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 500 毫秒
1.
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   

2.
3.
The catalytic α-subunits of Na,K- and H,K-ATPase require an accessory β-subunit for proper folding, maturation, and plasma membrane delivery but also for cation transport. To investigate the functional significance of the β-N terminus of the gastric H,K-ATPase in vivo, several N-terminally truncated β-variants were expressed in Xenopus oocytes, together with the S806C α-subunit variant. Upon labeling with the reporter fluorophore tetramethylrho da mine-6-maleimide, this construct can be used to determine the voltage-dependent distribution between E1P/E2P states. Whereas the E1P/E2P conformational equilibrium was unaffected for the shorter N-terminal deletions βΔ4 and βΔ8, we observed significant shifts toward E1P for the two larger deletions βΔ13 and βΔ29. Moreover, the reduced ΔF/F ratios of βΔ13 and βΔ29 indicated an increased reverse reaction via E2P → E1P + ADP → E1 + ATP, because cell surface expression was completely unaffected. This interpretation is supported by the reduced sensitivity of the mutants toward the E2P-specific inhibitor SCH28080, which becomes especially apparent at high concentrations (100 μm). Despite unaltered apparent Rb+ affinities, the maximal Rb+ uptake of these mutants was also significantly lowered. Considering the two putative interaction sites between the β-N terminus and α-subunit revealed by the recent cryo-EM structure, the N-terminal tail of the H,K-ATPase β-subunit may stabilize the pump in the E2P conformation, thereby increasing the efficiency of proton release against the million-fold proton gradient of the stomach lumen. Finally, we demonstrate that a similar truncation of the β-N terminus of the closely related Na,K-ATPase does not affect the E1P/E2P distribution or pump activity, indicating that the E2P-stabilizing effect by the β-N terminus is apparently a unique property of the H,K-ATPase.The gastric H,K-ATPase fulfills the remarkable task of pumping protons against a more than 106-fold concentration gradient. H+ extrusion is coupled to countertransport of an equal number of K+ ions for each ATP molecule hydrolyzed, resulting in an electroneutral overall process (1). Characteristic for all P-type ATPases, the enzyme cycles between the two principal conformational states (E1 and E2) and the corresponding phosphointermediates (E1P and E2P), which are formed by reversible phosphorylation of an aspartate residue in the highly conserved DKTGTLT motif. According to a Post-Albers-like reaction scheme (see Fig. 1A), the conformational E1P → E2P transition converts the high H+/low K+ affinity of the cation binding pocket into a low H+/high K+ affinity binding site, hence enabling proton release into the stomach lumen and subsequent binding of extracellular K+. Because the pump faces a lumenal proton concentration of ∼150 mm (2), proton release is probably the energetically most demanding step in the reaction cycle. Thus, during the conformational E1P → E2P transition, enormous pKa changes of the H+-coordinating residues have to occur that most likely involve the rearrangement of a positively charged lysine side chain (Lys-791 in rat H,K-ATPase) (3).Open in a separate windowFIGURE 1.Post-Albers scheme (A) and cryo-EM structural representation of pig gastric H,K-ATPase in the fluoroaluminate-bound pseudo-E2P state (B). A, Post-Albers scheme of the proposed reaction cycle of the gastric H,K-ATPase. E1P/E2P conformational states giving rise to voltage jump-induced fluorescence changes of TMRM-labeled H,K-ATPase molecules are highlighted (gray box). B, structural representation based on the cryo-EM structure of the pig gastric H,K-ATPase (surface or mesh, contoured at 1 σ; EM Data Bank code 5104) and the corresponding homology model (schematic; Protein Data Bank code 3IXZ). Inset, a close-up view (from the right side of the molecule) showing the putative interaction sites of the β-subunit N terminus with the P-domain (red arrow) and αTM3 (black arrow), respectively. Color coding is indicated in the figure.All P2-type ATPases share a common catalytic α-subunit, composed of 10 transmembrane domains harboring the ion-binding sites and a large cytoplasmic loop with the nucleotide-binding domain, the phosphorylation domain (P-domain),2 and the actuator domain (A-domain) (4). However, a unique feature of K+-transporting Na,K- and H,K-ATPase enzymes is the requirement for an accessory β-subunit, which is indispensable for proper folding, maturation, and plasma membrane delivery (5, 6). Despite only 20–30% overall sequence identity between the H,K-ATPase β-subunit and the Na,K β-isoforms, the topogenic structure is similar: a short N-terminal cytoplasmic tail, followed by a single transmembrane segment and a large extracellular C-terminal domain with glycosylation sites and disulfide-bridging cysteines. Numerous studies have demonstrated that the β-subunit of the Na,K-ATPase is more than just a chaperone for the α-subunit, being also required for proper ion transport activity of the holoenzyme. In fact, it has been discovered that different cell- and tissue-specific β-isoforms have distinct effects on the cation affinities (79). Furthermore, it was shown that mutational changes in all three topogenic domains of the Na,K-ATPase β-subunit (1019) as well as chemical interference with disulfide-forming cysteines in the Na,K-ATPase β-subunit ectodomain (2022) affect the cation transport properties of the sodium pump. Finally, conformational changes in the β-subunit during the Na,K-ATPase reaction cycle were demonstrated by proteolytic digestion studies (23) and voltage clamp fluorometry (24).Far less is known about the functional significance of the single H,K-ATPase β-isoform, especially about its potential impact on cation transport (reviewed in Refs. 25 and 26). We have proven recently that E2P state-specific transmembrane interactions between residues in αTM7 and two highly conserved tyrosines in the βTM of both Na,K- and H,K-ATPase significantly stabilize the E2P conformation (19). Mutational disruptions of this interaction resulted in substantial shifts toward E1P and severely affected H+ secretion, which highlighted the physiological relevance of this E2P state stabilization. Notably, according to the recently published cryo-EM structure of pig gastric H,K-ATPase in the pseudo-E2P state (27), the N-terminal tail of the β-subunit makes direct contact with the phosphorylation domain of the α-subunit (see Fig. 1B), thus indicating an additional E2P state stabilization mediated by the β-N terminus. Although this idea was further supported by biochemical studies on N-terminally truncated mutants, direct evidence for this putative E2P-stabilizing interaction and its potential significance for ion transport in intact cells is still lacking.Here, we demonstrate for the first time the functional importance of the gastric H,K-ATPase β-subunit N terminus in living cells under in vivo conditions: voltage clamp fluorometry, Rb+ flux, and SCH28080 sensitivity measurements revealed E1P-shifted, ion transport-impaired phenotypes for two N-terminally truncated H,K β-variants, thus substantiating the E2P-stabilizing effect of the β-N terminus suggested by the recent cryo-EM structure.  相似文献   

4.
The neurological disorders familial hemiplegic migraine type 2 (FHM2), alternating hemiplegia of childhood (AHC), and rapid-onset dystonia parkinsonism (RDP) are caused by mutations of Na+,K+-ATPase α2 and α3 isoforms, expressed in glial and neuronal cells, respectively. Although these disorders are distinct, they overlap in phenotypical presentation. Two Na+,K+-ATPase mutations, extending the C terminus by either 28 residues (“+28” mutation) or an extra tyrosine (“+Y”), are associated with FHM2 and RDP, respectively. We describe here functional consequences of these and other neurological disease mutations as well as an extension of the C terminus only by a single alanine. The dependence of the mutational effects on the specific α isoform in which the mutation is introduced was furthermore studied. At the cellular level we have characterized the C-terminal extension mutants and other mutants, addressing the question to what extent they cause a change of the intracellular Na+ and K+ concentrations ([Na+]i and [K+]i) in COS cells. C-terminal extension mutants generally showed dramatically reduced Na+ affinity without disturbance of K+ binding, as did other RDP mutants. No phosphorylation from ATP was observed for the +28 mutation of α2 despite a high expression level. A significant rise of [Na+]i and reduction of [K+]i was detected in cells expressing mutants with reduced Na+ affinity and did not require a concomitant reduction of the maximal catalytic turnover rate or expression level. Moreover, two mutations that increase Na+ affinity were found to reduce [Na+]i. It is concluded that the Na+ affinity of the Na+,K+-ATPase is an important determinant of [Na+]i.  相似文献   

5.
6.
The Na+,K+-ATPase binds Na+ at three transport sites denoted I, II, and III, of which site III is Na+-specific and suggested to be the first occupied in the cooperative binding process activating phosphorylation from ATP. Here we demonstrate that the asparagine substitution of the aspartate associated with site III found in patients with rapid-onset dystonia parkinsonism or alternating hemiplegia of childhood causes a dramatic reduction of Na+ affinity in the α1-, α2-, and α3-isoforms of Na+,K+-ATPase, whereas other substitutions of this aspartate are much less disruptive. This is likely due to interference by the amide function of the asparagine side chain with Na+-coordinating residues in site III. Remarkably, the Na+ affinity of site III aspartate to asparagine and alanine mutants is rescued by second-site mutation of a glutamate in the extracellular part of the fourth transmembrane helix, distant to site III. This gain-of-function mutation works without recovery of the lost cooperativity and selectivity of Na+ binding and does not affect the E1-E2 conformational equilibrium or the maximum phosphorylation rate. Hence, the rescue of Na+ affinity is likely intrinsic to the Na+ binding pocket, and the underlying mechanism could be a tightening of Na+ binding at Na+ site II, possibly via movement of transmembrane helix four. The second-site mutation also improves Na+,K+ pump function in intact cells. Rescue of Na+ affinity and Na+ and K+ transport by second-site mutation is unique in the history of Na+,K+-ATPase and points to new possibilities for treatment of neurological patients carrying Na+,K+-ATPase mutations.  相似文献   

7.
Alkalosis impairs the natriuretic response to diuretics, but the underlying mechanisms are unclear. The soluble adenylyl cyclase (sAC) is a chemosensor that mediates bicarbonate-dependent elevation of cAMP in intracellular microdomains. We hypothesized that sAC may be an important regulator of Na+ transport in the kidney. Confocal images of rat kidney revealed specific immunolocalization of sAC in collecting duct cells, and immunoblots confirmed sAC expression in mouse cortical collecting duct (mpkCCDc14) cells. These cells exhibit aldosterone-stimulated transepithelial Na+ currents that depend on both the apical epithelial Na+ channel (ENaC) and basolateral Na+,K+-ATPase. RNA interference-mediated 60-70% knockdown of sAC expression comparably inhibited basal transepithelial short circuit currents (Isc) in mpkCCDc14 cells. Moreover, the sAC inhibitors KH7 and 2-hydroxyestradiol reduced Isc in these cells by 50-60% within 30 min. 8-Bromoadenosine-3′,5′-cyclic-monophosphate substantially rescued the KH7 inhibition of transepithelial Na+ current. Aldosterone doubled ENaC-dependent Isc over 4 h, an effect that was abolished in the presence of KH7. The sAC contribution to Isc was unaffected with apical membrane nystatin-mediated permeabilization, whereas the sAC-dependent Na+ current was fully inhibited by basolateral ouabain treatment, suggesting that the Na+,K+-ATPase, rather than ENaC, is the relevant transporter target of sAC. Indeed, neither overexpression of sAC nor treatment with KH7 modulated ENaC currents in Xenopus oocytes. ATPase and biotinylation assays in mpkCCDc14 cells demonstrated that sAC inhibition decreases catalytic activity rather than surface expression of the Na+,K+-ATPase. In summary, these results suggest that sAC regulates both basal and agonist-stimulated Na+ reabsorption in the kidney collecting duct, acting to enhance Na+,K+-ATPase activity.Maintenance of intracellular pH depends in part on the extracellular to intracellular Na+ gradient, and elevation of intracellular [Na+] can lead to acidification of the cytoplasm. It has been shown that acidification of the cytoplasm of cells from frog skin and toad bladder by increased partial pressure of CO2 reduces Na+ transport and permeability (1, 2). Conversely, the rise in plasma bicarbonate caused by metabolic alkalosis with chronic diuretic use has been shown to increase net renal Na+ reabsorption independently of volume status, electrolyte depletion, and/or increased aldosterone secretion (3, 4). However, the underlying mechanisms involved in these phenomena remain unclear.The soluble adenylyl cyclase (sAC)2 is a chemosensor that mediates the elevation of cAMP in intracellular microdomains (5-7). Unlike transmembrane adenylyl cyclases (tmACs), sAC is insensitive to regulation by forskolin or heterotrimeric G proteins (8) and is directly activated by elevations of intracellular calcium (9, 10) and/or bicarbonate ions (11). Thus, sAC mediates localized intracellular increases in cAMP in response to variations in bicarbonate levels or its closely related parameters, partial pressure of CO2 and pH. Mammalian sAC is more similar to bicarbonate-regulated cyanobacterial adenylyl cyclases than to other mammalian nucleotidyl cyclases, which may indicate that there is a unifying mechanism for the regulation of cAMP signaling by bicarbonate across biological systems. Although sAC appears to be encoded by a single gene, there is significant isoform diversity for this ubiquitously expressed enzyme (11, 12) generated by alternative splicing (reviewed in Ref. 13). sAC has been shown to regulate the subcellular localization and/or activity of membrane transport proteins such as the vacuolar H+-ATPase (V-ATPase) and cystic fibrosis transmembrane conductance regulator in epithelial cells (14, 15). Functional activity of sAC has been reported in the kidney (16), and sAC has been localized to epithelial cells in the distal nephron (14, 17).Given that natriuresis is decreased during metabolic alkalosis, when bicarbonate is elevated, and Na+ reabsorption is impaired by high partial pressure of CO2, we hypothesized that bicarbonate-regulated sAC may play a key role in the regulation of transepithelial Na+ transport in the distal nephron. Reabsorption of Na+ in the kidney and other epithelial tissues is mediated by the parallel operation of apical ENaC and basolateral Na+,K+-ATPase, and both transport proteins can be stimulated by cAMP via the cAMP-dependent protein kinase (PKA) (18, 53). The aims of this study were to investigate the role of sAC in the regulation of transepithelial Na+ transport in the kidney through the use of specific sAC inhibitors and electrophysiological measurements. We found that sAC inhibition blocks transepithelial Na+ reabsorption in polarized mpkCCDc14 cells under both basal and hormone-stimulated conditions. Selective membrane permeabilization studies revealed that although ENaC activity appears to be unaffected by sAC inhibition, flux through the Na+,K+-ATPase is sensitive to sAC modulation. Inhibiting sAC decreases ATPase activity without affecting plasma membrane expression of the pump; thus, tonic sAC activity appears to be required for Na+ reabsorption in kidney collecting duct.  相似文献   

8.
As a stable analog for ADP-sensitive phosphorylated intermediate of sarcoplasmic reticulum Ca2+-ATPase E1PCa2·Mg, a complex of E1Ca2·BeFx, was successfully developed by addition of beryllium fluoride and Mg2+ to the Ca2+-bound state, E1Ca2. In E1Ca2·BeFx, most probably E1Ca2·BeF3, two Ca2+ are occluded at high affinity transport sites, its formation required Mg2+ binding at the catalytic site, and ADP decomposed it to E1Ca2, as in E1PCa2·Mg. Organization of cytoplasmic domains in E1Ca2·BeFx was revealed to be intermediate between those in E1Ca2·AlF4 ADP (transition state of E1PCa2 formation) and E2·BeF3·(ADP-insensitive phosphorylated intermediate E2P·Mg). Trinitrophenyl-AMP (TNP-AMP) formed a very fluorescent (superfluorescent) complex with E1Ca2·BeFx in contrast to no superfluorescence of TNP-AMP bound to E1Ca2·AlFx. E1Ca2·BeFx with bound TNP-AMP slowly decayed to E1Ca2, being distinct from the superfluorescent complex of TNP-AMP with E2·BeF3, which was stable. Tryptophan fluorescence revealed that the transmembrane structure of E1Ca2·BeFx mimics E1PCa2·Mg, and between those of E1Ca2·AlF4·ADP and E2·BeF3. E1Ca2·BeFx at low 50–100 μm Ca2+ was converted slowly to E2·BeF3 releasing Ca2+, mimicking E1PCa2·Mg → E2P·Mg + 2Ca2+. Ca2+ replacement of Mg2+ at the catalytic site at approximately millimolar high Ca2+ decomposed E1Ca2·BeFx to E1Ca2. Notably, E1Ca2·BeFx was perfectly stabilized for at least 12 days by 0.7 mm lumenal Ca2+ with 15 mm Mg2+. Also, stable E1Ca2·BeFx was produced from E2·BeF3 at 0.7 mm lumenal Ca2+ by binding two Ca2+ to lumenally oriented low affinity transport sites, as mimicking the reverse conversion E2P· Mg + 2Ca2+E1PCa2·Mg.Sarcoplasmic reticulum Ca2+-ATPase (SERCA1a),2 a representative member of the P-type ion transporting ATPases, catalyze Ca2+ transport coupled with ATP hydrolysis (Fig. 1) (19). The enzyme forms phosphorylated intermediates from ATP or Pi in the presence of Mg2+ (1013). In the transport cycle, the enzyme is first activated by cooperative binding of two Ca2+ ions at high affinity transport sites (E2 to E1Ca2, steps 1–2) (14) and autophosphorylated at Asp351 with MgATP to form the ADP-sensitive phosphoenzyme (E1P, step 3), which reacts with ADP to regenerate ATP in the reverse reaction. Upon this E1P formation, the two bound Ca2+ are occluded in the transport sites (E1PCa2). Subsequent isomeric transition to the ADP-insensitive form (E2PCa2), i.e. loss of ADP sensitivity at the catalytic site, results in rearrangement of the Ca2+ binding sites to deocclude Ca2+, reduce the affinity, and open the lumenal gate, thus releasing Ca2+ into the lumen (E2P, steps 4–5). Finally Asp351-acylphosphate in E2P is hydrolyzed to form the Ca2+-unbound inactive E2 state (steps 6 and 7). Mg2+ bound at the catalytic site is required as a physiological catalytic cofactor in phosphorylation and dephosphorylation and thus for the transport cycle. The cycle is totally reversible, e.g. E2P can be formed from Pi in the presence of Mg2+ and absence of Ca2+, and subsequent Ca2+ binding at lumenally oriented low affinity transport sites of E2P reverses the Ca2+-releasing step and produces E1PCa2, which is then decomposed to E1Ca2 by ADP.Open in a separate windowFIGURE 1.Ca2+ transport cycle of Ca2+-ATPase.Various intermediate structural states in the transport cycle were fixed as their structural analogs produced by appropriate ligands such as AMP-PCP (non-hydrolyzable ATP analog) or metal fluoride compounds (phosphate analogs), and their crystal structures were solved at the atomic level (1522). The three cytoplasmic domains, N, P, and A, largely move and change their organization state during the transport cycle, and the changes are coupled with changes in the transport sites. Most remarkably, in the change from E1Ca2·AlF4·ADP (the transition state for E1PCa2 formation, E1PCa2·ADP·Mg) to E2·BeF3 (the ground state E2P·Mg) (2325), the A domain largely rotates by more than 90° approximately parallel to the membrane plane and associates with the P domain, thereby destroying the Ca2+ binding sites, and opening the lumenal gate, thus releasing Ca2+ into the lumen (see Fig. 2). E1PCa2·Ca·AMP-PN formed by CaAMP-PNP without Mg2+ is nearly the same as E1Ca2·AlF4·ADP and E1Ca2·CaAMP-PCP in their crystal structures (17, 18, 22).Open in a separate windowFIGURE 2.Structure of SERCA1a and its change during processing of phosphorylated intermediate. E1Ca2·AlF4·ADP (the transition state analog for phosphorylation E1PCa2·ADP·Mg) and E2·BeF3 (the ground state E2P analog (25)) were obtained from the Protein Data Bank (PDB accession code 1T5T (17) and 2ZBE (21), respectively). Cytoplasmic domains N (nucleotide binding), P (phosphorylation), and A (actuator), and 10 transmembrane helices (M1–M10) are indicated. The arrows on the domains, M1′ and M2 (Tyr122) in E1Ca2·AlF4·ADP, indicate their approximate motions predicted for E1PCa2·ADP·MgE2P·Mg. The phosphorylation site Asp351, TGES184 of the A domain, Arg198 (tryptic T2 site) on the Val200 loop (DPR198AV200NQD) of the A domain, and Thr242 (proteinase K site) on the A/M3-linker are shown. Seven hydrophobic residues gather in the E2P state to form the Tyr122-hydrophobic cluster (Y122-HC); Tyr122/Leu119 on the top part of M2, Ile179/Leu180/Ile232 of the A domain, and Val705/Val726 of the P domain. The overall structure of E1Ca2·AlF4·ADP is virtually the same as those of E1Ca2·CaAMP-PCP and E1PCa2·Ca·AMP-PN (17, 18, 22).Despite these atomic structures, yet unsolved is the structure of E1PCa2·Mg, the genuine physiological intermediate E1PCa2 with bound Mg2+ at the catalytic site without the nucleotide. Its stable structural analog has yet to be developed. E1PCa2·Mg is the major intermediate accumulating almost exclusively at steady state under physiological conditions. Its rate-limiting isomerization results in Ca2+ deocclusion/release producing E2P·Mg as a key event for Ca2+ transport. In E1Ca2·CaAMP-PCP, E1Ca2·AlF4·ADP, and E1PCa2·Ca·AMP-PN, the N and P domains are cross-linked and strongly stabilized by the bound nucleotide and/or Ca2+ at the catalytic site, thus they are crystallized (17, 18, 22). Kinetically, E1PCa2·Ca formed with CaATP is markedly stabilized due to Ca2+ binding at the catalytic Mg2+ site, and its isomerization to E2P is strongly retarded in contrast to E1PCa2·Mg (26, 27). Thus, the bound Ca2+ at the catalytic Mg2+ site likely produces a significantly different structural state from that with bound Mg2+.Therefore, it is now essential to develop a genuine E1PCa2·Mg analog without bound nucleotide and thereby gain further insight into the structural mechanism in the Ca2+ transport process. It is also crucial to further clarify the structural importance of Mg2+ as the physiological catalytic cation. In this study, we successfully developed the complex E1Ca2·BeFx, most probably E1Ca2·BeF3, as the E1PCa2·Mg analog by adding beryllium fluoride (BeFx) to the E1Ca2 state without any nucleotides. For its formation, Mg2+ binding at the catalytic site was required and Ca2+ substitution for Mg2+ was absolutely unfavorable, revealing a likely structural reason for its preference as the physiological cofactor. In E1Ca2·BeF3, two Ca2+ ions bound at the high affinity transport sites are occluded. It was also produced from E2·BeF3 by lumenal Ca2+ binding at the lumenally oriented low affinity transport sites, mimicking E2P·Mg + 2Ca2+E1PCa2·Mg. All properties of the newly developed E1Ca2·BeF3 fulfilled the requirements as the E1PCa2·Mg analog, and hence we were able to uncover the hitherto unknown nature of E1PCa2·Mg as well as structural events occurring in the phosphorylation and isomerization processes. Also, we successfully found the conditions that perfectly stabilize the E1Ca2·BeF3 complex.  相似文献   

9.
10.
The airway surface liquid (ASL) is the thin fluid layer lining airway surface epithelial cells, whose volume and composition are tightly regulated and may be abnormal in cystic fibrosis (CF). We synthesized a two-color fluorescent dextran to measure ASL [K+], TAC-Lime-dextran-TMR, consisting of a green-fluorescing triazacryptand K+ ionophore-Bodipy conjugate, coupled to dextran, together with a red fluorescing tetramethylrhodamine reference chromophore. TAC-Lime-dextran-TMR fluorescence was K+-selective, increasing >4-fold with increasing [K+] from 0 to 40 mm. In well differentiated human airway epithelial cells, ASL [K+] was 20.8 ± 0.3 mm and decreased by inhibition of the Na+/K+ pump (ouabain), ENaC (amiloride), CF transmembrane conductance regulator (CFTRinh-172), or K+ channels (TEA or XE991). ASL [K+] was increased by forskolin but not affected by Na+/K+/2Cl cotransporter inhibition (bumetanide). Functional and expression studies indicated the involvement of [K+] channels KCNQ1, KCNQ3, and KCNQ5 as determinants of ASL [K+]. [K+] in CF cultures was similar to that in non-CF cultures, suggesting that abnormal ASL [K+] is not a factor in CF lung disease. In intact airways, ASL [K+] was also well above extracellular [K+]: 22 ± 1 mm in pig trachea ex vivo and 16 ± 1 mm in mouse trachea in vivo. Our results provide the first noninvasive measurements of [K+] in the ASL and indicate the involvement of apical and basolateral membrane ion transporters in maintaining a high ASL [K+].The airway surface liquid (ASL)2 is the thin layer of aqueous fluid that lines the mucosal surface of the airways, forming the interface between airway epithelial cells and the gas phase. The ASL contains water, ions, and macromolecules. It is believed that ASL volume and composition are tightly regulated to maintain a nonviscous fluid layer for mucociliary clearance by underlying epithelial cells and to support the intrinsic antimicrobial function of defensins and other macromolecules (1, 2). Abnormalities in ASL volume and/or composition are proposed to be important in the pathogenesis of cystic fibrosis (CF) lung disease (3). The ASL is formed by a combination of fluid secretion by airway submucosal glands, convective fluid transport up the airway tree, and water/ion transport by airway surface epithelial cells, the latter likely playing a key role in active regulation of ASL volume and ionic composition.Although early data suggested that salt concentration in the ASL is low in normal airways (4), the current view is that the ASL is approximately isotonic in both normal and CF airways (57). A concern with older measurements of ASL composition involving fluid collection by filter paper or microcapillaries is perturbation of the airway surface and the sampling, by capillary forces, of more fluid than contained in the very thin (tens of microns) ASL layer. Studies using ion-sensitive microelectrodes, although technically demanding and requiring direct contact with the ASL, provided evidence for a nearly isotonic ASL (5). Our laboratory developed ratioable fluorescent dyes to measure ASL [Na+] and [Cl], in which the ASL was fluorescently stained for determination of ion concentrations by ratio imaging microscopy (7). ASL salt concentration ([Na+] and [Cl]) was found to be approximately isotonic in airway epithelial cell cultures, mouse trachea and small airways, and ex vivo human airways, without differences in CFTR deficiency (7, 8). We also found the ASL to be approximately isosmolar with serum using fluorescent, osmotically sensitive liposomes (9).Relatively little is known about ASL potassium concentration ([K+]) or its regulation. As diagrammed in Fig. 1A, transcellular transport of K+ is believed to involve the coordinated activities of a Na+/K+ pump, Na+/K+/2Cl cotransporter, and K+ channel(s) at the basolateral membrane and a H+/K+ pump and K+ channel(s) at the apical membrane. The airway epithelium also has significant paracellular ion permeability. There is evidence for functional expression of several types of K+ channels in cell lines derived from airways/lung, including Ca2+-activated, cAMP-activated, and voltage-activated K+ channels (1013). Steady state [K+] in a stationary ASL (without fluid convection) should depend on the activities of cell membrane K+ pumps and ion transporters, as well as non-K+ ion channels, such as CFTR and ENaC, which are involved in establishing membrane potentials and thus the electrochemical driving forces for K+ transport.Open in a separate windowFIGURE 1.Cell model and perfusion chamber for measurements of ASL K+ concentration. A, schematic of airway epithelium showing principal ion transporters on the apical and basolateral plasma membranes, and paracellular pathway. B, schematic of perfusion chamber. Cells on a porous filter, facing upward, are imaged from above after fluorescent dye staining of the ASL. The under surface of the porous filter is perfused continuously. C, short circuit current (Isc) in HBE cell monolayers in response to the additions of amiloride (10 μm), forskolin (10 μm), CFTRinh-172 (10 μm), ATP (100 μm), and CaCCinh-A01 (30 μm), a CaCC-specific inhibitor. D, transepithelial PD in response to amiloride, forskolin, CFTRinh-172, ATP, and CaCCinh-A01. The data in C and D are representative of four sets of measurements.The purpose of this study was to develop a noninvasive fluorescence method to measure [K+] in the ASL and to establish the major determinants of [K+] regulation. Following several years of synthetic chemistry, we developed a series of water-soluble K+ sensors, the first being TAC-Red, in which K+ binding to a triazacryptand (TAC) K+ ionophore results in fluorescence enhancement of a conjugated xanthylium chromophore by a charge transfer quenching mechanism (14). TAC-Red was used to follow K+ waves in the extracellular space in brain in a neuroexcitation model of cortical spreading depression. Second generation K+ sensors of different colors, TAC-Crimson (15) and TAC-Lime (16), work by a similar K+-sensing mechanism but utilize different chromophores. These indicators are selective for K+ under physiological conditions and respond to changes in [K+] in milliseconds or less (15). For the measurements here, we synthesized a dextran conjugate containing TAC-Lime, which has K+-sensitive green fluorescence, and tetramethylrhodamine, a reference chromophore with K+-insensitive red fluorescence. The indicator allowed technically straightforward determination of ASL [K+] in cell culture and in vivo models by fluorescence ratio imaging.  相似文献   

11.
12.
Ion translocation by the sarcoplasmic reticulum Ca2+-ATPase depends on large movements of the A-domain, but the driving forces have yet to be defined. The A-domain is connected to the ion-binding membranous part of the protein through linker regions. We have determined the functional consequences of changing the length of the linker between the A-domain and transmembrane helix M3 (“A-M3 linker”) by insertion and deletion mutagenesis at two sites. It was feasible to insert as many as 41 residues (polyglycine and glycine-proline loops) in the flexible region of the linker without loss of the ability to react with Ca2+ and ATP and to form the phosphorylated Ca2E1P intermediate, but the rate of the energy-transducing conformational transition to E2P was reduced by >80%. Insertion of a smaller number of residues gave effects gradually increasing with the length of the insertion. Deletion of two residues at the same site, but not replacement with glycine, gave a similar reduction as the longest insertion. Insertion of one or three residues in another part of the A-M3 linker that forms an α-helix (“A3 helix”) in E2/E2P conformations had even more profound effects on the ability of the enzyme to form E2P. These results demonstrate the importance of the length of the A-M3 linker and of the position and integrity of the A3 helix for stabilization of E2P and suggest that, during the normal enzyme cycle, strain of the A-M3 linker could contribute to destabilize the Ca2E1P state and thereby to drive the transition to E2P.The sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA)2 is a membrane-bound ion pump that transports Ca2+ against a steep concentration gradient, utilizing the energy derived from ATP hydrolysis (13). It belongs to the family of P-type ATPases, in which the γ-phosphoryl group of ATP is transferred to a conserved aspartic acid residue during the reaction cycle. Both phospho and dephospho forms of the enzyme undergo transitions between so-called E1 and E2 conformations (Scheme 1). The E1 and E1P states display specificity for reaction with ATP and ADP, respectively (“kinase activity”), whereas E2P and E2 react with water and Pi instead of nucleotide (“phosphatase activity”). The E1 dephosphoenzyme of the Ca2+-ATPase binds two Ca2+ ions with high affinity from the cytoplasmic side, thereby triggering the phosphorylation from ATP. In E1P, the Ca2+ ions are occluded with no access to either side of the membrane, and Ca2+ is released to the luminal side after the conformational transition to E2P, likely in exchange for protons being countertransported. The structural organization and domain movements leading to Ca2+ translocation have recently been elucidated by crystallization of SERCA in various conformational states thought to represent intermediates in the pump cycle (47). SERCA is made up of 10 membrane-spanning mostly helical segments, M1–M10 (numbered from the N terminus), of which M4–M6 and M8 contribute liganding groups for Ca2+ binding, and a cytoplasmic headpiece separated into three distinct domains, named A (“actuator”), P (“phosphorylation”), and N (“nucleotide binding”). The A-domain appears to undergo considerable movement during the functional cycle. In the E1/E1P states, the highly conserved TGE183S loop of the A-domain is at great distance from the catalytic center containing nucleotide-binding residues and the phosphorylated Asp351 of the P-domain, but during the Ca2E1P → E2P transition, the A-domain rotates ∼90° around an axis perpendicular to the membrane, thereby moving the TGE183S loop into close contact with the catalytic site such that Glu183 can catalyze dephosphorylation of E2P (8, 9). During the dephosphorylation, Glu183 likely coordinates the water molecule attacking the aspartyl phosphoryl bond and withdraws a hydrogen. Hence, the movement of the A-domain during the Ca2E1P → E2P transition is the event that changes the catalytic specificity from kinase activity to phosphatase activity. During the dephosphorylation of E2P → E2, there is only a slight change of the position of the A-domain, and a large back-rotation is needed to reach the E1 form from E2; thus, the A-domain rotation defines the difference between the E1/E1P class of conformations and the E2/E2P class. Because the A-domain is physically connected to transmembrane helices M1–M3 through the linker segments A-M1, A-M2, and A-M3, the A-domain movement occurring during the Ca2E1P → E2P transition may be a key event in the opening of the Ca2+ sites toward the lumen, thus explaining the coupling of ATP hydrolysis to Ca2+ translocation. An important unanswered question is, however, how the movement of the A-domain is brought about. Which are the driving forces that destabilize Ca2E1P and/or stabilize E2P such that the energy-transducing Ca2E1P → E2P transition takes place? To answer this, it seems important to elucidate the exact roles of the linkers. Intriguing results have been obtained by Suzuki and co-workers, who demonstrated the importance of the A-M1 linker in connection with luminal release of Ca2+ from E2P (10). In this study, we have addressed the role of the A-M3 linker. An alignment of two crystal structures thought to resemble the Ca2E1P and E2·Pi forms (5), respectively, is shown in Fig. 1. The A-domain rotation is associated with formation of a helix (“A3 helix”) in the N-terminal part of the A-M3 linker, and this helix seems to interact with a helix bundle consisting of the P5–P7 helices of the P-domain, a feature exhibited by all published crystal structures of the E2 type (cf. supplemental Fig. S1 and Ref. 11). Moreover, when structures of similar crystallographic resolution are compared (as in Fig. 1), the non-helical part of the A-M3 linker in E2-type structures has a higher relative temperature factor (“B-factor”) than the corresponding segment in Ca2E1P (Fig. 1C, thick part colored orange-red for high temperature factor), thus suggesting a higher degree of freedom of movement relative to Ca2E1P. Hence, the A-M3 linker appears more strained in Ca2E1P compared with E2 forms, and the greater flexibility of the linker in E2 forms may promote the formation of the A3 helix.Open in a separate windowSCHEME 1.Ca2+-ATPase reaction cycle.Open in a separate windowFIGURE 1.A-M3 linker configuration in E1- and E2-type crystal structures. Crystal structures with Protein Data Bank codes 2zbd (Ca2E1P analog) and 1wpg (E2·Pi analog) are shown aligned. A, overview of structure 2zbd in bluish colors with green A-M3 linker and structure 1wpg in reddish colors with wheat A-M3 linker. B, magnification of the A-M3 linker (corresponding to the red box in A) with arrows indicating site 1, between Glu243 and Gln244, and site 2, between Gly233 and Lys234, in both conformations. The green A-M3 linker to the right is structure 2zbd. The wheat A-M3 linker to the left is structure 1wpg. Note the kinked A3 helix forming part of the latter structure. C, same A-M3 linker structures as in B but with the magnitude of the temperature factor (B-factor) indicated in colors (red > orange > yellow > green > blue) and by tube diameter. Because the two crystal structures selected here as E1- and E2-type representatives have similar crystallographic resolution (2.40 and 2.30 Å, respectively), the differences in temperature factor in specific regions provide direct information about chain flexibility.Here, we have determined the functional consequences of changing the length (and thereby likely the strain) of the A-M3 linker. Polyglycine and glycine-proline loops of varying lengths were inserted at two different sites in the linker (Fig. 1), and deletions were also studied. Rather unexpectedly, we were able to insert as many as 41 residues in one of the sites without loss of expression or ability to react with Ca2+ and ATP, forming Ca2E1P, but the Ca2E1P → E2P transition was greatly affected.  相似文献   

13.
Phospholemman (PLM) phosphorylation mediates enhanced Na/K-ATPase (NKA) function during adrenergic stimulation of the heart. Multiple NKA isoforms exist, and their function/regulation may differ. We combined fluorescence resonance energy transfer (FRET) and functional measurements to investigate isoform specificity of the NKA-PLM interaction. FRET was measured as the increase in the donor fluorescence (CFP-NKA-α1 or CFP-NKA-α2) during progressive acceptor (PLM-YFP) photobleach in HEK-293 cells. Both pairs exhibited robust FRET (maximum of 23.6 ± 3.4% for NKA-α1 and 27.5 ± 2.5% for NKA-α2). Donor fluorescence depended linearly on acceptor fluorescence, indicating a 1:1 PLM:NKA stoichiometry for both isoforms. PLM phosphorylation induced by cAMP-dependent protein kinase and protein kinase C activation drastically reduced the FRET with both NKA isoforms. However, submaximal cAMP-dependent protein kinase activation had less effect on PLM-NKA-α2 versus PLM-NKA-α1. Surprisingly, ouabain virtually abolished NKA-PLM FRET but only partially reduced co-immunoprecipitation. PLM-CFP also showed FRET to PLM-YFP, but the relationship during progressive photobleach was highly nonlinear, indicating oligomers involving ≥3 monomers. Using cardiac myocytes from wild-type mice and mice where NKA-α1 is ouabain-sensitive and NKA-α2 is ouabain-resistant, we assessed the effects of PLM phosphorylation on NKA-α1 and NKA-α2 function. Isoproterenol enhanced internal Na+ affinity of both isoforms (K½ decreased from 18.1 ± 2.0 to 11.5 ± 1.9 mm for NKA-α1 and from 16.4 ± 2.5 to 10.4 ± 1.5 mm for NKA-α2) without altering maximum transport rate (Vmax). Protein kinase C activation also decreased K½ for both NKA-α1 and NKA-α2 (to 9.4 ± 1.0 and 9.1 ± 1.1 mm, respectively) but increased Vmax only for NKA-α2 (1.9 ± 0.4 versus 1.2 ± 0.5 mm/min). In conclusion, PLM associates with and modulates both NKA-α1 and NKA-α2 in a comparable but not identical manner.Cardiac Na/K-ATPase (NKA)3 regulates intracellular Na+, which in turn affects intracellular Ca2+ and contractility via Na+/Ca2+ exchange. Members of the FXYD family of small, single membrane-spanning proteins, including phospholemman (PLM) and the NKA γ-subunit (1), have emerged recently as tissue-specific regulators of NKA. PLM is the only FXYD protein known to be highly expressed in cardiac myocytes and is also unique within the family in that it is phosphorylated at two or more sites by cAMP-dependent protein kinase (PKA) and protein kinase C (PKC) (2, 3). In the heart, PLM is a major phosphorylation target for both PKA and PKC.Co-immunoprecipitation experiments have demonstrated that PLM is physically associated with NKA (48), and this is not affected by PLM phosphorylation (6, 7). We have shown recently (9) that PLM and NKA are in very close proximity, such that fluorescence resonance energy transfer (FRET) occurs. PLM phosphorylation by either PKA or PKC reduces the FRET significantly, suggesting that although PLM and NKA are not physically dissociated upon phosphorylation, their interaction is altered. PLM inhibits NKA (4, 8, 10, 11), mostly by reducing the affinity of the pump for internal Na+. PLM phosphorylation relieves this inhibition and thus mediates the enhancement of NKA function by α- and β-adrenergic stimulation in mouse ventricular myocytes (10, 11).There are multiple NKA isoforms in cardiac myocytes. NKA-α1 is the dominant, ubiquitous isoform, whereas NKA-α2 and NKA-α3 are present in relatively small amounts and in a species-dependent manner (12). For instance, the adult rodent heart expresses NKA-α1 and NKA-α2, although dogs and monkeys do not have the NKA-α2 subunit (13). In humans all three NKA-α isoforms can be detected (14). It has been suggested that NKA-α2 and NKA-α3 are located mainly in the T-tubules, at the junctions with the sarcoplasmic reticulum, where they could regulate local Na+/Ca2+ exchange and thus cardiac myocyte Ca2+. There is rather convincing evidence supporting such a model in the smooth muscle (15). However, things are less clear in the heart. The functional density of NKA-α2 is significantly higher in the T-tubules (versus external sarcolemma) in cardiac myocytes from both rats (16, 17) and mice (18), but their precise localization with respect to the junctions with the sarcoplasmic reticulum is not known. Based on Ca2+ transients from heterozygous NKA-α1+/− and NKA-α2+/− mice, James et al. (19) concluded that NKA-α2 is involved in cardiac myocyte Ca2+ regulation, whereas NKA-α1 is not. Further support for this idea came from the observation that replacing mouse NKA-α2 with a low affinity mutant leads to a loss of glycoside inotropy (20), and increased expression of NKA-α2 decreased the Na+/Ca2+ exchange current and Ca2+ transients (21). However, other findings challenge the preferential role of NKA-α2 in regulating intracellular Ca2+ and contractility. Moseley et al. (22) showed that NKA-α1+/− mice were severely compromised, and Dostanic et al. (23) showed that NKA-α1 is also physically and functionally associated with the Na+/Ca2+ exchanger.In this context, it is important to determine whether NKA-α1 and NKA-α2 interact differently with PLM. The data available so far on this are contradictory. We have found (7) that NKA-α1, NKA-α2, and NKA-α3 isoforms co-immunoprecipitate PLM, both unphosphorylated and phosphorylated, in rabbit heart. In contrast, Silverman et al. (8) reported that NKA-α1 but not NKA-α2 co-immunoprecipitate with PLM in ventricular myocytes from guinea pig. The functional data are also contradictory. PLM was found to reduce the affinity for Na+ of both NKA-α1 and NKA-α2 isoforms in a heterologous expression system (4), whereas Silverman et al. (8) reported that forskolin-induced PLM phosphorylation results in a higher NKA-α1-mediated current and no change in the current generated by NKA-α2.Here we used two methods to investigate whether the interaction and functional effects of PLM on NKA are NKA-α isoform-specific. First, we used FRET to assess the interaction between PLM-YFP and CFP-NKA-α1/CFP-NKA-α2 transfected in HEK-293 cells and how PLM phosphorylation by PKA and PKC affects this interaction. Second, we measured NKA function in myocytes isolated from wild-type (WT) mice and mice where NKA isoforms have swapped ouabain affinities (SWAP; NKA-α1 is ouabain-sensitive, whereas NKA-α2 is ouabain-resistant) (23). In this way we could test the effect of β-adrenergic stimulation separately on NKA-α1 and NKA-α2 isoforms in the native myocyte environment, as an indicator of the functional interaction with PLM. Our results indicate that NKA-α1 and NKA-α2 interact similarly with PLM, and this interaction is equally affected by PLM phosphorylation.  相似文献   

14.
  • 1.1. The (Na+ + K+)- and Na+-ATPases, both present in kidney microsomes of Sparus auratus L., have different activities and optimal assay conditions as, in the first of the two stocks of fish used (A), the spec. act. of the former is 51.7 μmol Pi mg prot−1 hr−1 at pH 7.5, 100 mM Na+, 10 mM K+, 17.5 mM Mg2+, 7.5 mM ATP and that of the latter is 6.5 μmol Pi mg prot−1 hr−1 at pH 6.5, 40 mM Na+, 4.0 mM Mg2+, 2.5 mM ATP.
  • 2.2. Ouabain and vanadate specifically inhibit the (Na+ + K+)-ATPase but not the Na+-ATPase that is preferentially inhibited by ethacrynic acid.
  • 3.3. While the (Na+ + K+)-ATPase is strictly specific for ATP and Na+, Na+-ATPase can be activated by various monovalent cations and, apart from ATP, hydrolyses CTP, though less efficiently.
  • 4.4. The second stock B, subjected to higher salinity than A, shows an acidic shifted Na+-ATPase optimal pH, opposed to the stability of that of the (Na+ + K+)-ATPase, a decreased (Na+ + K+)-ATPase and a strikingly depressed Na+-ATPase.
  • 5.5. The results are compared with literature data and discussed on the basis of the presumptive different roles as well as functional prevalence in various salinities of the two ATPases.
  相似文献   

15.
Roles of hydrogen bonding interaction between Ser186 of the actuator (A) domain and Glu439 of nucleotide binding (N) domain seen in the structures of ADP-insensitive phosphorylated intermediate (E2P) of sarco(endo)plasmic reticulum Ca2+-ATPase were explored by their double alanine substitution S186A/E439A, swap substitution S186E/E439S, and each of these single substitutions. All the mutants except the swap mutant S186E/E439S showed markedly reduced Ca2+-ATPase activity, and S186E/E439S restored completely the wild-type activity. In all the mutants except S186E/E439S, the isomerization of ADP-sensitive phosphorylated intermediate (E1P) to E2P was markedly retarded, and the E2P hydrolysis was largely accelerated, whereas S186E/E439S restored almost the wild-type rates. Results showed that the Ser186-Glu439 hydrogen bond stabilizes the E2P ground state structure. The modulatory ATP binding at sub-mm∼mm range largely accelerated the EP isomerization in all the alanine mutants and E439S. In S186E, this acceleration as well as the acceleration of the ATPase activity was almost completely abolished, whereas the swap mutation S186E/E439S restored the modulatory ATP acceleration with a much higher ATP affinity than the wild type. Results indicated that Ser186 and Glu439 are closely located to the modulatory ATP binding site for the EP isomerization, and that their hydrogen bond fixes their side chain configurations thereby adjusts properly the modulatory ATP affinity to respond to the cellular ATP level.Sarcoplasmic reticulum Ca2+-ATPase (SERCA1a)2 is a representative member of P-type ion-transporting ATPases and catalyzes Ca2+ transport coupled with ATP hydrolysis (Fig. 1) (19). In the catalytic cycle, the enzyme is activated by binding of two Ca2+ ions at the transport sites (E2 to E1Ca2, steps 1–2) and then autophosphorylated at Asp351 with MgATP to form ADP-sensitive phosphoenzyme (E1P, step 3), which can react with ADP to regenerate ATP. Upon formation of this EP, the bound Ca2+ ions are occluded in the transport sites (E1PCa2). The subsequent isomeric transition to ADP-insensitive form (E2P) results in a change in the orientation of the Ca2+ binding sites and reduction of their affinity, and thus Ca2+ release into lumen (steps 4 and 5). Finally, the hydrolysis takes place and returns the enzyme into an unphosphorylated and Ca2+-unbound form (E2, step 6). E2P can also be formed from Pi in the presence of Mg2+ and the absence of Ca2+ by reversal of its hydrolysis.Open in a separate windowFIGURE 1.Reaction cycle of sarco(endo)plasmic reticulum Ca2+-ATPase.The cytoplasmic three domains N, A, and P largely move and change their organization states during the Ca2+ transport cycle (1022). These changes are linked with the rearrangements in the transmembrane helices. In the EP isomerization (loss of ADP sensitivity) and Ca2+ release, the A domain largely rotates (by ∼110° parallel to membrane plane), intrudes into the space between the N and P domains, and the P domain largely inclines toward the A domain. Thus in E2P, these domains produce the most compactly organized state (see Fig. 2 for the change E1Ca2·AlF4·ADP →E2·MgF42− as the model for the overall process E1PCa2·ADPE2·Pi).Open in a separate windowFIGURE 2.Structure of SERCA1a and formation of Ser186-Glu439 hydrogen bond between the A and N domains. The coordinates for the structures E1Ca2·AlF4·ADP, (the analog for the transition state of the phosphoryl transfer E1PCa2·ADP, left panel) and E2·MgF42− (E2·Pi analog (21), right panel) of Ca2+-ATPase were obtained from the Protein Data Bank (PDB accession code 1T5T and 1WPG, respectively (12, 14)). The arrows indicate approximate movements of the A and P domains in the change from E1Ca2·AlF4 ·ADP to E2·MgF42−. Ser186 and Glu439 are depicted as van der Waals spheres. These two residues form a hydrogen bond in E2·MgF42− (see inset). The phosphorylation site Asp351, two Ca2+ at the transport sites and ADP with AlF4 at the catalytic site in E1Ca2·AlF4·ADP, MgF42− bound at the catalytic site in E2·MgF42− are depicted. The TGES184 loop and Val200 loop of the A domain and Tyr122 on the top part of M2 are shown. These elements produce three interaction networks between A and P domains and M2 (Tyr122) in E2·MgF42− (2326). M1′ and M1-M10 are also indicated.We have found that the interactions between the A and P domains at the Val200-loop (Asp196-Asp203) with the residues of the P domain (Arg678/Glu680/Arg656/Asp660) (23) and at the Tyr122 hydrophobic cluster (2426) (see Fig. 2) play critical roles for Ca2+ deocclusion/release in E2PCa2E2P + 2Ca2+ after the loss of ADP sensitivity (E1PCa2 to E2PCa2 isomerization). The proper length of the A/M1′ linker is critical for inducing the inclining motion of the A and P domains for the Ca2+ deocclusion and release from E2PCa2 (27, 28). The importance of the interdomain interaction between Arg678 (P) and Asp203 (A) in stabilizing the E2P and E2 intermediates and its influence on modulatory ATP activation were pointed out by the mutation R678A (29). Regarding the N domain, the importance of Glu439 in the EP isomerization and E2P hydrolysis was previously noted by its alanine substitution, and possible importance of its interaction with Ser186 on the A domain has been suggested since Glu439 forms a hydrogen bond with Ser186 in the E2P analog structures (29) (see Fig. 2). The Darier disease-causing mutations of Ser186 of SERCA2b, S186P and S186F also alter the kinetics of the EP processing and its importance as the residue in the immediate vicinity of TGES184 has been pointed out (30, 31). Notably also, Glu439 is situated near the adenine binding pocket and its importance in the ATP binding and ATP-induced structural change have been shown (32, 33). In the structure E2(TG)AMPPCP (E2·ATP), Glu439 interacts with the modulatory ATP binding via Mg2+, and is involved in the acceleration of the Ca2+-ATPase cycle (16).Considering these critical findings on each of Glu439 and Ser186, it is crucial to reveal the role of the Ser186-Glu439 hydrogen-bonding interaction between the A and N domains in the EP processing and its ATP modulation (i.e. regulatory ATP-induced acceleration). We therefore made a series of mutants on both Ser186 and Glu439 including the swap substitution mutant, S186A, E439A, S186A/E439A, S186E, E439S, S186E/E439S, and explored their kinetic properties. Results showed that the Ser186-Glu439 hydrogen bond is critical for the stabilization of the E2P ground state structure, and possibly functioning as to make the E2P resident time long enough for Ca2+ release (E2PCa2E2P + 2Ca2+) thus to avoid its hydrolysis without Ca2+ release. Results also revealed that the side-chain configurations of Ser186 and Glu439 are fixed by their hydrogen bond, thereby conferring the proper modulatory ATP binding to occur at the cellular ATP level to accelerate the rate-limiting EP isomerization.  相似文献   

16.
Magi 4, now renamed δ-hexatoxin-Mg1a, is a 43-residue neurotoxic peptide from the venom of the hexathelid Japanese funnel-web spider (Macrothele gigas) with homology to δ-hexatoxins from Australian funnel-web spiders. It binds with high affinity to receptor site 3 on insect voltage-gated sodium (NaV) channels but, unlike δ-hexatoxins, does not compete for the related site 3 in rat brain despite being previously shown to be lethal by intracranial injection. To elucidate differences in NaV channel selectivity, we have undertaken the first characterization of a peptide toxin on a broad range of mammalian and insect NaV channel subtypes showing that δ-hexatoxin-Mg1a selectively slows channel inactivation of mammalian NaV1.1, NaV1.3, and NaV1.6 but more importantly shows higher affinity for insect NaV1 (para) channels. Consequently, δ-hexatoxin-Mg1a induces tonic repetitive firing of nerve impulses in insect neurons accompanied by plateau potentials. In addition, we have chemically synthesized and folded δ-hexatoxin-Mg1a, ascertained the bonding pattern of the four disulfides, and determined its three-dimensional solution structure using NMR spectroscopy. Despite modest sequence homology, we show that key residues important for the activity of scorpion α-toxins and δ-hexatoxins are distributed in a topologically similar manner in δ-hexatoxin-Mg1a. However, subtle differences in the toxin surfaces are important for the novel selectivity of δ-hexatoxin-Mg1a for certain mammalian and insect NaV channel subtypes. As such, δ-hexatoxin-Mg1a provides us with a specific tool with which to study channel structure and function and determinants for phylum- and tissue-specific activity.Voltage-gated sodium (NaV)4 channels are responsible for the generation and propagation of electrical signals in excitable cells. At least nine different genes encoding distinct NaV channels isoforms have been identified, and functionally expressed, in mammals (1). They are characterized by their sensitivity to TTX, with NaV1.5, NaV1.8, and NaV1.9 being TTX-insensitive or TTX-resistant, and the remaining subtypes being sensitive to nanomolar concentrations of TTX. In addition, localization of the subtypes also varies, with NaV1.1–1.3 mostly distributed in the central nervous system, NaV1.6–1.9 principally located in the peripheral nervous system, and NaV1.4 and NaV1.5 found in skeletal and cardiac muscle, respectively. The structural diversity of NaV channels also coincides with variations in physiological and pharmacological properties (2). In contrast, insects express only one gene (para) that undergoes extensive alternative splicing and RNA editing (3). The para-encoded NaV channel is exceptionally well conserved across diverse orders of insects, with the level of identity ranging from 87 to 98% (3). This is one reason why insecticides that target insect NaV channels have broad activity across many insect orders. In contrast, para-type NaV channels have significantly lower levels of identity with the various types of mammalian NaV channels with the level of identity typically around 50–60% (3). This explains why a high degree of phylogenetic specificity can be achieved with both NaV channel toxins and insecticides that target the NaV channel.At least seven distinct toxin-binding sites have been identified by radioligand binding and electrophysiological studies on vertebrate and insect NaV channels (4, 5). Toxins interacting with these neurotoxin receptor sites have been instrumental in the study of NaV channel topology, function, and pharmacology (6). In particular, a wide range of scorpion α-toxins, sea anemone toxins, and spider δ-hexatoxins (formerly δ-atracotoxins (7)) compete for binding to receptor site-3 on the extracellular surface of NaV channels. These polypeptide toxins all inhibit the fast inactivation of NaV channels to prolong Na+ currents (INa), despite huge diversity in primary and tertiary structures (8, 9). Nevertheless, receptor site-3 has not yet been fully characterized but is believed to involve domains DI/S5-S6, DIV/S5-S6, as well as DIV/S3-S4 (9). Most importantly, however, toxin characterization is often limited to studies using whole-cell INa or binding studies on neuronal membranes where there are mixed populations of NaV channel subtypes. For all of these toxins, the precise pattern of NaV channel subtype selectivity is either unknown or at best is incomplete.Recently, it was found that receptor site-3 was also recognized by a 43-residue spider toxin, originally named Magi 4, from the hexathelid spider Macrothele gigas (Iriomote, Japan). It binds with high affinity to insect NaV channels but, similar to scorpion α-like toxins, does not compete for the related site-3 in rat brain synaptosomes, despite being lethal by intracranial injection (10). Magi 4 shares significant homology to four δ-hexatoxin (HXTX)-1 family peptides and δ-actinopoditoxin-Mb1a (formerly δ-missulenatoxin-Mb1a; Fig. 1) but no sequence homology to scorpion α-toxins. Neurochemical studies have shown that δ-HXTX-1 toxins compete at nanomolar concentrations with both anti-mammalian (e.g. Aah2 and Lqh2) and anti-insect (e.g. LqhαIT) scorpion toxins for site-3 (1113). The three-dimensional structures of δ-HXTX-Ar1a and δ-HXTX-Hv1a peptides have been determined (14, 15) and possess core β regions stabilized by four disulfide bonds, placing them in the inhibitory cystine knot (ICK) structural family (16).Open in a separate windowFIGURE 1.Primary and secondary structure of δ-HXTX-Mg1a. A, comparison of the primary sequence of δ-HXTX-Mg1a and δ-HXTX-Mg1b (formerly Magi 14) with currently known members of the δ-HXTX-1 family and δ-AOTX-Mb1a (δ-actinopoditoxin-Mb1a, formerly δ-missulenatoxin-Mb1a). Homologies are shown relative to δ-HXTX-Mg1a; identities are boxed in gray, and conservative substitutions are in gray italic text. Gaps (dashes) have been inserted to maximize alignment. The disulfide bonding pattern for the strictly conserved cysteine residues determined for δ-HXTX-Mg1a (this study), δ-HXTX-Ar1a (55), and δ-HXTX-Hv1a (15) is indicated above the sequences; it is assumed that δ-AOTX-Mb1a (36), δ-HXTX-Hs20.1a (8), and δ-HXTX-Hv1b (56) have the same disulfide bonding pattern. The percentage identity and homology with δ-HXTX-Mg1a is shown to the right of the sequences. B, summary of δ-HXTX-Mg1a NMR data. Sequential NOEs, classified as very weak, weak, medium, and strong, are represented by the thickness of bars. Filled diamonds indicate backbone amide protons that form hydrogen bonds. 3JNHCα coupling constants are indicated by ↑ (>8 Hz) and ↓ (<5.5 Hz). Secondary structure is shown at the bottom of the figure where rectangles represent β-turns (the type of turn is indicated in the rectangle) and arrows represent β-sheets.The aim of this study was to first determine the solution structure of Magi 4 and second to investigate the ability of Magi 4 to discriminate between different NaV channels subtypes. Here we report the tertiary structure of Magi 4 by 1H NMR and show its disulfide bonding pattern and three-dimensional structure are homologous to δ-HXTX-1 toxins. We highlight the key residues in Magi 4 that appear to be topologically similar to those residues known to be part of the pharmacophore for site-3 scorpion α-toxins, despite Magi 4 having a different overall structure to scorpion α-toxins (11). In addition, we provide a detailed characterization of the selectivity and mode of action of Magi 4 on nine cloned mammalian and insect NaV channel subtypes, including a detailed characterization on insect neurotransmission. Given that the toxin potently slows the inactivation of NaV channels, it should be renamed δ-hexatoxin-Mg1a (δ-HXTX-Mg1a) in accordance with the rational nomenclature recently proposed for naming spider peptide toxins (7) (see ArachnoServer spider toxin data base).  相似文献   

17.
Rat C6 glioma cells were cultured for 4 days in MEM medium supplemented with 10% bovine serum and Na+,K+-ATPase activity was determined in homogenates of harvested cells. Approximately 50% of enzyme activity was attained at 1.5 mM K+ and the maximum (2.76±0.13 mol Pi/h/mg protein) at 5 mM K+. The specific activity of Na+,K+-ATPase was not influenced by freezing the homogenates or cell suspensions before the enzyme assay. Ten minutes' exposure of glioma cells to 10–4 or 10–5 M noradrenaline (NA) remained without any effect on NA+,K+-ATPase activity. Neither did the presence of NA in the incubation medium, during the enzyme assay, influence the enzyme activity. The nonresponsiveness of Na+,K+-ATPase of C6 glioma cells to NA is consistent with the assumption that (+) form of the enzyme may be preferentially sensitive to noradrenaline. Na+,K+-ATPase was inhibited in a dose-dependent manner by vanadate and 50% inhibition was achieved at 2×10–7 M concentration. In spite of the fact that Na+,K+-ATPase of glioma cells was not responsive to NA, the latter could at least partially reverse vanadate-induced inhibition of the enzyme. Although the present results concern transformed glial cells, they suggest the possibility that inhibition of glial Na+,K+-ATPase may contribute to the previously reported inhibition by vanadate of Na+,K+-ATPase of the whole brain tissue.  相似文献   

18.
19.
ATP synthase uses a unique rotational mechanism to convert chemical energy into mechanical energy and back into chemical energy. The helix-turn-helix motif, termed “DELSEED-loop,” in the C-terminal domain of the β subunit was suggested to be involved in coupling between catalysis and rotation. Here, the role of the DELSEED-loop was investigated by functional analysis of mutants of Bacillus PS3 ATP synthase that had 3–7 amino acids within the loop deleted. All mutants were able to catalyze ATP hydrolysis, some at rates several times higher than the wild-type enzyme. In most cases ATP hydrolysis in membrane vesicles generated a transmembrane proton gradient, indicating that hydrolysis occurred via the normal rotational mechanism. Except for two mutants that showed low activity and low abundance in the membrane preparations, the deletion mutants were able to catalyze ATP synthesis. In general, the mutants seemed less well coupled than the wild-type enzyme, to a varying degree. Arrhenius analysis demonstrated that in the mutants fewer bonds had to be rearranged during the rate-limiting catalytic step; the extent of this effect was dependent on the size of the deletion. The results support the idea of a significant involvement of the DELSEED-loop in mechanochemical coupling in ATP synthase. In addition, for two deletion mutants it was possible to prepare an α3β3γ subcomplex and measure nucleotide binding to the catalytic sites. Interestingly, both mutants showed a severely reduced affinity for MgATP at the high affinity site.F1F0-ATP synthase catalyzes the final step of oxidative phosphorylation and photophosphorylation, the synthesis of ATP from ADP and inorganic phosphate. F1F0-ATP synthase consists of the membrane-embedded F0 subcomplex, with, in most bacteria, a subunit composition of ab2c10, and the peripheral F1 subcomplex, with a subunit composition of α3β3γδε. The energy necessary for ATP synthesis is derived from an electrochemical transmembrane proton (or, in some organisms, a sodium ion) gradient. Proton flow down the gradient through F0 is coupled to ATP synthesis on F1 by a unique rotary mechanism. The protons flow through (half) channels at the interface of the a and c subunits, which drives rotation of the ring of c subunits. The c10 ring, together with F1 subunits γ and ε, forms the rotor. Rotation of γ leads to conformational changes in the catalytic nucleotide binding sites on the β subunits, where ADP and Pi are bound. The conformational changes result in the formation and release of ATP. Thus, ATP synthase converts electrochemical energy, the proton gradient, into mechanical energy in the form of subunit rotation and back into chemical energy as ATP. In bacteria, under certain physiological conditions, the process runs in reverse. ATP is hydrolyzed to generate a transmembrane proton gradient, which the bacterium requires for such functions as nutrient import and locomotion (for reviews, see Refs. 16).F1 (or F1-ATPase) has three catalytic nucleotide binding sites located on the β subunits at the interface to the adjacent α subunit. The catalytic sites have pronounced differences in their nucleotide binding affinity. During rotational catalysis, the sites switch their affinities in a synchronized manner; the position of γ determines which catalytic site is the high affinity site (Kd1 in the nanomolar range), which site is the medium affinity site (Kd2 ≈ 1 μm), and which site is the low affinity site (Kd3 ≈ 30–100 μm; see Refs. 7 and 8). In the original crystal structure of bovine mitochondrial F1 (9), one of the three catalytic sites, was filled with the ATP analog AMP-PNP,2 a second was filled with ADP (plus azide) (see Ref. 10), and the third site was empty. Hence, the β subunits are referred to as βTP, βDP, and βE. The occupied β subunits, βTP and βDP, were in a closed conformation, and the empty βE subunit was in an open conformation. The main difference between these two conformations is found in the C-terminal domain. Here, the “DELSEED-loop,” a helix-turn-helix structure containing the conserved DELSEED motif, is in an “up” position when the catalytic site on the respective β subunit is filled with nucleotide and in a “down” position when the site is empty (Fig. 1A). When all three catalytic sites are occupied by nucleotide, the previously open βE subunit assumes an intermediate, half-closed (βHC) conformation. It cannot close completely because of steric clashes with γ (11).Open in a separate windowFIGURE 1.The βDELSEED-loop. A, interaction of the βTP and βE subunits with theγ subunit.β subunits are shown in yellow andγ in blue. The DELSEED-loop (shown in orange, with the DELSEED motif itself in green)of βTP interacts with the C-terminal helixγ and the short helix that runs nearly perpendicular to the rotation axis. The DELSEED-loop of βE makes contact with the convex portion of γ, formed mainly by the N-terminal helix. A nucleotide molecule (shown in stick representation) occupies the catalytic site of βTP, and the subunit is in the closed conformation. The catalytic site on βE is empty, and the subunit is in the open conformation. This figure is based on Protein Data Bank file 1e79 (32). B, deletions in the βDELSEED-loop. The loop was “mutated” in silico to represent the PS3 ATP synthase. The 3–4-residue segments that are removed in the deletion mutants are color-coded as follows: 380LQDI383, pink; 384IAIL387, green; 388GMDE391, yellow; 392LSD394, cyan; 395EDKL398, orange; 399VVHR402, blue. Residues that are the most involved in contacts with γ are labeled. All figures were generated using the program PyMOL (DeLano Scientific, San Carlos, CA).The DELSEED-loop of each of the three β subunits makes contact with the γ subunit. In some cases, these contacts consist of hydrogen bonds or salt bridges between the negatively charged residues of the DELSEED motif and positively charged residues on γ. The interactions of the DELSEED-loop with γ, its movement during catalysis, the conservation of the DELSEED motif (see 1214). Thus, the finding that an AALSAAA mutant in the α3β3γ complex of ATP synthase from the thermophilic Bacillus PS3, where several hydrogen bonds/salt bridges to γ are removed simultaneously, could drive rotation of γ with the same torque as the wild-type enzyme (14) came as a surprise. On the other hand, it seems possible that it is the bulk of the DELSEED-loop, more so than individual interactions, that drives rotation of γ. According to a model favored by several authors (6, 15, 16) (see also Refs. 1719), binding of ATP (or, more precisely, MgATP) to the low affinity catalytic site on βE and the subsequent closure of this site, accompanied by its conversion into the high affinity site, are responsible for driving the large (80–90°) rotation substep during ATP hydrolysis, with the DELSEED-loop acting as a “pushrod.” A recent molecular dynamics (20) study supports this model and implicates mainly the region around several hydrophobic residues upstream of the DELSEED motif (specifically βI386 and βL387)3 as being responsible for making contact with γ during the large rotation substep.

TABLE 1

Conservation of residues in the DELSEED-loop Amino acids found in selected species in the turn region of the DELSEED-loop. Listed are all positions subjected to deletions in the present study. Residue numbers refer to the PS3 enzyme. Consensus annotation: p, polar residue; s, small residue; h, hydrophobic residue; –, negatively charged residue; +, positively charged residue.Open in a separate windowIn the present study, we investigated the function of the DELSEED-loop using an approach less focused on individual residues, by deleting stretches of 3–7 amino acids between positions β380 and β402 of ATP synthase from the thermophilic Bacillus PS3. We analyzed the functional properties of the deletion mutants after expression in Escherichia coli. The mutants showed ATPase activities, which were in some cases surprisingly high, severalfold higher than the activity of the wild-type control. On the other hand, in all cases where ATP synthesis could be measured, the rates where below or equal to those of the wild-type enzyme. In Arrhenius plots, the hydrolysis rates of the mutants were less temperature-dependent than those of wild-type ATP synthase. In those cases where nucleotide binding to the catalytic sites could be tested, the deletion mutants had a much reduced affinity for MgATP at high affinity site 1. The functional role of the DELSEED-loop will be discussed in light of the new information.  相似文献   

20.
Na+/H+ antiporters influence proton or sodium motive force across the membrane. Synechocystis sp. PCC 6803 has six genes encoding Na+/H+ antiporters, nhaS1–5 and sll0556. In this study, the function of NhaS3 was examined. NhaS3 was essential for growth of Synechocystis, and loss of nhaS3 was not complemented by expression of the Escherichia coli Na+/H+ antiporter NhaA. Membrane fractionation followed by immunoblotting as well as immunogold labeling revealed that NhaS3 was localized in the thylakoid membrane of Synechocystis. NhaS3 was shown to be functional over a pH range from pH 6.5 to 9.0 when expressed in E. coli. A reduction in the copy number of nhaS3 in the Synechocystis genome rendered the cells more sensitive to high Na+ concentrations. NhaS3 had no K+/H+ exchange activity itself but enhanced K+ uptake from the medium when expressed in an E. coli potassium uptake mutant. Expression of nhaS3 increased after shifting from low CO2 to high CO2 conditions. Expression of nhaS3 was also found to be controlled by the circadian rhythm. Gene expression peaked at the beginning of subjective night. This coincided with the time of the lowest rate of CO2 consumption caused by the ceasing of O2-evolving photosynthesis. This is the first report of a Na+/H+ antiporter localized in thylakoid membrane. Our results suggested a role of NhaS3 in the maintenance of ion homeostasis of H+, Na+, and K+ in supporting the conversion of photosynthetic products and in the supply of energy in the dark.Na+/H+ antiporters are integral membrane proteins that transport Na+ and H+ in opposite directions across the membrane and that occur in virtually all cell types. These transporters play an important role in the regulation of cytosolic pH and Na+ concentrations and influence proton or sodium motive force across the membrane (1, 2). In Escherichia coli, three Na+/H+ antiporters (NhaA, NhaB, and ChaA) have been described in detail. Of these, NhaA is the functionally best characterized transporter. The crystal structure of NhaA has been resolved (3). In addition, mutants of nhaA, nhaB, and chaA as well as the triple mutant have been generated (4). The triple mutant was shown to be hypersensitive to extracellular Na+. The genome of the cyanobacterium Synechocystis sp. PCC 6803 contains six genes encoding Na+/H+ antiporters (NhaS1–5 and sll0556). NhaS1 (slr1727) has also been designated SynNhaP (5, 6). Null mutants of nhaS1, nhaS2, nhaS4, and nhaS5 have been generated; however, a null mutant of nhaS3 could not be obtained, indicating that it is an essential gene (68). By heterologous expression in E. coli, Na+/H+ exchange activities could be shown for NhaS1–5 (5, 6). Inactivation of nhaS1 and nhaS2 results in retardation of growth of Synechocystis (5, 6). It has been reported that in these mutants the concentration of Na+ in cytosol and intrathylakoid space (lumen) increases and impairs the photosynthetic and/or respiratory activity of the cell (9, 10). Therefore the Na+ extrusion by Synechocystis Na+/H+ antiporters similar to E. coli NhaA, NhaB, and ChaA is essential for the adaptation to salinity stress.In contrast to the case in E. coli, Na+ is an essential element for the growth of some cyanobacteria (11, 12). Interestingly, the Na+/H+ antiporter homolog NhaS4 was identified as an uptake system for Na+ from the medium during a screen for mutations in Synechocystis that result in lack of growth at low Na+ concentrations (7). The requirement of a Na+ uptake antiporter for cell growth is consistent with the physiology of Synechocystis. Specifically, photoautotrophic bacteria like cyanobacteria share some components (plastoquinone, cytochrome b6f, and c6) of the thylakoid membrane for electron transport for both photophosphorylation and respiratory oxidative phosphorylation. Na+/H+ antiporters therefore may coordinate both H+ and Na+ gradients across the plasma and thylakoid membranes to adapt to daily environmental changes (11). It remains to be determined whether the six Na+/H+ antiporters are localized to the plasma membrane or to the thylakoid membrane in Synechocystis. Information on the membrane localization will also provide information on the physiological role in Synechocystis. In this study, we explored the membrane localization of NhaS3, the role of specific amino acid residues for its function, and the effect of CO2 concentration and circadian rhythms on the expression pattern of nhaS3 to gain insight into the physiological role of NhaS3 in Synechocystis.  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