intact tissue. Moreover, the low recovery of Na+,K+-ATPase activity often obtained raised doubts as to whether the Na+,K+-ATPase in the samples tested was representative of the entire population of enzyme molecules present in the starting tissue. These problems were analyzed in a review, which found that, among 12 papers, only 3 obtained a recovery >5% of the Na+,K+-ATPase present in the starting material and, in 9 of these papers, recovery was in the range of 0.2 to 3.5% (Hansen and Clausen, 1988). To obtain better purification of the Na+,K+-ATPase in sarcolemma, giant vesicles were later produced from the sarcolemmal membranes of rat hind-limb muscles. However, they only contained 0.3% of the total content of Na+,K+ pumps present per gram of muscle (Juel et al., 2001). Thus, it is difficult to evaluate to what extent down-regulation of the content of Na+,K+ pumps might reduce Na+,K+ pump–mediated K+ uptake in skeletal muscle and thereby impair the clearance of K+ from extracellular space and plasma.The Na+,K+-ATPase activity has been measured in plasma membranes prepared from various rat muscles. In total membranes, Na+ (0–80 mM) and K+ (0–10 mM) were shown to activate ouabain-suppressible ATPase (Juel, 2009). Surprisingly, Vmax for Na+ given as µmoles of ATP split per milligram protein per hour was higher in total membranes than in the sarcolemmal membranes, which might be expected to show a higher value, because of partial purification of the Na+,K+-ATPase. In total membranes, as well as in plasma membranes, 30 min of running gave only a minor increase in Vmax, which was only significant in a few instances, suggesting only modest translocation of the Na+,K+ pumps from an intracellular pool of endosomal membranes to the plasma membrane (see the section below, Translocation of Na+,K+ pumps compared…, for alternatives to translocation) (Juel, 2009). Because of a lack of accurate information about plasma membrane recovery, however, these observations could not readily be translated to pmol of Na+,K+ pumps per gram wet weight.

3-O-methylfluorescein phosphatase (3-O-MFPase) assay.

Using 3-O-methylfluorescein phosphate as a substrate, K+-dependent and ouabain-suppressible phosphatase activity has been measured in a fluorimetric assay for the Na+,K+-ATPase in skeletal muscle (Nørgaard et al., 1984). Early studies on rat skeletal muscle showed good agreement between the 3-O-MFPase activity of crude homogenates and the number of [3H]ouabain-binding sites measured in the same intact muscles or biopsies thereof (Hansen and Clausen, 1988), suggesting that this assay provided a reliable measure of Na+,K+-ATPase activity. Moreover, the 3-O-MFPase activity showed similar decreases with aging or K+ depletion of the rats as the [3H]ouabain-binding capacity. However, studies on biopsies of human vastus lateralis muscle showed that exercise to fatigue reduced 3-O-MFPase activity by 13% (P < 0.001), whereas [3H]ouabain binding remained constant (Leppik et al., 2004). Several other studies have also reported activity-dependent decreases in 3-O-MFPase activity in muscle: Another study from the same laboratory showed that acute exercise induced a 22% drop in the 3-O-MFPase activity of human vastus lateralis muscle (McKenna et al., 2006). In human vastus lateralis muscle, 16 h of heavy intermittent cycle exercise induced a 15% reduction in 3-O-MFPase (Green et al., 2007). In rats, running exercise induced a 12% decrease in 3-O-MFPase in all muscles (Fowles et al., 2002). However, other studies have not. For instance, electrical stimulation of the isolated rat soleus muscle for 15 or 90 min induced a 40 or 53% increase in 3-O-MFPase activity of crude homogenate, respectively (Sandiford et al., 2005). In keeping with this, giant vesicles representing sarcolemmal membrane obtained from rat muscles after 3 min of running exercise showed a 37% increase in 3-O-MFPase activity (Kristensen et al., 2008). Moreover, a full Na+,K+-ATPase assay showed that 30 min of treadmill running increased Vmax by 12 to 39% in sarcolemmal membranes from rat hind-limb muscles (Juel, 2009). In contrast, high frequency stimulation of isolated rat EDL muscles caused no significant change in the content of 3-O-MFPase (Goodman et al., 2009). These discrepancies indicate that measurements of 3-O-MFPase provide inconsistent information about the content of Na+,K+-ATPase in skeletal muscle. Moreover, as pointed out by Juel (2009), the 3-O-MFPase cannot be used to measure Na+-dependent activation. Therefore, further use of the 3-O-MFPase assay requires closer analysis of the causes of these discrepancies.

[3H]Ouabain binding.

[3H]Ouabain and other labeled cardiac glycosides bind stoichiometrically to the extracellular surface of the α subunit of the Na+,K+ pump (one drug molecule per Na+,K+-ATPase molecule). Therefore, incubation of tissues, cells, or plasma membranes with [3H]ouabain enables the quantification of the content of Na+,K+-ATPase in molar units by liquid scintillation counting of extracts of the tissue, cells, or membranes (Clausen and Hansen, 1974, 1977; Hansen and Clausen, 1988). Because K+ interferes markedly with the binding of cardiac glycosides, studies of [3H]ouabain binding to intact rat, mouse, or guinea pig muscles are performed using K+-free (KR) buffer (Clausen and Hansen, 1974). [3H]Ouabain binding is saturable and reversible without showing [3H]ouabain penetration into the intracellular space. In isolated soleus muscle obtained from 4-wk-old rats, [3H]ouabain binding amounts to 720 pmol/g wet wt, corresponding to 3,350 molecules of Na+,K+-ATPase per µ2 of sarcolemma (not including t-tubules). When injected intraperitoneally, [3H]ouabain binds rapidly to the outer surface of the muscle cells, showing saturation and the same content of [3H]ouabain-binding sites per gram tissue wet weight as measured in intact muscles incubated with [3H]ouabain in vitro (Clausen et al., 1982; Murphy et al., 2008). The [3H]ouabain binding in vivo is faster than in vitro, reaching saturation in 20 min, partly because of the higher temperature and better access to the muscle cells via the capillaries. When bound to the outer surface of intact muscles, [3H]ouabain is rapidly displaced by the addition of an excess of unlabeled ouabain during a subsequent wash performed both after binding in vitro and in vivo, indicating that the [3H]ouabain is not internalized into the cytoplasm (Clausen and Hansen, 1974; Clausen et al., 1982).It has been estimated that, in rat skeletal muscle, most of the Na+,K+ pumps (around 80%) are the α2-subunit isoform, which has high affinity for [3H]ouabain. A minor fraction (around 20%) is the α1 isoform, which has a lower affinity for [3H]ouabain (Hansen, 2001) and therefore may not be detected at the concentration of [3H]ouabain used in the standard assay for quantification of [3H]ouabain-binding sites (10−6 to 5 × 10−6 M). However, these values for α1 and α2 isoforms depend on the antibody used and its interaction with newly synthesized α subunits and are therefore inconclusive (for detailed discussion, see Clausen, 2003). Another study in which relative subunit composition was determined with antibodies indicated that in mouse EDL muscle, the α2 subunit accounted for 87% of the total amount of the α subunit and the α1 subunit for 13% (He et al., 2001). More importantly, measurements of the maximum rate of ouabain-suppressible active 86Rb uptake in Na+-loaded rat soleus muscles have given values in good agreement with those obtained in the standard [3H]ouabain-binding assay, indicating that this assay quantifiably measures the majority of the total content of functional Na+,K+ pumps (Clausen et al., 1987).The phosphate analogue vanadate (VO4) binds to the intracellular surface of the Na+,K+ pumps and thereby facilitates [3H]ouabain binding to the outer surface of the Na+,K+ pumps in the plasma membranes (Hansen and Clausen, 1988). This enabled development of an assay in which cut muscle tissue segments are incubated with [3H]ouabain in a Tris buffer–containing Tris vanadate (Nørgaard et al., 1983). Because the muscle segments are cut, vanadate gains ready access to the cytoplasm and can bind to the inner surface of the Na+,K+-ATPase. This assay gives the same values for [3H]ouabain binding as measurements of [3H]ouabain binding in intact muscles incubated in K+-free KR. A comparison of the binding kinetics of [3H]ouabain obtained using the vanadate-facilitated assay showed that the α1-, α2-, and α3-subunit isoforms of the pumps in human tissues have similar affinity for [3H]ouabain (Wang et al., 2001). This indicates that the vanadate-facilitated [3H]ouabain-binding assay quantifies the sum of the three Na+,K+-pump isoforms present in human muscle biopsies (for details, see Clausen, 2003). Another major advantage of the vanadate-facilitated [3H]ouabain-binding assay with small (2–5-mg) cut muscle specimens is that it can be used to measure Na+,K+ pumps in frozen muscle samples. Storage in the freezer for at least 4 yr causes no change in the content of [3H]ouabain-binding sites of human muscle samples, and samples may be sent on dry ice by air freight. The assay is rapid, efficient, and inexpensive, and the results are reproducible. 17 studies on human skeletal muscle biopsies performed in seven different laboratories obtained similar values for the content of [3H]ouabain-binding sites of human skeletal muscle samples (in the range of 243 to 425 pmol/g wet wt; see Clausen, 2003). More recent measurements of [3H]ouabain-binding sites in human vastus lateralis muscle performed in four different laboratories showed results in the same range (326 ± 30 pmol/g wet wt [Nordsborg et al., 2005]; 272 ± 10 pmol/g wet wt [Green et al., 2004]; 240 ± 10 pmol/g wet wt [Boon et al., 2012]; 350 ± 108 pmol/g wet wt [McKenna et al., 2012]). [3H]ouabain binds to both the outer surface of sarcolemma and the inner surface of the t-tubular lumen. Lau et al. (1979) showed that t-tubules from rabbit muscle bind [3H]ouabain with the same affinity (dissociation constant around 5 × 10−8 M) and capacity (700 pmol/g wet wt) as the intact rat soleus muscle (Clausen and Hansen, 1974).The rate of [3H]ouabain binding depends on the rate of active Na+,K+ transport (Clausen and Hansen, 1977; Andersen and Clausen, 1993; Clausen, 2000, 2003). [3H]Ouabain binding requires a certain configuration of the binding site on the extracellular portion of the α subunit that occurs only once during each Na+,K+-pumping cycle (Schwartz et al., 1975; Clausen and Hansen, 1977). Hence, an increased rate of cycling augments the chances for the [3H]ouabain molecule to bind. Therefore, an increase in the rate of [3H]ouabain binding likely indicates that the Na+,K+-pumping rate is augmented.

Regulation of the Na+,K+ pumps in skeletal muscle

Regulation of the Na+,K+ pumps evaluated by measurement of Na+ and K+ fluxes and [3H]ouabain binding. Short-term regulation of Na+,K+-pump activity.

The Na+ ionophore monensin (10−5 M) induces a graded increase in Na+ influx in skeletal muscle, enabling the evaluation of the effects of increased intracellular Na+ on the rate of active Na+,K+ transport (measured as ouabain-suppressible 42K or 86Rb influx) and of possible changes in pump affinity for intracellular Na+ (Buchanan et al., 2002). The Na+,K+ pumps undergo short-term and long-term regulation, often defined as acute changes in transport activity (e.g., µmoles of Na+ extruded per gram wet weight per minute) or slower and more long-lasting changes in the content of Na+,K+ pumps (expressed as pmol of [3H]ouabain bound per gram wet weight), respectively. In skeletal muscle, the most common cause of an acute increase in Na+,K+-pump activity is excitation-induced rise in intracellular Na+ caused by increased Na+ influx. This and around 20 other conditions or agents shown to stimulate the Na+,K+ pump in skeletal muscle are listed in Hodgkin and Horowicz, 1959). In close agreement with this value, intact rat EDL muscle stimulated for 5 s at 90 Hz to produce isometric tetanic contraction showed a Na+ influx of 12 nmol/g wet wt/action potential and a K+ efflux of 10 nmol/g wet wt/action potential, close to the 1:1 exchange of Na+ and K+ causing the action potential (Clausen et al., 2004). The excitation-induced increase in intracellular Na+ primarily takes place at the inner surface of the plasma membrane, which is close enough to the Na+,K+ pumps to induce prompt activation of their transport activity. An early increase in Na+,K+-pump activity may also involve an excitation-dependent increase in Na+,K+-pump affinity for [Na+]i that is independent of the rise in intracellular Na+ per se (Buchanan et al., 2002). Thus, as shown in Fig. 2, electrical stimulation induces a leftward shift of the curve relating intracellular Na+ to Na+,K+ pump–mediated 86Rb uptake. Electrical stimulation of rat soleus for 10 s at 60 Hz at 30°C induced an immediate 58% increase in intracellular Na+ content. During subsequent rest at 30°C, reextrusion of Na+ was complete in 2 min, and this was followed by a statistically significant undershoot in [Na+]i (P < 0.001) (Everts and Clausen, 1994). During the rapid early decrease in Na+ content, it can be assumed that activity of the Na+,K+ pumps is stimulated 15-fold, even though [Na+]i has decreased from its peak value down to the range found in resting muscle. The excitation-induced increase in Na+ affinity was also detected by a highly significant (P < 0.001–0.05) decrease in intracellular Na+ of ∼25% that lasted up to 30 min after 60 s of electrical stimulation of isolated rat soleus muscles. Similar poststimulatory undershoots in [Na+]i was seen in rat EDL muscle. It was blocked by ouabain or cooling to 0°C and was assumed to be mediated by activation of the Na+,K+ pumps by CGRP released from nerves in the muscles. Thus, reducing CGRP content by capsaicin or by prior denervation of the muscles prevented both excitation-induced force recovery and the drop in intracellular Na+ (Nielsen and Clausen, 1997). This provides further evidence that activity-dependent stimulation of the Na+,K+ pump in muscle is not solely a function of the rise in [Na+]i but might be caused by a mechanism similar to that in the electric organ of Narcine brasiliensis, where electrical stimulation markedly augments the activity of the Na+,K+-ATPase within fractions of a second despite a minimal change in intracellular Na+ (Blum et al., 1990).Open in a separate windowFigure 2.Electrical stimulation affects the relationship between [Na+]i and 86Rb+ uptake rate. Rat soleus muscles were mounted isometrically on electrodes and were stimulated for 10 s at 60 Hz. After a 2-min rest, they were transferred into solutions containing 86Rb+, incubated for 2 min before being washed to remove extracellular 86Rb and Na+, and blotted, and intracellular 86Rb+ and [Na+] were determined. [Na+]i was manipulated by preincubating in buffer in which the Na+ content had been reduced or in standard KR buffer containing the Na+ ionophore monensin. •, resting muscles; ○, muscles stimulated. Each point represents the mean ± SEM of measurements performed on three muscles. The curves were fitted using a computer program. Reprinted with permission from The Journal of Physiology (Buchanan et al., 2002).The slowly hydrolyzed cholinergic antagonist carbachol induces a long-lasting activation of the nicotinic acetylcholine receptors. Our experiments showed that in isolated rat soleus muscle, carbachol augments Na+ influx leading to depolarization, increased intracellular Na+, and loss of force (Macdonald et al., 2005). This reduction in force is significantly (P < 0.05) restored by stimulating the Na+,K+ pumps with epinephrine, salbutamol, or CGRP. All these effects are likely to be indirect, mediated by cAMP generated by stimulation of the adenylate cyclase. CGRP is found in sensory nerve endings from which it may be released by capsaicin or electrical stimulation to stimulate the Na+,K+ pumps in the muscle cells (Nielsen et al., 1998). Thus, CGRP release and the ensuing stimulation of pump activity may contribute to the above-mentioned long-lasting drop in intracellular Na+ caused by electrical stimulation. Because the excitation-induced force recovery seen at increased [K+]o is suppressed by the CGRP analogue CGRP-(8–37), which acts as a competitive CGRP antagonist, it is likely mediated by local release of CGRP from nerve endings in the muscle (Macdonald et al., 2008). In the isolated rat soleus muscle, repeated electrical stimulation caused hyperpolarization (11 mV) and increased amplitude of the action potentials. These effects were abolished by ouabain, cooling, or omission of K+ from the buffer, suggesting that they resulted from of the electrogenic Na+,K+ pumps (Hicks and McComas, 1989) (see below in The electrogenic action of the Na+,K+ pumps).As listed in Clausen and Flatman, 1977). Like electrical stimulation (Fig. 2), salbutamol causes a leftward shift of the curve relating intracellular Na+ to Na+,K+ pump–mediated uptake of 86Rb in rat soleus muscle (Buchanan et al., 2002). The rapid stimulating effect of the above-mentioned agents on the Na+,K+ pumps, their mechanisms, and the physiological importance have been described in detail (Clausen, 2003). The stimulatory effects of these agents have been exploited in the treatment of hyperkalemia and the ensuing muscular weakness or paralysis. Thus, salbutamol was introduced for the treatment of paralytic attacks in patients with hyperkalemic periodic paralysis (Wang and Clausen, 1976). In muscles isolated from knock-in mice with the same genetic anomaly, stimulation of the Na+,K+ pumps with salbutamol or CGRP restored muscle force (Clausen et al., 2011). These observations suggest that the beneficial effect of mild exercise on severe weakness seen during attacks of hyperkalemic periodic paralysis is likely related to the stimulatory effect of locally released CGRP on the Na+,K+ pumps.In rat soleus muscle, ATP induces a twofold increase in Na+,K+ pump–mediated 86Rb uptake (Broch-Lips et al., 2010). ATP induces a marked recovery of force and M-waves (the sum of action potentials as recorded from the surface of the muscle) in muscles inhibited by increasing [K+]o, indicating that ATP-induced stimulation of the Na+,K+ pumps enhances excitability. Similar effects are exerted by ADP and are mediated by purinergic receptors. These effects are of particular interest during intense exercise, where local release of ATP or ADP from the muscle cells may reduce the inhibitory effect on excitability of the concomitant increase in [K+ ]o (Broch-Lips et al., 2010). The importance of the electrogenic action of the Na+,K+ pumps is also evident from the observation that at increased [K+]o, where M-wave amplitude and area are decreased, stimulation with salbutamol or ATP restores the M-waves (Overgaard et al., 1999; Broch-Lips et al., 2010). In conclusion, the short-term increase in Na+,K+-pump activity in response to excitation has at least two components: (1) a direct stimulatory effect of [Na+]i on the Na+,K+ pump; and (2) an increase in the affinity of the Na+,K+ pump for Na+, potentiating the stimulating effect of [Na+]i. This seems to be mediated by cAMP, protein kinases, or ATP, and may involve release of a local hormone (CGRP). For more details, see below in Molecular mechanisms of pump regulation.

Long-term regulation of Na+,K+-pump content.

Long-term increases in the muscle content of Na+,K+ pumps is most commonly seen after training. As measured using [3H]ouabain binding, such an increase has been observed in nine different species and in numerous studies on human and rat muscle (see Clausen, 2003). The relative increase observed in the different studies depends on the duration and intensity of the training and ranges from 14 to 46%. There is a great deal of evidence for such changes in Na+,K+-pump content, which can also be induced by electrical stimulation. For instance, the content of [3H]ouabain-binding sites in the tibialis anterior muscle of the rabbit was more than doubled after 20 d of electrical stimulation in vivo. Furthermore, it was correlated to the amplitude of M-waves (P < 0.01; r = 0.8) (Hicks et al., 1997). In pigs in which shivering is induced by lowered temperature, [3H]ouabain binding is increased by 58–84% after weeks (for details, see Clausen, 2003). Running 100 km in 10.7 h produced a significant (P < 0.05) 13% increase in the content of [3H]ouabain-binding sites in human vastus lateralis muscle (Clausen, 2003). Cycling exercise for 3 d increased the content of [3H]ouabain-binding sites in human vastus lateralis muscle by 8% (P < 0.05) (Green et al., 2004). A recent study showed that rats performing voluntary free wheel running covered a distance of 13 km/day. After 8 wk, their soleus muscles contained 22% more [3H]ouabain-binding sites than those from untrained controls. Moreover, the soleus muscle of the trained rats showed considerably better contractile tolerance to increased [K+]o (9 mM) than those from untrained controls (Broch-Lips et al., 2011). In rats exposed to hypoxia for 6 wk (causing chronic stimulation of respiration), the content of [3H]ouabain-binding sites in the diaphragm was increased by 24% and the contractile endurance of the muscle was significantly augmented (McMorrow et al., 2011).Conversely, a decrease in the muscle content of Na+,K+ pumps is seen during reduced activity or immobilization. In rats, guinea pigs, sheep, and humans, immobilization reduced the content of [3H]ouabain-binding sites by 20–27%, changes that were reversible after mobilization (Clausen, 2003). Thus, after 2 wk of decreased mobility, the content of [3H]ouabain-binding sites in guinea pig gastrocnemius muscles was 258 ± 13 pmol/g wet wt (25% below that of the contra-lateral freely mobile muscle; P < 0.02). After 3 wk of reduced mobility and 3 wk of training by running, the content of [3H]ouabain-binding sites in the gastrocnemius muscle of these animals increased by 57%, whereas those that had not been immobilized reached 498 ± 25 pmol/g wet wt, which is 93% higher than that of the muscles with 2 wk of reduced mobility (Leivseth et al., 1992). This indicates that the regulatory range of Na+,K+-pump content in muscle represents roughly a doubling from the immobilized to the trained muscle. These results are particularly notable because, compared with rats or mice, all the Na+,K+ pumps in guinea pig muscle have high affinity for [3H]ouabain, indicating complete occupancy of the [3H]ouabain-binding sites.Immobilization induced by denervation, a plaster cast, or tenotomy (tendon release) reduced the contents of [3H]ouabain-binding sites in the skeletal muscles of mice, rats, guinea pigs, and sheep (Clausen, 2003; Clausen et al., 1982). In keeping with this, in human subjects complete spinal cord injury and partial denervation caused 58% reduction in the content of [3H]ouabain-binding sites in the vastus lateralis muscle (Ditor et al., 2004). A more recent study confirmed this observation, showing ∼50% reduction in the content of [3H]ouabain-binding sites in skeletal muscles of patients with complete cervical spinal cord injury (Boon et al., 2012). The mechanisms of exercise-related increases or decreases in the content of Na+,K+ pumps was explored in chick embryo leg muscle using monoclonal antibodies for estimating the relative changes in the number of Na+,K+ pumps. Increased intracellular Na+ induced by activating the Na+ channels with veratridine induced a 60–100% increase in the content of Na+,K+ pumps. Conversely, inhibition of the Na+ channels with tetrodotoxin blocked this effect and induced a rapid decrease (Fambrough et al., 1987).

Hormonal regulation of Na+,K+-pump content.

Thyroid hormone exerts the most potent hormonal stimulation of the synthesis of Na+,K+ pumps in skeletal muscle (Asano et al., 1976; for details, see Clausen, 2003). In rats, daily subcutaneous injection of thyroid hormone increases the content of [3H]ouabain-binding sites by 75 pmol/g muscle wet wt/d (Everts and Clausen, 1988). The content of [3H]ouabain-binding sites in the human vastus lateralis muscle varied from 100 pmol/g wet wt to 550 pmol/g wet wt in hypothyroid, euthyroid, and hyperthyroid subjects, with close linear correlation to free T4 index (r = 0.87; P < 0.001) (Clausen, 2003). [3H]ouabain binding in biopsies from vastus lateralis muscle of nine hyperthyroid patients was 89% higher value than in those form euthyroid controls. This increase in [3H]ouabain binding, the concomitant increase in energy expenditure, and plasma thyroid hormone concentration were all restored to the control levels by the standard treatment of hyperthyroidism with the anti-thyroid drug methimazole, which inhibits the synthesis of thyroid hormone (Riis et al., 2005).During fasting, the plasma concentration of thyroid hormones decreases (Spencer et al., 1983). In rats, total fasting for 72 h led to an ∼50% decrease in the content of [3H]ouabain-binding sites in plasma membranes obtained from soleus muscle (Swann, 1984). 3 wk on reduced caloric intake caused a 45–53% decrease in rat plasma thyroid hormone concentration and a 25% decrease in the content of [3H]ouabain-binding sites in the skeletal muscles (see Clausen, 2003). Because reduced caloric intake also decreases muscle mass, K+ clearance from the extracellular space is likely to be further impaired. Worldwide, a starvation-induced decrease in Na+,K+-pump content is probably the most common muscle Na+,K+-pump disorder in human subjects, causing reduced tolerance to the hyperkalemia arising during K+ ingestion and intensive work, leading to impaired physical performance. However, there is no information available about the effects of starvation on the content of Na+,K+ pumps in human skeletal muscle. The increase in Na+,K+-pump content in rat skeletal muscle induced by injection of thyroid hormone is preceded by an increase in the content of Na+ channels in the muscles measured using 3H-labeled saxitoxin as well as increased intracellular Na+ measured by flame photometry (Clausen, 2003). Increased Na+ influx in resting soleus muscle also induces an increase in the Na+,K+-pump content in knock-in mice with hyperkalemic periodic paralysis (Clausen et al., 2011). These observations are in keeping with the stimulatory effect of increased intracellular Na+ on the synthesis of Na+,K+ pumps in cultured muscle cells (Fambrough et al., 1987). Adrenal steroids also influence the content of Na+,K+ pumps in skeletal muscle. Thus, in 36 patients treated for chronic obstructive lung disease with the glucocorticoid dexamethasone, the content of [3H]ouabain-binding sites in needle biopsies of the vastus lateralis muscle was 31% higher than in 23 age- and sex-matched control subjects (P < 0.001) (Ravn and Dørup, 1997). In rats, 7–14 d of continuous infusion of dexamethasone via osmotic mini-pumps induced a 27–42% increase (P < 0.001– 0.01) in the content of [3H]ouabain-binding sites in soleus, EDL, gastrocnemius, and diaphragm muscles. In contrast, acute stimulation of Na+,K+-pump activity by infusion of the β2 agonist terbutaline produced no significant change in [3H]ouabain binding. A more detailed study (Dørup and Clausen, 1997) showed that the up-regulation of Na+,K+ pumps induced by dexamethasone could not be attributed to a mineralocorticoid action. Indeed, the mineralocorticoid aldosterone induced a decrease in the content of [3H]ouabain-binding sites in rat skeletal muscle, which was closely correlated to a concomitant reduction in the content of K+ in the muscles, induced by aldosterone stimulation of renal K+ excretion This is in keeping with the down-regulation of Na+,K+ pumps in skeletal muscle observed in rats maintained on K+-deficient fodder and in patients developing K+ deficiency during treatment with diuretics (Clausen, 1998). In rat skeletal muscles, dexamethasone infusion for 14 d induced a relative rise in Na+,K+-ATPase α2-subunit isoform of 53–78%, and in mRNA for the α2 subunit a 6.5-fold increase was found (Thompson et al., 2001). In young human subjects, the ingestion of dexamethasone (2 mg twice daily for 5 d) increased the content of [3H]ouabain-binding sites by 18% in the vastus lateralis muscle (P < 0.001) and by 24% in the deltoid muscle (P < 0.01). In the same subjects, the content of 3-O-MFPase in vastus lateralis and deltoid muscles increased by 14 and 18%, respectively (P < 0.05) (Nordsborg et al., 2005). Another study by the same group showed that the same dose of dexamethasone induced 17% relative increase in α1- and α2-subunit expression (P < 0.05) in vastus lateralis muscle, a significantly lower exercise-induced net K+ release from the thigh muscles and a borderline significant prolongation of time to exhaustion (P = 0.07) (Nordsborg et al., 2008).In rats made diabetic by streptozotocin pretreatment, the content of [3H]ouabain-binding sites in skeletal muscle was reduced by 24% in soleus and 48% in EDL (Schmidt et al., 1994). These changes were completely restored by insulin treatment (Kjeldsen et al., 1987).

Translocation of Na+,K+ pumps compared with other types of activation.

In skeletal muscle, the Na+,K+-pump molecules are located in the sarcolemma (primarily the α1-subunit isoform) and in the t-tubular membranes (primarily the α2-subunit isoform) (Marette et al., 1993; Williams et al., 2001; Cougnon et al., 2002; Radzyukevich et al., 2013). This localization seems to be strategic, reflecting a particular need for transport of Na+ and K+ across the t-tubular membranes during and after work. Thus, a large fraction of the Na+,K+ exchange takes place via the t-tubular membranes, and because of the small volume of the t-tubules, the intra-tubular concentration of K+ is likely to reach high levels (36.6 mM in 1 s of stimulation at 100 Hz, sufficient to cause severe interference with excitability; Kirsch et al., 1977). A recent simulation study on single rat EDL muscle fibers showed that stimulation at 30 Hz increased the K+ concentration in the t-tubules to a plateau of 14 mM within 1 s (Fraser et al., 2011). Therefore, during intense work, there is an urgent need for increased active transport of K+ out of the t-tubular lumen and into the cytoplasm. The most efficient way of meeting this demand would be augmented affinity of the Na+,K+ pumps for Na+ and/or K+. Alternatively, the entire Na+,K+-pump molecule might move to positions in the plasma membrane where there is more ready access to Na+ or K+. It has long been known that insulin induces a 1.7-fold increase in the binding of [3H]ouabain in frog sartorius muscle (Erlij and Grinstein, 1976) that was proposed to reflect an “unmasking” of an intracellular pool of inactive Na+,K+ pumps. Furthermore, Tsakiridis et al. (1996) showed that in rats, 60 min of treadmill running induced a significant relative increase (43–94%) in the α2 polypeptide of the Na+,K+-ATPase (now termed the α2-subunit isoform) detected by immunoblotting of plasma membranes prepared from hind-limb muscles. This was interpreted as indicating that exercise might induce translocation from an intracellular pool of endosomal membranes to the plasma membrane. Surprisingly, however, there was no concomitant decrease in the α2 polypeptide detectable in the intracellular pool, and there was no information about the recovery of the Na+,K+ pumps in the plasma membranes (Tsakiridis et al., 1996). A later study used differential centrifugation of a whole homogenate to separate sarcolemmal membranes from endosomal membranes (Sandiford et al., 2005). After electrical stimulation in vivo via the nerves, the maximum 3-O-MFPase activity in the sarcolemmal fraction had increased by ∼40%. In the same fraction, the α2 subunit had increased by 38–40%, and in the endosomal fraction, the α2 subunit had decreased by 42%, in keeping with a translocation of Na+,K+ pumps from the endosomal membranes to the sarcolemmal membranes. However, in a more recent study, Na+,K+ pumps could barely be detected in the intracellular pool of membranes (the putative source for the translocation of Na+,K+ pumps to the plasma membrane) (Zheng et al., 2008). In rats, 60 min of treadmill running induced a 29% increase in [3H]ouabain labeling of sarcolemmal giant vesicles obtained from mixed hind-limb muscles, as well as 19–32% increases in the contents of the α1-2 and β1-2 subunits (Juel et al., 2001). The increase in subunits was reversible with a half-life of 20 min, but there is no information about the reversibility of the [3H]ouabain binding. However, as mentioned in the paper (Juel et al., 2001), the content of Na+,K+ pumps in the vesicular membranes measured as [3H]ouabain-binding sites was only 0.3% of the total content of Na+,K+ pumps in rat hind-limb muscles. Therefore, it could not be concluded that the samples of Na+,K+ pumps in the giant vesicles were representative of the entire pool of Na+,K+ pumps in the exercising muscles, raising doubts as to whether translocation of Na+,K+ pumps had taken place.Virtually all the evidence for translocation of Na+,K+ pumps has been obtained using membrane fractions isolated from muscle homogenates. In contrast, studies on intact muscles or cut muscle specimens in vitro or intact muscles in vivo have consistently failed to detect translocation. Thus, measurements of [3H]ouabain binding to intact rat soleus muscles performed under equilibrium conditions showed no effect of insulin or other conditions known to stimulate the activity of the Na+,K+ pumps (electrical stimulation, epinephrine, insulin-like growth factor I, and amylin; see also Clausen and Hansen, 1977; Dørup and Clausen, 1995; Clausen, 2003; McKenna et al., 2003; Murphy et al., 2008). It cannot be excluded, therefore, that the 70% “unmasking” or “translocation” of Na+,K+ pumps by insulin (Erlij and Grinstein, 1976) reflects the increase in the [3H]ouabain binding taking place before binding equilibrium has been reached (see the section above, [3H]Ouabain binding). This is usually caused by inadequate saturation of the ouabain-binding sites with [3H]ouabain (Clausen and Hansen, 1977). Translocation has often been proposed as the cause of Na+,K+-pump stimulation induced by electrical stimulation or exercise. This hypothesis was tested by comparing effects of electrical stimulation on Na+ extrusion and [3H]ouabain binding in isolated rat soleus. Electrical stimulation in vitro at 120 Hz for 10 s caused a rapid 70% rise in intracellular Na+, followed by a subsequent 18-fold increase in net Na+ extrusion from rat soleus muscle but no significant change in the content of [3H]ouabain-binding sites (McKenna et al., 2003). A later study showed that after 10–60 min of running exercise, there were highly significant increases (18–80%; P < 0.001– 0.01) in the Na+ content of intact rat soleus muscles in vivo but no significant change in [3H]ouabain binding (Murphy et al., 2008). In isolated rat EDL muscles, 15 s of 60-Hz stimulation caused a 7.3-fold increase in net Na+ extrusion (Clausen, 2011). Another study on rat EDL showed that a 10-s stimulation at 60 Hz induced a fivefold increase in Na+,K+-pump activity (Nielsen and Clausen, 1997). It seems unlikely that such pronounced increases in Na+ extrusion mediated by the Na+,K+ pumps could be accounted for by the relatively modest translocation of Na+,K+ pumps or binding of [3H]ouabain (no change detectable) described above.Either in vivo or in vitro administration of salbutamol augments K+ content and reduces Na+ content in rat soleus muscle (Murphy et al., 2008), reflecting increased affinity for intracellular Na+ (Buchanan et al., 2002) probably caused by phosphorylation of FXYD1. Conversely, inhibition of the Na+,K+ pumps with ouabain reduces intracellular K+ and increases intracellular Na+ in vivo (Murphy et al., 2008).Measurement of maximum binding capacity for [3H]ouabain in isolated rat soleus (0.72 nmol/g wet wt) would predict that if all Na+,K+ pumps were operating at full speed, the theoretical maximum K+ uptake should reach 0.72 nmol × 8,900 = 6,408 nmol/g/min at 30°C, which is corrected for temperature using the observed Q10 of 2.3 for the rate of Na+,K+ pumping (Clausen and Kohn, 1977). Are such high values seen in the intact soleus muscle? When isolated rat soleus muscles are loaded with Na+ by preincubation in K+-free KR buffer without Ca2+ or Mg2+, intracellular Na+ reaches 126 mM (Clausen et al., 1987). When these muscles are subsequently incubated for 3 min in KR buffer containing 42K or 86Rb and between 5 and 130 mM K+, the ouabain-suppressible rates of 42K or 86Rb uptake can be measured and the results plotted in an Eadie–Hofstee plot (Fig. 4 in Clausen et al., 1987). The maximum rates of 42K and 86Rb uptake determined from this plot reach 6,150 nmol/g wet wt/min, corresponding to 96% of the above-mentioned theoretical maximum K+ uptake at 30°C. When the same Na+-loading procedure was used, the maximum rate of ouabain-suppressible 86Rb uptake was correlated (r = 0.95; P < 0.001) to the content of [3H]ouabain-binding sites over a wide range of values obtained by varying thyroid status, age, or K+ depletion (Fig. 5 in the present paper). These high 86Rb uptake values were not suppressed by Ba2+ or the local anesthetic tetracaine, indicating that they were mediated by the Na+,K+ pumps and not by K+ channels (Clausen et al., 1987).Open in a separate windowFigure 4.Effects of salbutamol, rat CGRP, dbcAMP, and ouabain on the response of tetanic force to electroporation (EP). Force was measured in rat soleus muscles mounted for isometric contractions in KR buffer at 30°C using 2-s pulse trains of 60 Hz. Trains were repeated three times to obtain an initial determination of tetanic force, and data are presented as a percentage of this value. After transfer to an electroporation cuvette, muscles either underwent electroporation (eight pulses of 500 V/cm and 0.1-ms duration) or were left without electroporation. Force was subsequently recorded at the indicated intervals in buffer without additions (EP controls) or in buffer containing salbutamol (10−5 M), rat CGRP (10−7 M), dbcAMP (1 mM), or salbutamol (10−5 M) plus ouabain (10−3 M). Each point represents the mean of observations on 4–13 muscles, with bars denoting SEM. Reprinted with permission from Acta Physiologica (Clausen and Gissel, 2005).Open in a separate windowFigure 5.Correlation between ouabain-suppressible 86Rb uptake and [3H]ouabain-binding site content in soleus muscles of rats with K+ deficiency and varying age or thyroid status. Ouabain-suppressible 86Rb uptake was measured as described in Clausen et al. (1987). [3H]Ouabain-binding site content was determined on muscle samples using the vanadate-facilitated assay. After correction for nonspecific uptake, [3H]ouabain binding was corrected for radioisotopic purity, incomplete saturation, and loss of specifically bound [3H]ouabain. The line was constructed using linear-regression analysis of unweighted values (r = 0.95; P < 0.001). Reprinted with permission from The Journal of Physiology (Clausen et al., 1987).In conclusion, translocation of Na+,K+ pumps happens in skeletal muscle, but it is difficult to define and quantify. As a mechanism for stimulation of the Na+,K+ pumps it is too slow to explain the documented evidence for the acute Na+,K+-pump activation, which may reach the theoretical maximum in less than 1 min.

Subunit isoforms of the Na+,K+-ATPase

In the mammalian brain, the Na+,K+-ATPase exists in two distinct molecular forms that differ with respect to affinity for ouabain (Sweadner, 1979). Similarly, two subunit isoforms of the enzyme, α1 and α2, were detected in rat skeletal muscle (Tsakiridis et al., 1996). The functional Na+,K+ pump is composed of one of four catalytic α subunits (α1–α4) and one of three structural β subunits (β1–β3). In skeletal muscle, a third subunit, FXYD1, is coexpressed with the α subunits and involved in the regulation of Na+,K+-pump activity (Fig. 3).Open in a separate windowFigure 3.Hypothetical relationship between phospholemman (PLM or FXYD1) and the Na+/K+ATPase α subunit indicating possible effects of PLM phosphorylation on their interaction. The PLM cytoplasmic tail can be phosphorylated at Ser63, Ser68, and Thr69 (Ser69 in mouse in panel a). Phosphorylation at any or all of these sites (or, for example, at the site depicted as undergoing phosphorylation by PKA in panel b) may alter orientation of the cytoplasmic tail, thereby affecting its interaction with the Na+/K+ ATPase α subunit to increase ion transport. Reprinted with permission from Current Opinion in Pharmacology (Shattock, 2009).At present, the subunit isoforms of the Na+,K+-ATPase as detected by immunoblotting cannot be quantified and given in molar units for the contents per gram muscle wet weight (Tsakiridis et al., 1996; Clausen, 2003). A recent study (McKenna et al., 2012) indicates that in human vastus lateralis muscle, there is a relative decrease of −24% in the abundance of the α2 subunit from 24 to 67 yr of age (P < 0.05). This decrease was associated with a 36% reduction in muscle strength and peak O2 consumption. However, the same study showed no significant age-dependent change in the content of [3H]ouabain-binding sites (Bangsbo et al., 2009). More recent studies on human vastus lateralis show that 10 d of training increases the abundance of α1 and α2 subunit by 113 and 49%, respectively (Benziane et al., 2011). In rat hind-limb muscle, 60 min of treadmill running (20 m/min) induces a 55–64% increase in the abundance of the Na+,K+-ATPase α1 subunit and a 43–94% increase in that of the α2 subunit (Tsakiridis et al., 1996).In the isolated rat EDL muscle, the mechanisms for the excitation-induced up-regulation of mRNA for the α1-3 subunits were examined. Three bouts of electrical stimulation (60 Hz for10 s) had no immediate effect on the abundance of the mRNA-encoding α1–α3-subunit isoforms (Murphy et al., 2006). However, after 3 h of poststimulatory rest, there were 223, 621, and 892% increases in the abundance of the mRNA-encoding subunits α1, α2, and α3, respectively. However, in rat EDL muscle, even massive increases (598–769%) in intracellular Na+ induced by pretreatment with ouabain, veratridine, or monensin produced no significant change in the mRNA encoding any of the subunit isoforms. Increasing intracellular Ca2+ by preincubation with the Ca2+ ionophore A23187 produced no increase in the mRNA for the two major subunit isoforms (α1 and α2). Thus, a combination of multiple signals seems to be required to recapitulate the increase in abundance of the mRNAs that trigger the increase in synthesis of Na+,K+ pumps induced by exercise. In conclusion, the abundance of the α1 and α2 subunits seems to respond to training and age, but the observed relative changes show considerable variation and cannot yet be “translated” to molar units for Na+,K+ pumps per gram muscle tissue. As shown in Radzyukevich et al., 2013). These defects could be reproduced in isolated muscles from control animals by selectively inhibiting the α2 subunit with 5 µM ouabain.

Molecular mechanisms of pump regulation. Acute regulation of Na+,K+-ATPase by hormones.

Insulin has a hypokalemic action, prompting its use in the therapy of hyperkalemia (Harrop and Benedict, 1924) and inspiring studies with isolated rat muscles that showed that insulin increases the uptake of K+ and the extrusion of Na+ (Creese, 1968), leading to reduced intracellular Na+ and hyperpolarization (Zierler and Rabinowitz, 1964). Because the effects of insulin on Na+,K+ fluxes and membrane potential are suppressed by ouabain, they are likely caused by stimulation of the Na+,K+ pumps (Clausen and Kohn, 1977, Clausen, 2003). The mechanism seems to reflect increased affinity for Na+ on the inner surface of the Na+,K+ pumps (Clausen, 2003). As discussed in the section above on translocation (Translocation of Na+,K+ pumps compared…), the stimulating action of insulin on the Na+,K+ pumps may not be attributed to translocation of Na+,K+ pumps. Studies on cultured rat muscle cells indicate that the stimulatory effect of insulin on 86Rb uptake is abolished by ouabain and reduced by phorbol ester, indicating that it is mediated by Na+,K+ pumps and caused by activation of PKC (Sampson et al., 1994).Similarly, the hypokalemic action of catecholamines inspired studies with isolated muscles demonstrating stimulatory effects of epinephrine and norepinephrine on the uptake of K+ and the net extrusion of Na+ (Clausen and Flatman, 1977). Because of the 3:2 exchange of Na+ versus K+, the Na+,K+ pump is electrogenic and its stimulation leads to hyperpolarization. Detailed studies showed that this stimulation depended on a β2 adrenoceptor–mediated activation of adenylate cyclase increasing the cytoplasmic concentration of cAMP (Clausen and Flatman, 1977; Clausen, 2003). cAMP, via PKA, induces phosphorylation of FXYD1serine68, causing increased Na+ affinity of the Na+,K+ pumps (Shattock, 2009) and ensuing reduction of intracellular Na+. A similar sequence of events explain the cAMP-mediated, Na+,K+ pump–stimulatory action of other hormones and pharmaceuticals (including calcitonin, CGRP, amylin, isoprenaline, and theophylline; see Clausen, 1998, and Fig. 6 in Clausen, 2003). More detailed molecular insight is discussed by Shattock (2009) and in this paper.

Na+,K+-ATPase regulation by auxiliary proteins.

The Na+,K+ pumps are subject to regulation by a series of proteins. The β subunit (Fig. 1) is a glycoprotein required for the transfer of the entire Na,K-ATPase enzyme molecule from its site of synthesis in the endoplasmic reticulum to its site of insertion in the plasma membrane. The γ subunit, originally identified in the kidney, augments the affinity of the myocardial Na+,K+-ATPase for Na+ (Bibert et al., 2008). In mammals, there is a family of at least seven regulatory proteins termed FXYD (“fix-it”) that are associated with the Na+,K+-ATPase in a tissue-specific way (Mahmmoud et al., 2000; Cornelius and Mahmmoud, 2003). The first FXYD identified, called FXYD1 and also known as phospholemman, was originally found in myocardial cells (Palmer et al., 1991). FXYD1 is a major substrate for PKA and PKC and was later found to regulate the Na+,K+-ATPase in heart and skeletal muscle (Shattock, 2009). Increased cAMP activates cAMP-dependent kinase (PKA), which, in turn, phosphorylates serine-68 on FXYD1 (Fig. 3). This disinhibits the Na+,K+ pumps and stimulates their activity by raising Vmax and the sensitivity of the pumps to intracellular Na+. Phosphorylation of FXYD increases the maximal Na+,K+-pump current of α2/β isozymes (Bibert et al., 2008). cAMP seems to function as a common signaling molecule acting via PKA to augment the affinity of the Na+,K+ pump for [Na+]i in muscle (Shattock, 2009). In addition to Ser68, FXYD can also be phosphorylated at Ser63 (by PKC) and at Thr69 (by an unidentified kinase).In the human vastus lateralis muscle, 30 s of intense exercise clearly increased phosphorylation of FXYD (Thomassen et al., 2011). Another study demonstrated that 60 min of acute exercise increases phosphorylation of FXYD from human vastus lateralis by 75% (Benziane et al., 2011). In soleus muscle of wild-type mice, 10 min of electrical stimulation caused a 59% increase in phosphorylation of FXYD. This phosphorylation was not seen in PKCα knockout mice and therefore seems to depend on PKC (Thomassen et al., 2011).

Physiology and pathophysiology of the Na+,K+-ATPase in skeletal muscle

The electrogenic action of the Na+,K+ pumps.

Because each Na+,K+-pump molecule in the plasma membrane exchanges three Na+ ions for two K+ ions in each cycle, each cycle leads to a net extrusion of one positive charge. Measurements on isolated rat soleus muscle show that the resting ouabain-suppressible efflux of 22Na at 30°C is 0.287 µmol/g wet wt/min, and the concomitant influx of 42K is 0.196 µmol/g wet wt/min (Clausen and Kohn, 1977). Thus, the ratio between Na+,K+ pump–mediated Na+ efflux and K+ influx is 0.287/0.196 = 1.46, close to the 3:2 ratio. This electrogenic action of the Na+,K+ pumps is consistent with the observation that, in rat soleus muscle, blocking the Na+,K+ pumps with ouabain causes a depolarization of ∼10 mV in 10 min (Clausen and Flatman, 1977). Conversely, acute stimulation of the rate of active Na+,K+ transport by epinephrine in isolated rat soleus muscle causes a hyperpolarization of ∼8 mV, which is abolished by ouabain. The β2 agonist salbutamol induces a similar effect in vitro, and when given intravenously, it causes 15-mV hyperpolarization in rat soleus (Clausen and Flatman, 1980). The following agents, which stimulate active Na+,K+ transport, all induce significant hyperpolarization in rat, mouse, or guinea pig soleus: epinephrine, norepinephrine, isoprenaline, salbutamol, dbcAMP, phenylephrine, rat, human or salmon calcitonin, CGRP, and insulin (Clausen and Flatman, 1977, 1987; Andersen and Clausen, 1993). Rat soleus muscles can be loaded with Na+ by a 90-min preincubation in K+-free KR. When subsequently transferred to KR with normal K+, the rate of 22Na efflux is more than doubled. This effect is blocked by ouabain and associated with an ∼10-mV hyperpolarization (Clausen and Flatman, 1987). These observations indicate that the electrogenic action of hormonal Na+,K+-pump stimulation can be mimicked by Na+ loading.

The Na+,K+ pumps compensate functional defects caused by plasma membrane leakage

Apart from the channel-mediated selective Na+ and K+ leaks that take place during excitation, the plasma membrane may develop nonspecific leaks caused by loss of integrity during intense work, anoxia, physical damage, electroporation, excessive swelling, rhabdomyolysis, or various muscle diseases. The response to such leaks and the possible role of the Na+,K+ pumps in compensating the ensuing functional disorders have been examined using electroporation of isolated rat soleus and EDL muscles (Clausen and Gissel, 2005). In these studies, muscles were mounted in an electroporation cuvette and exposed to short-lasting (0.1-ms) pulses of an electrical field of 100–800 V/cm across the muscles. This induces rapid formation of pores in the plasma membrane, increasing its permeability, but not in the membranes of intracellular organelles. This allows reversible influx of Na+, loss of K+ and excitability, release of intracellular proteins, and penetration of extracellular markers into the cytoplasm (Clausen and Gissel, 2005). As shown in Fig. 4, eight electroporation pulses of 500 V/cm induces immediate complete loss of tetanic force, which in 200 min is followed by spontaneous recovery to around 30% of the force measured before electroporation. The initial phase of this recovery is considerably (183–433%) enhanced by stimulating the Na+,K+ pumps with salbutamol, rat CGRP, or dibutyryl cAMP. Both the spontaneous 30% force recovery and the 50% force recovery induced by salbutamol were abolished by ouabain, indicating that they were caused by Na+,K+-pump stimulation. The electroporation caused a depolarization from −70 to around −20 mV, followed by a partial spontaneous recovery. Salbutamol (10−5 M) further improved the repolarization by 15 mV (Clausen and Gissel, 2005). In isolated rat EDL muscles, 30–60 min of intermittent stimulation caused a drop in tetanic force to 12% of the initial level, followed by slow spontaneous recovery to 20–25% of the initial force level. This was associated with 11–15-mV depolarization and marked loss of the intracellular protein lactic acid dehydrogenase. Subsequent stimulation of the Na+,K+ pumps with salbutamol restored membrane potential to normal level. Salbutamol, epinephrine, rat CGRP, and dibutyryl cAMP all induced a significant increase (40–90%) in the force recovery after intermittent stimulation (Mikkelsen et al., 2006). In EDL muscle, anoxia caused massive reductions in the content of phosphocreatine and ATP and a marked loss of force, which was clearly reduced by the two β2 agonists salbutamol and terbutaline. The force recovery seen after reoxygenation was markedly improved by salbutamol, the long-acting β2 agonist salmeterol, theophylline (a phosphodiesterase inhibitor that inhibits the breakdown of cAMP), and dibutyryl cAMP (causing the following relative increases in force: 55–262%). The salbutamol-induced increase in force recovery could be related to partial restoration of excitability. In anoxic muscles, salbutamol had no effect on phosphocreatine or ATP but decreased intracellular Na+ concentration and increased 86Rb uptake and K+ content, indicating that the mechanism is cAMP-mediated stimulation of the electrogenic Na+,K+ pumps (Fredsted et al., 2012) and cannot be related to recovery of energy status. Thus, the Na+,K+ pumps may restore excitability even when energy status is low. This interpretation is in keeping with a detailed analysis in isolated mouse soleus and EDL muscles, suggesting that impaired excitability is the main contributor to severe fatigue, discounting anoxia as the major contributor to fatigue in isolated muscles (Cairns et al., 2009).

Functional significance of Na+,K+ pumps.

The evidence described above indicates that the content of Na+,K+ pumps measured using [3H]ouabain binding represents Na+,K+ pumps, which may all become operational when activated by intracellular Na+ loading and increased extracellular K+, by hormones or excitation. Is it possible to induce a rapid activation of all the Na+,K+ pumps in the intact muscle? In rat soleus muscle, electrical stimulation at 120 Hz for 10 s increases intracellular Na+ to ∼50 mM (Nielsen and Clausen, 1997). When the muscles were subsequently allowed to rest in standard KR at 30°C, the net efflux of Na+ recorded over the first 30 s was shown to reach 9,000 nmol/g wet wt/min, 97% of the theoretical maximum Na+ transport rate of 9,300 nmol/g wet wt/min (Fig. 5 in Nielsen and Clausen, 1997). This indicates that during intense exercise, most likely to produce a high degree of Na+,K+-pump use, a 97% utilization has been documented. Apparently, it requires the combination of large increases in intracellular Na+ and extracellular K+, which, as described above, can be mimicked by Na+ loading and subsequent incubation at high [K+]o (Clausen et al., 1987). In rat soleus, 60-Hz stimulation induces a total efflux of K+ of 5,580 nmol/g wet wt/min (Clausen et al., 2004). Only full activation of all Na+,K+ pumps would enable K+ reuptake sufficient to allow extracellular K+ clearance to keep pace with excitation-induced loss of K+. As already noted, the theoretical maximum Na+,K+ pump–mediated K+ uptake in rat soleus should reach 6,408 nmol K+/g wet wt/min, slightly exceeding the above-mentioned total loss of K+. Such large increases in active Na+,K+ transport would appear energetically unlikely. However, as stated previously, both in resting and working muscles, the energy requirement of the Na+,K+ pumps only amounts to 2–10% of the total energy turnover. This evaluation provides an example of the importance of accurate quantification of the Na+,K+ pumps and their energy requirements.This information raises the question of whether such combinations of high [Na+]i and [K+]o occur in working intact muscle in vitro or in vivo. Several studies indicate that isolated muscles and muscle fibers, when stimulated at resting length, take up substantial amounts of Na+ and release nearly equimolar amounts of K+ (Hodgkin and Horowicz, 1959; Clausen, 2008b, 2011, 2013; Clausen et al., 2004). In vitro experiments show that when the amount of K+ released from the cells (µmoles/gram wet weight as measured by flame photometry) is divided by the extracellular space (milliliter/gram wet weight), it can be calculated that in isolated soleus and EDL muscles, [K+]o reaches levels of 20 or 50 mM, respectively, sufficient to suppress excitability (Clausen, 2008a). More detailed studies showed that in isolated EDL muscles, 60 s of stimulation at 20 Hz increases [K+]o to around 45 mM, leading to loss of excitability (Clausen, 2011). Recent in vitro and in vivo studies show that during direct or indirect electrical stimulation (at 5 Hz for 300 s or 60 Hz for 60 s) of rat EDL muscles, the mean extracellular K+ in the stimulated muscles may reach up to 50–70 mM (Clausen, 2013). This was combined with an appreciable decrease in [Na+]o (46 mM), causing further loss of force. This provides new evidence that in the intact organism, excitation-induced rise in [K+]o and decrease in [Na+]o can be major causes of muscle fatigue. Moreover, as described above, the Na+,K+ pumps available in the muscles may fully use their maximum transport capacity to counterbalance the excitation-induced rise in [K+]o. An important function of the Na+,K+ pumps is to protect muscle excitability against increases in [K+]o. When exposed to 10–15 mM K+, contractility of rat soleus or EDL muscles is reduced or lost. Force may rapidly be restored by stimulating the Na+,K+ pumps with salbutamol or insulin (Clausen, 2003; Clausen and Nielsen, 2007).The excitation-induced rise in [K+]o depends on muscle fiber type (slow-twitch or fast-twitch fibers) and the frequency of contractions. In isolated rat EDL muscle (predominantly fast-twitch fibers), the excitation-induced influx of Na+ and release of K+ per action potential is, respectively, 6.5- and 6.6-fold larger than in soleus (predominantly slow-twitch fibers) (Clausen et al., 2004). When stimulated continuously at 60 Hz, the rate of force decline over the first 20 s is 5.9-fold faster in EDL than in soleus muscle. When tested using continuous stimulation in the frequency range of 10 to 200 Hz, the rates of force decline in rat soleus and EDL muscles are closely correlated (r2 = 0.93; P < 0.002 and 0.99; P < 0.01, respectively) to the excitation-induced measured increase in [K+]o (Clausen, 2008b). These observations indicate that the faster rate of force decline and fatigability seen in fast-twitch muscles as compared with slow-twitch muscles is caused by the much larger excitation-induced passive Na+,K+ fluxes. In view of this difference, leading to a larger work-induced increase in intracellular Na+, it would be expected that the content of Na+,K+ pumps in EDL would undergo long-term up-regulation to a level considerably over and above that seen in soleus. However, measurements of [3H]ouabain-binding sites in vitro and in vivo show that rat EDL only contains 20–30% more Na+,K+ pumps per gram wet weight than soleus (Clausen et al., 1982; Murphy et al., 2008).Control soleus muscle, ouabain-pretreated soleus, or soleus from K+-deficient rats show a graded reduction in functional Na+,K+ pumps from 756 pmol/g wet wt down to 110 pmol/g wet wt. The content of [3H]ouabain-binding sites is significantly correlated (r2 = 0.88; P < 0.013) to the rate of force decline developing over 30 s of continuous stimulation at 60 Hz (Nielsen and Clausen, 1996). This illustrates the crucial role of the Na+,K+-pump capacity in maintaining contractile endurance. In keeping with this, K+ deficiency in human subjects, which reduces the content of Na+,K+ pumps, causes reduced grip strength and fatigue (Clausen, 1998, 2003). Conversely, 19 studies have documented that exercise training leads to an increase in the content of Na+,K+ pumps measured as [3H]ouabain binding (Clausen, 2003). This up-regulation improved the clearance of plasma K+ during exercise, a mechanism for reducing fatigue (McKenna et al., 1993). During contractile activity, the K+ released from the muscle cells is normally cleared via the circulation. When the blood vessels are compressed because of static contractions when carrying heavy burdens (Barcroft and Millen, 1939), this clearance is reduced or abolished. Therefore, extracellular K+ may undergo further increase, and its clearance will depend on the transport capacity and activity of the Na+,K+ pumps. Through a similar mechanism, deficient circulation may lead to reduction of contractile force.

Conclusions

Recent studies indicate that in skeletal muscle, excitation-induced net gain of Na+ and loss of K+ are considerably larger than reported previously and represent important causes of fatigue. This implies that to counterbalance these rapid passive Na+,K+ fluxes, working muscles have to make full use of the transport capacity of their Na+,K+ pumps. This capacity can be quantified using [3H]ouabain binding and expressed in picomoles of Na+,K+ pumps per gram muscle wet weight, offering values that can be confirmed by measurements of maximum ouabain-suppressible fluxes of Na+ and K+. This allows quantification of the capacity to clear extracellular K+ in patients with reduced content of Na+,K+ pumps in their muscles. Numerous studies have shown that the content of Na+,K+ pumps in skeletal muscles can be quantified, documenting the up-regulation of the content of Na+,K+ pumps induced by exercise and hormones as well as the down-regulation seen in various diseases associated with reduced physical performance or fatigue. Assays based on immunoblotting provide relative changes in abundance that cannot yet be translated into molar units. The activity of the Na+,K+ pumps may increase rapidly (by electrical stimulation up to 20-fold in 10 s) or by Na+ loading, insulin, catecholamines, calcitonins, amylin, and theophylline.  相似文献   

12.
Using the antibiotic Nystatin, we have developed a systematic method for the preparation of red blood cells with independently selected levels of intracellular Na+ concentrations and water content. Such cells provided an experimental model to study the effect of Na+/K+ pump stimulation on red cell water content. Even in initially dehydrated cells, stimulation of the Na+/K+ pump by elevated intracellular Na+ caused subsequent further loss of cell water. Cell water loss was reflected in decreased monovalent cation content per unit mass of hemoglobin and by a shift in the density distribution of the cell populations to higher densities on discontinuous Stractan gradients. We conclude that the 3 Naout+ : 2 Kin+ stoichiometry of the Na+/K+ pump results in a net desalting effect with increased pump activity. Under the conditions of these experiments, the cell appears to have no effective mechanism to compensate for a net loss of ions and water.  相似文献   

13.
To examine the extracellular Na+ sensitivity of a renal inwardly rectifying K+ channel, we performed electrophysiological experiments on Xenopus oocytes or a human kidney cell line, HEK293, in which we had expressed the cloned renal K+ channel, ROMK1 (Kir1.1). When extracellular Na+ was removed, the whole-cell ROMK1 currents were markedly suppressed in both the oocytes and HEK293 cells. Single-channel ROMK1 activities recorded in the cell-attached patch on the oocyte were not affected by removal of Na+ from the pipette solution. However, macro-patch ROMK1 currents recorded on the oocyte were significantly suppressed by Na+ removal from the bath solution. A blocker of Na+/H+ antiporters, amiloride, largely inhibited the Na+ removal-induced suppression of whole-cell ROMK1 currents in the oocytes. The pH-insensitive K80M mutant of ROMK1 was much less sensitive to Na+ removal. Na+ removal was found to induce a significant decrease in intracellular pH in the oocytes using H+-selective microelectrodes. Coexpression of ROMK1 with NHE3, which is a Na+/H+ antiporter isoform of the kidney apical membrane, conferred increased sensitivity of ROMK1 channels to extracellular Na+ in both the oocytes and HEK293 cells. Thus, it is concluded that the ROMK1 channel is regulated indirectly by extracellular Na+, and that the interaction between NHE transporter and ROMK1 channel appears to be involved in the mechanism of Na+ sensitivity of ROMK1 channel via regulating intracellular pH. Received: 13 April 1999/Revised: 15 July 1999  相似文献   

14.
Previous studies in expression systems have found different ion activation of the Na+/K+-ATPase isozymes, which suggest that different muscles have different ion affinities. The rate of ATP hydrolysis was used to quantify Na+,K+-ATPase activity, and the Na+ affinity of Na+,K+-ATPase was studied in total membranes from rat muscle and purified membranes from muscle with different fiber types. The Na+ affinity was higher (K m lower) in oxidative muscle compared with glycolytic muscle and in purified membranes from oxidative muscle compared with glycolytic muscle. Na+,K+-ATPase isoform analysis implied that heterodimers containing the β1 isoform have a higher Na+ affinity than heterodimers containing the β2 isoform. Immunoprecipitation experiments demonstrated that dimers with α1 are responsible for approximately 36% of the total Na,K-ATPase activity. Selective inhibition of the α2 isoform with ouabain suggested that heterodimers containing the α1 isoform have a higher Na+ affinity than heterodimers containing the α2 isoform. The estimated K m values for Na+ are 4.0, 5.5, 7.5 and 13 mM for α1β1, α2β1, α1β2 and α2β2, respectively. The affinity differences and isoform distributions imply that the degree of activation of Na+,K+-ATPase at physiological Na+ concentrations differs between muscles (oxidative and glycolytic) and between subcellular membrane domains with different isoform compositions. These differences may have consequences for ion balance across the muscle membrane.  相似文献   

15.
[3H]-Ouabain binding to muscle preparations was utilized to estimate the number of Na+,K+-ATPase enzyme units in hindlimbs from 8 week old lean and obese mice. Specific [3H]-ouabain binding per mg particulate protein was 36% lower in obese mice; whereas, the affinity of the binding sites for ouabain was similar in obese and lean mice. Since obese mice had less muscle than lean mice, the number of Na+,K+-ATPase enzyme units in hindlimbs from obese mice was less than half the number observed in lean mice.  相似文献   

16.
Effects of long-term, subtotal inhibition of Na+-K+ transport, either by growth of cells in sublethal concentrations of ouabain or in low-K+ medium, are described for HeLa cells. After prolonged growth in 2 × 10?8 M ouabain, the total number of ouabain molecules bound per cell increases by as much as a factor of three, mostly due to internalization of the drug. There is only about a 20% increase in ouabain-binding sites on the plasma membrane, representing amodest induction of Na+, K+-ATPase. In contrast, after long-term growth in low K+ there can be a twofold or greater increase in ouabain binding per cell, and in this case the additional sites are located in the plasma membrane. The increase is reversible. To assess the corresponding transport changes, we have separately estimated the contributions of increased intracellular [Na+] and of transport capacity (number of transport sites) to transport regulation. During both induction and reversal, short-term regulation is achieved primarily by changes in [Na+]i. More slowly, long-term regulation is achieved by changes in the number of functional transporters in the plasma membrane as assessed by ouabain binding, Vmax for transport, and specific phosphorylation. Parallel exposure of cryptic Na+, K+-ATPase activity with sodium dodecyl sulfate in the plasma membranes of both induced and control cells showed that the induction cannot be accounted for by an exposure of preexisting Na+, K+-ATPase in the plasma membrane. Analysis of the kinetics of reversal indicates that it may be due to a post-translational event.  相似文献   

17.
Lettré cells maintain a plasma membrane potential near — 60mV, yet are scarcely depolarized by 80 mM Rb+ and are relatively impermeable to 86Rb+. They are depolarized by ouabain without a concomitant change in intracellular cation content. Addition of K+ to cells suspended in a K+ free medium, or of Na+ to cells in a Na+ free medium, hyperpolarizes the cells. They contain electroneutral transport mechanisms for Na+, K+ and H+ which can function as Na+:K+ and Na+:H+ exchanges. It is concluded that plasma membrane potential of Lettré cells, in steady-state for Na+ and K+, is produced by an electrogenic Na+ pump sustained by electroneutral exchanges, and restricted by anion leakage.  相似文献   

18.
The Na+-K+ pumps in the transverse tubular (T) system of a muscle fiber play a vital role keeping K+ concentration in the T-system sufficiently low during activity to prevent chronic depolarization and consequent loss of excitability. These Na+-K+ pumps are located in the triad junction, the key transduction zone controlling excitation-contraction (EC) coupling, a region rich in glycolytic enzymes and likely having high localized ATP usage and limited substrate diffusion. This study examined whether Na+-K+ pump function is dependent on ATP derived via the glycolytic pathway locally within the triad region. Single fibers from rat fast-twitch muscle were mechanically skinned, sealing off the T-system but retaining normal EC coupling. Intracellular composition was set by the bathing solution and action potentials (APs) triggered in the T-system, eliciting intracellular Ca2+ release and twitch and tetanic force responses. Conditions were selected such that increased Na+-K+ pump function could be detected from the consequent increase in T-system polarization and resultant faster rate of AP repriming. Na+-K+ pump function was not adequately supported by maintaining cytoplasmic ATP concentration at its normal resting level (8 mM), even with 10 or 40 mM creatine phosphate present. Addition of as little as 1 mM phospho(enol)pyruvate resulted in a marked increase in Na+-K+ pump function, supported by endogenous pyruvate kinase bound within the triad. These results demonstrate that the triad junction is a highly restricted microenvironment, where glycolytic resynthesis of ATP is critical to meet the high demand of the Na+-K+ pump and maintain muscle excitability. muscle fatigue; sodium-potassium-adenosinetriphosphatase; excitation-contraction coupling; T-system; excitability  相似文献   

19.
Summary Rabbit erythrocytes are well known for possessing highly active Na+/Na+ and Na+/H+ countertransport systems. Since these two transport systems share many similar properties, the possibility exists that they represent different transport modes of a single transport molecule. Therefore, we evaluated this hypothesis by measuring Na+ transport through these exchangers in acid-loaded cells. In addition, selective inhibitors of these transport systems such as ethylisopropyl-amiloride (EIPA) and N-ethylmaleimide (NEM) were used. Na+/Na+ exchange activity, determined as the Na o + -dependent22Na efflux or Na i + -induced22Na entry was completely abolished by NEM. This inhibitor, however, did not affect the H i + -induced Na+ entry sensitive to amiloride (Na+/H+ exchange activity). Similarly, EIPA, a strong inhibitor of the Na+/H+ exchanger, did not inhibit Na+/Na countertransport, suggesting the independent nature of both transport systems. The possibility that the NEM-sensitive Na+/Na+ exchanger could be involved in Na+/H+ countertransport was suggested by studies in which the net Na+ transport sensitive to NEM was determined. As expected, net Na+ transport through this transport system was zero at different [Na+] i /[Na+] o ratios when intracellular pH was 7.2. However, at pH i =6.1, net Na+ influx occurred when [Na+] i was lower than 39mm. Valinomycin, which at low [K+] o was lower than 39mm. Valinomycin, which at low [K+] o clamps the membrane potential close to the K+ equilibrium potential, did not affect the net NEM-sensitive Na+ entry but markedly stimulated, the EIPA-and NEM-resistant Na+ uptake. This suggest that the net Na+ entry through the NEM-sensitive pathway at low pH i , is mediated by an electroneutral process possibly involving Na+/H+ exchange. In contrast, the EIPA-sensitive Na+/H+ exchanger is not involved in Na+/Na+ countertransport, because Na+ transport through this mechanism is not affected by an increase in cell Na from 0.4 to 39mm. Altogether, these findings indicate that both transport systems: the Na+/Na+ and Na+/H+ exchangers, are mediated by distinct transport proteins.  相似文献   

20.
Two of five Zygosaccharomyces rouxii mutants defective in salt tolerance, 152S (sat1) and 1717S (SAT3), were inviable in a nutrient medium (YPD) containing more than 1% NaCl. These two mutant cells contained significantly higher amounts of Na+ (298 μmol and 285 μmol per g cells of 152S and 1717S, respectively) but lower amounts of K+ (242 μmol and 176 μmol per g cells of 152S and 1717S, respectively) than three other mutants, 41S (sat2-1 [98 μmol Na+ and 326 μmol K+/g cells]), 197S (sat2-2 [103μmol Na+ and 336 μmol K+/g cells]), 1611S (SAT4 [139 μmol Na+ and 294 μmol K+/g cells]), as well as a wild-type strain, AN39 (61 μmol Na+ and 349 μmol K+/g cells), when cultured in YPD medium containing 0.8% NaCl. A KCl supplement, optimally 0.6 M, added to the medium somewhat restored the NaCl-hypersensitivity of 152S and 1717S with a concomitant decrease of intracellular Na+. This finding suggests that the NaCl-hypersensitive mutations are due to a defect in the Na+-regulating mechanism. The other three mutants showed weak responses to KCl in high NaCl-YPD. These five salt sensitive mutants and the wild-type strain retained the same levels of intracellular glycerol and arabitol when transferred into NaCl (5%)-YPD from YDP medium. This suggests that polyol accumulation is not the only mechanism of salt tolerance in Z. rouxii.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号