Comprehensive Physiology Wiley Online Library

Modulation of Voltage‐Gated Ion Channels by Sialylation

Full Article on Wiley Online Library



Abstract

Control and modulation of electrical signaling is vital to normal physiology, particularly in neurons, cardiac myocytes, and skeletal muscle. The orchestrated activities of variable sets of ion channels and transporters, including voltage‐gated ion channels (VGICs), are responsible for initiation, conduction, and termination of the action potential (AP) in excitable cells. Slight changes in VGIC activity can lead to severe pathologies including arrhythmias, epilepsies, and paralyses, while normal excitability depends on the precise tuning of the AP waveform. VGICs are heavily posttranslationally modified, with upward of 30% of the mature channel mass consisting of N‐ and O‐glycans. These glycans are terminated typically by negatively charged sialic acid residues that modulate voltage‐dependent channel gating directly. The data indicate that sialic acids alter VGIC activity in isoform‐specific manners, dependent in part, on the number/location of channel sialic acids attached to the pore‐forming alpha and/or auxiliary subunits that often act through saturating electrostatic mechanisms. Additionally, cell‐specific regulation of sialylation can affect VGIC gating distinctly. Thus, channel sialylation is likely regulated through two mechanisms that together contribute to a dynamic spectrum of possible gating motifs: a subunit‐specific mechanism and regulated (aberrant) changes in the ability of the cell to glycosylate. Recent studies showed that neuronal and cardiac excitability is modulated through regulated changes in voltage‐gated Na+ channel sialylation, suggesting that both mechanisms of differential VGIC sialylation contribute to electrical signaling in the brain and heart. Together, the data provide insight into an important and novel paradigm involved in the control and modulation of electrical signaling. © 2012 American Physiological Society. Compr Physiol 2:1269‐1301, 2012.

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

Download a PowerPoint presentation of all images


Figure 1. Figure 1.

External [Ca2+] shifts voltage‐gated ion channel (VGIC) activation voltage. Decreased [Cao2+] results in a 10‐ to 30‐mV shift in squid giant axon Na+ current activation voltage. Adapted, with permission, from B Frankenhaeuser and AL Hodgkin. J Physiol 137: 218‐244, 1957 59.

Reprinted with permission.
Figure 2. Figure 2.

External negative surface charge impacts membrane potential sensed by VGIC. Illustration of the putative effects of external negative charges on the membrane electric field and potential. Surface potential theory predicts that an increase in external divalent cations nullifies the impact of the surface charge on the membrane potential. Figure adapted, with permission, from B Hille. Ion channels of Excitable Membranes. 3rd edition 84.

Reprinted, with permission, from Sinauer Associates.
Figure 3. Figure 3.

Extracellular glycosylation is a sequential process that occurs in the endoplasmic reticulum (ER) and Golgi and involves the activity of hundreds of glycogenes. Illustration of the process of protein N‐glycosylation as it occurs in the ER and Golgi. Figure adapted, with permission, from A Helenius and M Aebi. Science 291: 2364‐2369, 2001 79.

Reprinted, with permission, from AAAS.
Figure 4. Figure 4.

Voltage‐dependent gating of the skeletal muscle Nav was shifted nearly identically under each of three independent conditions of reduced sialylation. Conductance‐voltage (G‐V) relationships for Nav1.4 expressed under conditions of full sialylation (+SA) and three independent conditions of reduced sialylation [−SA; neuraminidase‐treated, to remove sialic acids (top), in the essentially nonsialylating Lec2 cell line (middle), and two deletion mutants in which four and five N‐glycosylation sites were removed through mutagenesis (bottom)]. G‐V relationships here and generally, were determined by measuring the peak conductance during a brief depolarizing test pulse from a negative resting potential. Adapted, with permission, from E Bennett et al. J Gen Physiol 109: 327‐343, 1997 17.

Figure 5. Figure 5.

Gating of less‐sialylated Nav are not as sensitive to changes in [Cao2+]. Conductance‐voltage (G‐V) curves for Nav1.4 expressed under conditions of full sialylation (top) and reduced sialylation (bottom) at 2.0 (filled circles) and 0.2 (open circles) mmol/L Ca2+. Note the much larger hyperpolarizing shifts in G‐V relationship under reduced [Cao2+] for the fully sialylated channels. Figure adapted, with permission, from E Bennett et al. J Gen Physiol 109: 327‐343, 1997 17.

Figure 6. Figure 6.

Functional channel sialic acids are localized to the DIS5‐S6 loop. Conductance‐voltage (G‐V) curves for Nav1.4 (hSkM1, top left), Nav1.5 (hH1, top right), and chimeras in which the DIS5‐S6 loop of Nav1.5 replaced the Nav1.4 loop (hSkM1P1, bottom left) or vice versa (hH1P1, bottom right) as expressed in fully sialylating Pro5 (circles) and nonsialylating Lec2 (squares) cells. Figure adapted, with permission, from ES Bennett. J Physiol 538: 675‐690, 2002 18.

Figure 7. Figure 7.

The fully sialylated β1‐subunit modulates gating of three Nav α subunits. Conductance‐voltage (G‐V) curves for four Nav α‐subunits ± the β1‐subunit under conditions of full and reduced sialylation. Filled symbols: α‐subunit alone. Open symbols: α + β1. Circles: in Pro5 cells (+SA). Squares: in Lec2 cells (−SA). Solid lines: α‐subunit alone. Dashed lines: α + β1. A, Nav1.4. B, Nav1.5. C, Nav1.7. D, Nav1.2. Figure adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.

Figure 8. Figure 8.

The β1‐subunit had no effect on Nav α‐subunit gating following removal of β1 N‐glycosylation sites. Conductance‐voltage (G‐V) relationships for hSkM1P1 (A), Nav1.5 (B), Nav1.7 (C), and Nav1.2 (D) under fully sialylated conditions alone (filled circles with solid lines; n = 9‐13 for each) or coexpressed with β1 (filled squares with dashed lines; n = 9‐12 for each) or with β1‐ΔN, a mutant β1 in which all four putative N‐linked glycosylation Asparagine residues were mutated to Serine residues (filled triangles with dotted lines; n = 4‐6 for each). Removal of β1 N‐linked sugars prevented fully the β1 sialic acid‐induced hyperpolarizing shifts in channel gating. Filled triangles with dotted line, β1‐ΔN. All other symbols and lines are as described in Figure 7. Figure and legend adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.

Figure 9. Figure 9.

The impact of channel sialic acids on Nav gating is apparently a saturating phenomenon. Voltage‐dependent steady state and kinetic gating for hSkM1P1 ± β1 ± SA. Circles, hSkM1P1 alone; squares, hSkM1P1 ± β1. Filled symbols, in Pro5 cells; open symbols, in Lec2 cells. n = 8‐11 for each condition. A schematic of hSkM1P1 structure illustrates that the chimera consists of Nav1.4 with the less‐glycosylated Nav1.5 DIS5‐S6 loop replacing the Nav1.4 DIS5‐S6. (A) Conductance‐voltage (G‐V) relationship. (B) Steady‐state channel availability curve (SSI). Peak currents were measured during a short, maximally depolarizing pulse that was preceded by a 500 ms prepulse to increasingly depolarized potentials. (C) Fast inactivation time constants. Fractional current that recovered from fast inactivation during recovery prepulses of variable duration was used to determine time constants. (D) Time constants for recovery from fast inactivation. Figure and legend adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.

Figure 10. Figure 10.

Model predicting the putative saturating effects of α and β1 sialic acids on Nav gating. (A) Examples of possible interactions of β1 sialic acids with Nav1.4 (left) and the glycosylation‐deficient chimera, hSkM1P1 (right). The data suggest that β1 sialic acids cannot contribute further to the gating of Nav1.4 but do contribute to the gating of the other α‐subunits through an apparent electrostatic mechanism. Thus, the left panel shows ineffectual β1 sialic acids as distant from Nav1.4, whereas the right panel illustrates that fewer α‐subunit functional DIS5‐S6 sialic acids may allow β1 sialic acids to interact more intimately with the α‐subunit and contribute to channel gating. (B) A theoretical concentration response curve comparing relative contributions to Va by functional sialic acids associated with various αβ1 combinations. A typical sigmoidal curve is shown, with the various αβ1‐subunit combinations aligned along the curve; the location of each is not precise but is consistent, relatively, with the data shown in Figures 7‐9. For example, Nav1.4 must contain essentially saturating levels of functional sialic acids in this experimental system. Thus, Nav1.4 alone and Nav1.4 ± β1 are pictured along the maximum of the curve. In a second example, Nav1.5 alone cannot contain many DIS5‐S6 functional sialic acids, and therefore Nav1.5 expressed alone must lie somewhere along the minimum portions of the curve. When β1 is coexpressed with Nav1.5, the levels of functional sialic acids increase, causing a hyperpolarization in Va, and hence, Nav1.5 + β1 is shown somewhere along the rising slope of the curve. Although the data presented in Johnson et al. J Biol Chem, 2004, showed uniform effects of β1 sialic acids on measured voltage‐dependent Nav gating parameters, there is no reason, a priori, for this to occur. The impact of β1 sialic acids on specific Nav gating parameters will depend on several factors including the extent of coupling among channel kinetic states and/or any inherent voltage dependence of the transition from one state to another. Thus, for example, the interaction of β1 sialic acids with a specific α‐subunit may place the channel at a different position along the concentration response curves for shifts in Va versus shifts in Vi. This would likely be observed as differences in the impact of β1 sialic acids on Va and Vi. “P1” represents hSkM1P1. Figure and legend adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.

Figure 11. Figure 11.

The sialic acid‐dependent effects of β1 and β2 on Nav1.5 gating are additive. Filled symbols, Pro5 cells (+SA); open symbols, Lec2 cells (−SA); gold circles/lines, Nav1.5; red squares/lines, Nav1.5 + β2; cyan diamonds/lines, Nav1.5 + β1; purple triangles/lines, Nav1.5 + β1 + β2. n = 9‐12 for each condition tested. (A) Conductance‐voltage (G‐V) relationships. (B) Channel availability (SSI) curves. (C) Fast inactivation time constant. (D) Time constants of the recovery from fast inactivation. Figure and legend adapted, with permission, from D Johnson and ES Bennett. J Biol Chem 281: 25875‐25881, 2006 105.

Figure 12. Figure 12.

Sialic acids modulate Kv1.1 gating through apparent electrostatic mechanisms. Bar graphs illustrating the impact of changes in [Cao2+] on activation gating of Kv1.1 under three conditions of glycosylation (full, in Pro5; reduced sialylation, in Lec2; terminal galactose and sialic acid residues missing, in Lec8).

Figure adapted, with permission, from WB Thornhill et al. J Biol Chem 271: 19093‐19098, 1996 194. Reprinted with permission.
Figure 13. Figure 13.

Kv1.2 gating can be modulated by the addition or removal of N‐glycans. Conductance‐voltage (G‐V) curves for wild‐type Kv1.2 (contains one N‐glycosylation site) and two mutant Kv1.2 constructs in which N‐glycosylation sites were added (Kv1.2Tri‐N) or removed (Kv1.2N207Q) through mutagenesis.

Figure adapted, with permission, from I Watanabe et al. Brain Res 1144: 1‐18, 2007 211. Reprinted with permission.
Figure 14. Figure 14.

Cardiac transient outward K+ current, Ito, is sensitive to desialylation. (A) Whole cell IK traces from adult ventricular myocytes ± neuraminidase treatment. (B) Peak Ito density ± neuraminidase. (C) Conductance‐voltage (G‐V) relationships for Ito ± neuraminidase. (D) Steady‐state channel availability (SSI) curves for Ito ± neuraminidase. Figure adapted, with permission, from CA Ufret‐Vincenty et al. Am J Physiol Cell Physiol 281: C464‐C474, 2001b 200.

Figure 15. Figure 15.

Adult dorsal root ganglia (DRG) neuronal Nav is less sialylated and less sensitive to neuraminidase treatment than the neonatal Nav. Comparison of INa in neonatal and adult DRG neurons. (A) Current‐voltage (I‐V) relationships. (B) Conductance‐voltage (G‐V) relationships. (C) Steady‐state channel availability (SSI) curves.

Adapted, with permission, from L Tyrrell et al. J Neuroscience 21: 24 9629‐9637, 2001 198. Reprinted with permission.
Figure 16. Figure 16.

Neonatal ventricular Nav gate at more depolarized potentials. (A) Conductance‐voltage (G‐V) relationships. Inset: whole cell Na+ current (INa) from an isolated adult ventricular myocyte. (B) Steady‐state inactivation (SSI). Inset: INa measured during a series of +40 mV test pulses, each following a prepulse to a larger depolarization. (C) Time constant of fast inactivation (τh). Inset: INa during a −50 mV test pulse. Solid traces: neonatal atrial (NA), adult atrial (AA), and adult ventricular (NV) myocytes. Dotted trace: neonatal ventricular (NV) myocyte. (D) Voltage dependence of the recovery time constants. Inset: plotted recovery data and fits for all four conditions at a −130 mV recovery potential. Solid and dotted curves as described in (C). n = 6‐17. Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.

Figure 17. Figure 17.

Removal of Nav1.5 glycosylation and sialylation can account fully for gating difference in Nav1.5 among neonatal and adult atrial and ventricular Nav. Conductance‐voltage (G‐V) (A), and steady‐state inactivation (SSI) (B) curves, fast inactivation (C), and recovery (D) kinetics for all channel types following treatment with PNGase‐F (open symbols, triangles, NA; inverted triangles, AA; circles, NV; squares, AV), and for Nav from neonatal ventricles under control/untreated (solid circles) and neuraminidase‐treated (dotted, open circles) conditions. Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.

Figure 18. Figure 18.

Gel‐shift analyses suggest that Nav1.5 sialylation levels are regulated between cardiac chambers and during ventricular development. (A‐C) Immunoblots of Nav α subunits from neonatal and adult atria and ventricles ± glycosidase treatments. (A) Untreated lysates. (B) Lysates ± PNGase‐F treatment. (C) Lysates ± neuraminidase treatment. (D) Bar graph of average measured Nav α‐subunit Mr ± PNGase‐F ± neuraminidase. Significance was determined for the untreated data using a two‐tailed Student's t test comparing the Mr for the α‐subunit from neonatal ventricles to the Mr for the α‐subunits from neonatal atria and adult atria and ventricles. *, significant (P < 0.01). Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.

Figure 19. Figure 19.

Suggested mechanisms by which differential sialylation modulates voltage‐gated ion channel (VGIC) gating. Simplified model depicts subunit‐ and cell‐specific differential sialylation relevant to Now gating. Each mechanism potentially can produce a modest spectrum of channel gating motifs [shown for illustrative purposes only as different conductance‐voltage (G‐V) relationships]. Top left: subunit‐specific differential sialylation. Each α‐ (and β‐) subunit will have an inherent “glycosylation signature” that predetermines the level/location of glycosylation that can be attached to channel. Top right: cell‐specific differential sialylation. The relative activity of one to several of the 20 sialyltransferases (STs) will determine the amount and location of sialic acids added to one α‐ (and/or β‐) subunit. When combined together (bottom), perhaps hundreds of gating motifs could be realized. For example, by assuming that each of the Nav α‐subunits might be sialylated differently by the activity of each of the 20 known STs, 200 different gating motifs might be possible. This model is likely oversimplified, as it does not account for the even greater number of gating motifs that might be realized through regulated expression of the Nav β‐subunits, the activity of multiple STs in various combinations, or for the varied expression/activity of other glycogenes. Thus, with the regulated expression of specific α‐ and β‐subunits combined with the regulated expression and activity of specific STs, Nav gating might be modulated across a large and finely tuned spectrum of voltages. Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.

Figure 20. Figure 20.

Mutant polysialyltransferase modulates activity of Drosophila Nav. (A) Tetrodotoxin (TTX) sensitivity of the Drosophila excitatory junction potentials (EJP) is reduced in mutant DSiaT (S23), suggesting reduced Para (Drosophila Nav) activity in the mutant. (B) The TTX sensitivity of the EJP is rescued by the DSiaT rescue mutant.

Figure adapted, with permission, from E Repnikova et al. J Neuroscience 30: 186466‐6476, 2010 162. Reprinted with permission.
Figure 21. Figure 21.

Hippocampal Nav gating is modulated by sialylation. Nucleated patch (A and B) and whole cell recording (C and D) measurements of isolated hippocampal neuronal Nav conductance‐voltage (G‐V) (A and C) and SSI (B and D) ± neuraminidase treatment. In (C) and (D), G‐V and SSI curves measured following treatment with neuraminidase + inhibitor are essentially the same as the untreated curves.

Figure adapted, with permission, from D Isaev et al. J Neuroscience 27: 4311587‐11594, 2007 96. Reprinted with permission.
Figure 22. Figure 22.

Hippocampal excitability is modulated by apparent changes in hippocampal protein sialylation. (A) Field recordings ± neuraminidase ± neuraminidase blocker following multiple injections of higher [Ko+] used to induce epileptiform activity in vivo. (B) Summary data of seizure occurrence as a function of injection number. (C) Duration of ictal‐like activity. (D) Maximum population spike frequency.

Figure adapted, with permission, from D Isaev et al. J Neuroscience 27: 4311587‐11594, 2007 96. Reprinted with permission.
Figure 23. Figure 23.

Atrial action potential (AP) waveform is modulated by expression of the polysialyltransferase, ST8Sia2. APs were measured from isolated control and ST8sia2(−/−) neonatal atrial and ventricular myocytes. Top left: typical AP waveforms measured from control (in black) and ST8sia2(−/−) (in red) atrial myocytes. (A‐C) The mean ± SEM AP waveform parameters (relevant portions of AP marked with arrows). Black, control atrium (n = 7); red, ST8sia2 (−/−) atrium (n = 9); dark blue, control ventricle (n = 7); blue, ST8sia2(−/−) ventricle (n = 5). (A) Time to AP peak. (B) AP duration at 50% repolarization (APD50). (C) AP duration at 90% repolarization (APD90). Significance tested using a two‐tailed t test and comparing ST8sia2(−/−) to control parameters. *, significant (P<0.005); #, not significant (P>0.1). Figure and legend adapted, with permission, from ML Montpetit et al. Proc Natl Acad Sci, U S A 106: 16517‐16522, 2009 142.

Figure 24. Figure 24.

Atrial Nav gating is modulated by ST8Sia2 expression. ST8sia2(−/−) and control myocyte Nav gating. Left panels (A, C, E, and G): atrium, ST8sia2(−/−) (red; n = 10); control (black; n = 4). Right panels (B, D, F, and H): ventricle, ST8sia2(−/−) (blue; n = 4); control (dark blue; n = 6). (A and B) Steady‐state activation. Insets: half‐activation voltage. (C and D) Steady‐state channel availability (SSI). Dashed line: ST8sia2(−/−) data shifted along voltage axis by −7.3 mV (to mimic measured shift in half‐activation voltage). Insets: half‐inactivation voltage. (E and F) Fast inactivation time constants. Dashed line: The ST8sia2(−/−) data shifted along voltage axis by −7.3 mV. Insets: typical whole cell Na+ current traces elicited by stepping to a −40 mV test potential. (G and H) Time constants of recovery from fast inactivation at a −130 mV recovery potential. Significance tested using a two‐tailed t test and comparing ST8sia2(−/−) to control parameters. *, significant (P ≤ 0.005); #, not significant (P > 0.1). Figure and legend adapted, with permission, from ML Montpetit et al. Proc Natl Acad Sci, U S A 106: 16517‐16522, 2009 142.

Figure 25. Figure 25.

MLP−/− ventricular myocytes are susceptible to early afterdepolarizations. MLP−/− mice show extended QT intervals (A), increased action potential durations (B), and increased susceptibility to early after depolarizations (C and D).

Figure adapted, with permission, from CA Ufret‐Vincenty et al. J Biol Chem 276: 28197‐28203, 2001 199. Reprinted with permission.
Figure 26. Figure 26.

Gating of MLP−/− ventricular Nav is less sensitive to desialylation than the wild‐type channel. INa for desialylated control myocytes mimicked untreated and neuraminidase‐treated MLP−/− myocyte INa. (A) Current‐voltage (I‐V) relationships. (B) Conductance‐voltage (G‐V) relationships. (D) SSI curves. (E) Fast inactivation times constants.

Figure adapted, with permission, from CA Ufret‐Vincenty et al. J Biol Chem 276: 28197‐28203, 2001 199. Reprinted with permission.


Figure 1.

External [Ca2+] shifts voltage‐gated ion channel (VGIC) activation voltage. Decreased [Cao2+] results in a 10‐ to 30‐mV shift in squid giant axon Na+ current activation voltage. Adapted, with permission, from B Frankenhaeuser and AL Hodgkin. J Physiol 137: 218‐244, 1957 59.

Reprinted with permission.


Figure 2.

External negative surface charge impacts membrane potential sensed by VGIC. Illustration of the putative effects of external negative charges on the membrane electric field and potential. Surface potential theory predicts that an increase in external divalent cations nullifies the impact of the surface charge on the membrane potential. Figure adapted, with permission, from B Hille. Ion channels of Excitable Membranes. 3rd edition 84.

Reprinted, with permission, from Sinauer Associates.


Figure 3.

Extracellular glycosylation is a sequential process that occurs in the endoplasmic reticulum (ER) and Golgi and involves the activity of hundreds of glycogenes. Illustration of the process of protein N‐glycosylation as it occurs in the ER and Golgi. Figure adapted, with permission, from A Helenius and M Aebi. Science 291: 2364‐2369, 2001 79.

Reprinted, with permission, from AAAS.


Figure 4.

Voltage‐dependent gating of the skeletal muscle Nav was shifted nearly identically under each of three independent conditions of reduced sialylation. Conductance‐voltage (G‐V) relationships for Nav1.4 expressed under conditions of full sialylation (+SA) and three independent conditions of reduced sialylation [−SA; neuraminidase‐treated, to remove sialic acids (top), in the essentially nonsialylating Lec2 cell line (middle), and two deletion mutants in which four and five N‐glycosylation sites were removed through mutagenesis (bottom)]. G‐V relationships here and generally, were determined by measuring the peak conductance during a brief depolarizing test pulse from a negative resting potential. Adapted, with permission, from E Bennett et al. J Gen Physiol 109: 327‐343, 1997 17.



Figure 5.

Gating of less‐sialylated Nav are not as sensitive to changes in [Cao2+]. Conductance‐voltage (G‐V) curves for Nav1.4 expressed under conditions of full sialylation (top) and reduced sialylation (bottom) at 2.0 (filled circles) and 0.2 (open circles) mmol/L Ca2+. Note the much larger hyperpolarizing shifts in G‐V relationship under reduced [Cao2+] for the fully sialylated channels. Figure adapted, with permission, from E Bennett et al. J Gen Physiol 109: 327‐343, 1997 17.



Figure 6.

Functional channel sialic acids are localized to the DIS5‐S6 loop. Conductance‐voltage (G‐V) curves for Nav1.4 (hSkM1, top left), Nav1.5 (hH1, top right), and chimeras in which the DIS5‐S6 loop of Nav1.5 replaced the Nav1.4 loop (hSkM1P1, bottom left) or vice versa (hH1P1, bottom right) as expressed in fully sialylating Pro5 (circles) and nonsialylating Lec2 (squares) cells. Figure adapted, with permission, from ES Bennett. J Physiol 538: 675‐690, 2002 18.



Figure 7.

The fully sialylated β1‐subunit modulates gating of three Nav α subunits. Conductance‐voltage (G‐V) curves for four Nav α‐subunits ± the β1‐subunit under conditions of full and reduced sialylation. Filled symbols: α‐subunit alone. Open symbols: α + β1. Circles: in Pro5 cells (+SA). Squares: in Lec2 cells (−SA). Solid lines: α‐subunit alone. Dashed lines: α + β1. A, Nav1.4. B, Nav1.5. C, Nav1.7. D, Nav1.2. Figure adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.



Figure 8.

The β1‐subunit had no effect on Nav α‐subunit gating following removal of β1 N‐glycosylation sites. Conductance‐voltage (G‐V) relationships for hSkM1P1 (A), Nav1.5 (B), Nav1.7 (C), and Nav1.2 (D) under fully sialylated conditions alone (filled circles with solid lines; n = 9‐13 for each) or coexpressed with β1 (filled squares with dashed lines; n = 9‐12 for each) or with β1‐ΔN, a mutant β1 in which all four putative N‐linked glycosylation Asparagine residues were mutated to Serine residues (filled triangles with dotted lines; n = 4‐6 for each). Removal of β1 N‐linked sugars prevented fully the β1 sialic acid‐induced hyperpolarizing shifts in channel gating. Filled triangles with dotted line, β1‐ΔN. All other symbols and lines are as described in Figure 7. Figure and legend adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.



Figure 9.

The impact of channel sialic acids on Nav gating is apparently a saturating phenomenon. Voltage‐dependent steady state and kinetic gating for hSkM1P1 ± β1 ± SA. Circles, hSkM1P1 alone; squares, hSkM1P1 ± β1. Filled symbols, in Pro5 cells; open symbols, in Lec2 cells. n = 8‐11 for each condition. A schematic of hSkM1P1 structure illustrates that the chimera consists of Nav1.4 with the less‐glycosylated Nav1.5 DIS5‐S6 loop replacing the Nav1.4 DIS5‐S6. (A) Conductance‐voltage (G‐V) relationship. (B) Steady‐state channel availability curve (SSI). Peak currents were measured during a short, maximally depolarizing pulse that was preceded by a 500 ms prepulse to increasingly depolarized potentials. (C) Fast inactivation time constants. Fractional current that recovered from fast inactivation during recovery prepulses of variable duration was used to determine time constants. (D) Time constants for recovery from fast inactivation. Figure and legend adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.



Figure 10.

Model predicting the putative saturating effects of α and β1 sialic acids on Nav gating. (A) Examples of possible interactions of β1 sialic acids with Nav1.4 (left) and the glycosylation‐deficient chimera, hSkM1P1 (right). The data suggest that β1 sialic acids cannot contribute further to the gating of Nav1.4 but do contribute to the gating of the other α‐subunits through an apparent electrostatic mechanism. Thus, the left panel shows ineffectual β1 sialic acids as distant from Nav1.4, whereas the right panel illustrates that fewer α‐subunit functional DIS5‐S6 sialic acids may allow β1 sialic acids to interact more intimately with the α‐subunit and contribute to channel gating. (B) A theoretical concentration response curve comparing relative contributions to Va by functional sialic acids associated with various αβ1 combinations. A typical sigmoidal curve is shown, with the various αβ1‐subunit combinations aligned along the curve; the location of each is not precise but is consistent, relatively, with the data shown in Figures 7‐9. For example, Nav1.4 must contain essentially saturating levels of functional sialic acids in this experimental system. Thus, Nav1.4 alone and Nav1.4 ± β1 are pictured along the maximum of the curve. In a second example, Nav1.5 alone cannot contain many DIS5‐S6 functional sialic acids, and therefore Nav1.5 expressed alone must lie somewhere along the minimum portions of the curve. When β1 is coexpressed with Nav1.5, the levels of functional sialic acids increase, causing a hyperpolarization in Va, and hence, Nav1.5 + β1 is shown somewhere along the rising slope of the curve. Although the data presented in Johnson et al. J Biol Chem, 2004, showed uniform effects of β1 sialic acids on measured voltage‐dependent Nav gating parameters, there is no reason, a priori, for this to occur. The impact of β1 sialic acids on specific Nav gating parameters will depend on several factors including the extent of coupling among channel kinetic states and/or any inherent voltage dependence of the transition from one state to another. Thus, for example, the interaction of β1 sialic acids with a specific α‐subunit may place the channel at a different position along the concentration response curves for shifts in Va versus shifts in Vi. This would likely be observed as differences in the impact of β1 sialic acids on Va and Vi. “P1” represents hSkM1P1. Figure and legend adapted, with permission, from D Johnson et al. J Biol Chem 279: 44303‐44310, 2004 106.



Figure 11.

The sialic acid‐dependent effects of β1 and β2 on Nav1.5 gating are additive. Filled symbols, Pro5 cells (+SA); open symbols, Lec2 cells (−SA); gold circles/lines, Nav1.5; red squares/lines, Nav1.5 + β2; cyan diamonds/lines, Nav1.5 + β1; purple triangles/lines, Nav1.5 + β1 + β2. n = 9‐12 for each condition tested. (A) Conductance‐voltage (G‐V) relationships. (B) Channel availability (SSI) curves. (C) Fast inactivation time constant. (D) Time constants of the recovery from fast inactivation. Figure and legend adapted, with permission, from D Johnson and ES Bennett. J Biol Chem 281: 25875‐25881, 2006 105.



Figure 12.

Sialic acids modulate Kv1.1 gating through apparent electrostatic mechanisms. Bar graphs illustrating the impact of changes in [Cao2+] on activation gating of Kv1.1 under three conditions of glycosylation (full, in Pro5; reduced sialylation, in Lec2; terminal galactose and sialic acid residues missing, in Lec8).

Figure adapted, with permission, from WB Thornhill et al. J Biol Chem 271: 19093‐19098, 1996 194. Reprinted with permission.


Figure 13.

Kv1.2 gating can be modulated by the addition or removal of N‐glycans. Conductance‐voltage (G‐V) curves for wild‐type Kv1.2 (contains one N‐glycosylation site) and two mutant Kv1.2 constructs in which N‐glycosylation sites were added (Kv1.2Tri‐N) or removed (Kv1.2N207Q) through mutagenesis.

Figure adapted, with permission, from I Watanabe et al. Brain Res 1144: 1‐18, 2007 211. Reprinted with permission.


Figure 14.

Cardiac transient outward K+ current, Ito, is sensitive to desialylation. (A) Whole cell IK traces from adult ventricular myocytes ± neuraminidase treatment. (B) Peak Ito density ± neuraminidase. (C) Conductance‐voltage (G‐V) relationships for Ito ± neuraminidase. (D) Steady‐state channel availability (SSI) curves for Ito ± neuraminidase. Figure adapted, with permission, from CA Ufret‐Vincenty et al. Am J Physiol Cell Physiol 281: C464‐C474, 2001b 200.



Figure 15.

Adult dorsal root ganglia (DRG) neuronal Nav is less sialylated and less sensitive to neuraminidase treatment than the neonatal Nav. Comparison of INa in neonatal and adult DRG neurons. (A) Current‐voltage (I‐V) relationships. (B) Conductance‐voltage (G‐V) relationships. (C) Steady‐state channel availability (SSI) curves.

Adapted, with permission, from L Tyrrell et al. J Neuroscience 21: 24 9629‐9637, 2001 198. Reprinted with permission.


Figure 16.

Neonatal ventricular Nav gate at more depolarized potentials. (A) Conductance‐voltage (G‐V) relationships. Inset: whole cell Na+ current (INa) from an isolated adult ventricular myocyte. (B) Steady‐state inactivation (SSI). Inset: INa measured during a series of +40 mV test pulses, each following a prepulse to a larger depolarization. (C) Time constant of fast inactivation (τh). Inset: INa during a −50 mV test pulse. Solid traces: neonatal atrial (NA), adult atrial (AA), and adult ventricular (NV) myocytes. Dotted trace: neonatal ventricular (NV) myocyte. (D) Voltage dependence of the recovery time constants. Inset: plotted recovery data and fits for all four conditions at a −130 mV recovery potential. Solid and dotted curves as described in (C). n = 6‐17. Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.



Figure 17.

Removal of Nav1.5 glycosylation and sialylation can account fully for gating difference in Nav1.5 among neonatal and adult atrial and ventricular Nav. Conductance‐voltage (G‐V) (A), and steady‐state inactivation (SSI) (B) curves, fast inactivation (C), and recovery (D) kinetics for all channel types following treatment with PNGase‐F (open symbols, triangles, NA; inverted triangles, AA; circles, NV; squares, AV), and for Nav from neonatal ventricles under control/untreated (solid circles) and neuraminidase‐treated (dotted, open circles) conditions. Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.



Figure 18.

Gel‐shift analyses suggest that Nav1.5 sialylation levels are regulated between cardiac chambers and during ventricular development. (A‐C) Immunoblots of Nav α subunits from neonatal and adult atria and ventricles ± glycosidase treatments. (A) Untreated lysates. (B) Lysates ± PNGase‐F treatment. (C) Lysates ± neuraminidase treatment. (D) Bar graph of average measured Nav α‐subunit Mr ± PNGase‐F ± neuraminidase. Significance was determined for the untreated data using a two‐tailed Student's t test comparing the Mr for the α‐subunit from neonatal ventricles to the Mr for the α‐subunits from neonatal atria and adult atria and ventricles. *, significant (P < 0.01). Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.



Figure 19.

Suggested mechanisms by which differential sialylation modulates voltage‐gated ion channel (VGIC) gating. Simplified model depicts subunit‐ and cell‐specific differential sialylation relevant to Now gating. Each mechanism potentially can produce a modest spectrum of channel gating motifs [shown for illustrative purposes only as different conductance‐voltage (G‐V) relationships]. Top left: subunit‐specific differential sialylation. Each α‐ (and β‐) subunit will have an inherent “glycosylation signature” that predetermines the level/location of glycosylation that can be attached to channel. Top right: cell‐specific differential sialylation. The relative activity of one to several of the 20 sialyltransferases (STs) will determine the amount and location of sialic acids added to one α‐ (and/or β‐) subunit. When combined together (bottom), perhaps hundreds of gating motifs could be realized. For example, by assuming that each of the Nav α‐subunits might be sialylated differently by the activity of each of the 20 known STs, 200 different gating motifs might be possible. This model is likely oversimplified, as it does not account for the even greater number of gating motifs that might be realized through regulated expression of the Nav β‐subunits, the activity of multiple STs in various combinations, or for the varied expression/activity of other glycogenes. Thus, with the regulated expression of specific α‐ and β‐subunits combined with the regulated expression and activity of specific STs, Nav gating might be modulated across a large and finely tuned spectrum of voltages. Figure and legend adapted, with permission, from PJ Stocker and ES Bennett. J Gen Physiol 127: 253‐265, 2006 180.



Figure 20.

Mutant polysialyltransferase modulates activity of Drosophila Nav. (A) Tetrodotoxin (TTX) sensitivity of the Drosophila excitatory junction potentials (EJP) is reduced in mutant DSiaT (S23), suggesting reduced Para (Drosophila Nav) activity in the mutant. (B) The TTX sensitivity of the EJP is rescued by the DSiaT rescue mutant.

Figure adapted, with permission, from E Repnikova et al. J Neuroscience 30: 186466‐6476, 2010 162. Reprinted with permission.


Figure 21.

Hippocampal Nav gating is modulated by sialylation. Nucleated patch (A and B) and whole cell recording (C and D) measurements of isolated hippocampal neuronal Nav conductance‐voltage (G‐V) (A and C) and SSI (B and D) ± neuraminidase treatment. In (C) and (D), G‐V and SSI curves measured following treatment with neuraminidase + inhibitor are essentially the same as the untreated curves.

Figure adapted, with permission, from D Isaev et al. J Neuroscience 27: 4311587‐11594, 2007 96. Reprinted with permission.


Figure 22.

Hippocampal excitability is modulated by apparent changes in hippocampal protein sialylation. (A) Field recordings ± neuraminidase ± neuraminidase blocker following multiple injections of higher [Ko+] used to induce epileptiform activity in vivo. (B) Summary data of seizure occurrence as a function of injection number. (C) Duration of ictal‐like activity. (D) Maximum population spike frequency.

Figure adapted, with permission, from D Isaev et al. J Neuroscience 27: 4311587‐11594, 2007 96. Reprinted with permission.


Figure 23.

Atrial action potential (AP) waveform is modulated by expression of the polysialyltransferase, ST8Sia2. APs were measured from isolated control and ST8sia2(−/−) neonatal atrial and ventricular myocytes. Top left: typical AP waveforms measured from control (in black) and ST8sia2(−/−) (in red) atrial myocytes. (A‐C) The mean ± SEM AP waveform parameters (relevant portions of AP marked with arrows). Black, control atrium (n = 7); red, ST8sia2 (−/−) atrium (n = 9); dark blue, control ventricle (n = 7); blue, ST8sia2(−/−) ventricle (n = 5). (A) Time to AP peak. (B) AP duration at 50% repolarization (APD50). (C) AP duration at 90% repolarization (APD90). Significance tested using a two‐tailed t test and comparing ST8sia2(−/−) to control parameters. *, significant (P<0.005); #, not significant (P>0.1). Figure and legend adapted, with permission, from ML Montpetit et al. Proc Natl Acad Sci, U S A 106: 16517‐16522, 2009 142.



Figure 24.

Atrial Nav gating is modulated by ST8Sia2 expression. ST8sia2(−/−) and control myocyte Nav gating. Left panels (A, C, E, and G): atrium, ST8sia2(−/−) (red; n = 10); control (black; n = 4). Right panels (B, D, F, and H): ventricle, ST8sia2(−/−) (blue; n = 4); control (dark blue; n = 6). (A and B) Steady‐state activation. Insets: half‐activation voltage. (C and D) Steady‐state channel availability (SSI). Dashed line: ST8sia2(−/−) data shifted along voltage axis by −7.3 mV (to mimic measured shift in half‐activation voltage). Insets: half‐inactivation voltage. (E and F) Fast inactivation time constants. Dashed line: The ST8sia2(−/−) data shifted along voltage axis by −7.3 mV. Insets: typical whole cell Na+ current traces elicited by stepping to a −40 mV test potential. (G and H) Time constants of recovery from fast inactivation at a −130 mV recovery potential. Significance tested using a two‐tailed t test and comparing ST8sia2(−/−) to control parameters. *, significant (P ≤ 0.005); #, not significant (P > 0.1). Figure and legend adapted, with permission, from ML Montpetit et al. Proc Natl Acad Sci, U S A 106: 16517‐16522, 2009 142.



Figure 25.

MLP−/− ventricular myocytes are susceptible to early afterdepolarizations. MLP−/− mice show extended QT intervals (A), increased action potential durations (B), and increased susceptibility to early after depolarizations (C and D).

Figure adapted, with permission, from CA Ufret‐Vincenty et al. J Biol Chem 276: 28197‐28203, 2001 199. Reprinted with permission.


Figure 26.

Gating of MLP−/− ventricular Nav is less sensitive to desialylation than the wild‐type channel. INa for desialylated control myocytes mimicked untreated and neuraminidase‐treated MLP−/− myocyte INa. (A) Current‐voltage (I‐V) relationships. (B) Conductance‐voltage (G‐V) relationships. (D) SSI curves. (E) Fast inactivation times constants.

Figure adapted, with permission, from CA Ufret‐Vincenty et al. J Biol Chem 276: 28197‐28203, 2001 199. Reprinted with permission.
References
 1. Agnew WS, Levinson SR, Brabson JS, Raftery MA. Purification of the tetrodotoxin‐binding component associated with the voltage‐sensitive sodium channel from Electrophorus electricus electroplax membranes. Proc Natl Acad Sci U S A 75: 2606‐2610, 1978.
 2. Aguilan JT, Sundaram S, Nieves E, Stanley P. Mutational and functional analysis of Large in a novel CHO glycosylation mutant. Glycobiology 19: 971‐986, 2009.
 3. Akhavan A, Atanasiu R, Shrier A. Identification of a COOH‐terminal segment involved in maturation and stability of human ether‐a‐go‐go‐related gene potassium channels. J Biol Chem 278: 40105‐40112, 2003.
 4. Albach C, Klein RA, Schmitz B. Do rodent and human brains have different N‐glycosylation patterns? Biol Chem 382: 187‐194, 2001.
 5. Altmann F, Staudacher E, Wilson IB, Marz L. Insect cells as hosts for the expression of recombinant glycoproteins. Glycoconj J 16: 109‐123, 1999.
 6. Andersen OS, Feldberg S, Nakadomari H, Levy S, McLaughlin S. Electrostatic interactions among hydrophobic ions in lipid bilayer membranes. Biophys J 21: 35‐70, 1978.
 7. Angata K, Fukuda M. Polysialyltransferases: Major players in polysialic acid synthesis on the neural cell adhesion molecule. Biochimie 85: 195‐206, 2003.
 8. Angata T, Varki A. Chemical diversity in the sialic acids and related alpha‐keto acids: An evolutionary perspective. Chem Rev 102: 439‐469, 2002.
 9. Apweiler R, Hermjakob H, Sharon N. On the frequency of protein glycosylation, as deduced from analysis of the SWISS‐PROT database. Biochim Biophys Acta 1473: 4‐8, 1999.
 10. Arber S, Hunter JJ, Ross J Jr, Hongo M, Sansig G, Borg J, Perriard JC, Chien KR, Caroni P. MLP‐deficient mice exhibit a disruption of cardiac cytoarchitectural organization, dilated cardiomyopathy, and heart failure. Cell 88: 393‐403, 1997.
 11. Arhem P. Voltage sensing in ion channels: A 50‐year‐old mystery resolved? Lancet 363: 1221‐1223, 2004.
 12. Armstrong CM. Sodium channels and gating currents. Physiol Rev 61: 644‐683, 1981.
 13. Armstrong CM, Bezanilla F. Currents related to movement of the gating particles of the sodium channels. Nature 242: 459‐461, 1973.
 14. Armstrong CM, Cota G. Calcium block of Na+ channels and its effect on closing rate. Proc Natl Acad Sci U S A 96: 4154‐4157, 1999.
 15. Armstrong CM, Cota G. Calcium ion as a cofactor in Na channel gating. Proc Natl Acad Sci U S A 88: 6528‐6531, 1991.
 16. Aumiller JJ, Hollister JR, Jarvis DL. A transgenic insect cell line engineered to produce CMP‐sialic acid and sialylated glycoproteins. Glycobiology 13: 497‐507, 2003.
 17. Bennett E, Urcan MS, Tinkle SS, Koszowski AG, Levinson SR. Contribution of sialic acid to the voltage dependence of sodium channel gating. A possible electrostatic mechanism. J Gen Physiol 109: 327‐343, 1997.
 18. Bennett ES. Isoform‐specific effects of sialic acid on voltage‐dependent Na+ channel gating: Functional sialic acids are localized to the S5‐S6 loop of domain I. J Physiol 538: 675‐690, 2002.
 19. Bezanilla F, Armstrong CM. Kinetic properties and inactivation of the gating currents of sodium channels in squid axon. Philos Trans R Soc Lond B Biol Sci 270: 449‐458, 1975.
 20. Blaustein MP, Goldman DE. The action of certain polyvalent cations on the voltage‐clamped lobster axon. J Gen Physiol 51: 279‐291, 1968.
 21. Boccaccio A, Moran O, Conti F. Calcium dependent shifts of Na+ channel activation correlated with the state dependence of calcium‐binding to the pore. Eur Biophys J 27: 558‐566, 1998.
 22. Borjesson SI, Elinder F. Structure, function, and modification of the voltage sensor in voltage‐gated ion channels. Cell Biochem Biophys 52: 149‐174, 2008.
 23. Brockhausen I, Schachter H, Stanley P. O‐GalNAc Glycans. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, editors. Essentials of Glycobiology (2nd ed). Cold Spring Harbor Laboratory Press, 2009, chap 9.
 24. Brooks NL, Corey MJ, Schwalbe RA. Characterization of N‐glycosylation consensus sequences in the Kv3.1 channel. FEBS J 273: 3287‐3300, 2006.
 25. Buraei Z, Yang J. The β subunit of voltage‐gated Ca2+ channels. Physiol Rev 90: 1461‐1506, 2010.
 26. Burda P, Aebi M. The dolichol pathway of N‐linked glycosylation. Biochim Biophys Acta 1426: 239‐257, 1999.
 27. Cabral MG, Piteira AR, Silva Z, Ligeiro D, Brossmer R, Videira PA. Human dendritic cells contain cell surface sialyltransferase activity. Immunol Lett 131: 89‐96, 2010.
 28. Cartwright TA, Corey MJ, Schwalbe RA. Complex oligosaccharides are N‐linked to Kv3 voltage‐gated K+ channels in rat brain. Biochim Biophys Acta 1770: 666‐671, 2007.
 29. Cartwright TA, Schwalbe RA. Atypical sialylated N‐glycan structures are attached to neuronal voltage‐gated potassium channels. Biosci Rep 29: 301‐313, 2009.
 30. Castillo C, Diaz ME, Balbi D, Thornhill WB, Recio‐Pinto E. Changes in sodium channel function during postnatal brain development reflect increases in the level of channel sialidation. Brain Res Dev Brain Res 104: 119‐130, 1997.
 31. Cha SK, Hu MC, Kurosu H, Kuro‐o M, Moe O, Huang CL. Regulation of renal outer medullary potassium channel and renal K(+) excretion by Klotho. Mol Pharmacol 76: 38‐46, 2009.
 32. Cha SK, Ortega B, Kurosu H, Rosenblatt KP, Kuro OM, Huang CL. Removal of sialic acid involving Klotho causes cell‐surface retention of TRPV5 channel via binding to galectin‐1. Proc Natl Acad Sci U S A 105: 9805‐9810, 2008.
 33. Chandrasekharan K, Martin PT. Genetic defects in muscular dystrophy. Methods Enzymol 479: 291‐322, 2010.
 34. Chang Q, Hoefs S, van der Kemp AW, Topala CN, Bindels RJ, Hoenderop JG. The beta‐glucuronidase klotho hydrolyzes and activates the TRPV5 channel. Science 310: 490‐493, 2005.
 35. Chen W, Stanley P. Five Lec1 CHO cell mutants have distinct Mgat1 gene mutations that encode truncated N‐acetylglucosaminyltransferase I. Glycobiology 13: 43‐50, 2003.
 36. Chen X, Varki A. Advances in the biology and chemistry of sialic acids. ACS Chem Biol 5: 163‐176, 2010.
 37. Cohen SA, Levitt LK. Partial characterization of the rH1 sodium channel protein from rat heart using subtype‐specific antibodies. Circ Res 73: 735‐742, 1993.
 38. Conti LR, Radeke CM, Vandenberg CA. Membrane targeting of ATP‐sensitive potassium channel. Effects of glycosylation on surface expression. J Biol Chem 277: 25416‐25422, 2002.
 39. Creemers EE, Pinto YM. Molecular mechanisms that control interstitial fibrosis in the pressure‐overloaded heart. Cardiovasc Res 89: 265‐272, 2011.
 40. Cronin NB, O'Reilly A, Duclohier H, Wallace BA. Effects of deglycosylation of sodium channels on their structure and function. Biochemistry 44: 441‐449, 2005.
 41. Cruz TF, Gurd JW. Identification of intrinsic sialidase and sialoglycoprotein substrates in rat brain synaptic junctions. J Neurochem 40: 1599‐1604, 1983.
 42. Cukierman S, Zinkand WC, French RJ, Krueger BK. Effects of membrane surface charge and calcium on the gating of rat brain sodium channels in planar bilayers. J Gen Physiol 92: 431‐447, 1988.
 43. Cummins TR, Dib‐Hajj SD, Black JA, Akopian AN, Wood JN, Waxman SG. A novel persistent tetrodotoxin‐resistant sodium current in SNS‐null and wild‐type small primary sensory neurons. J Neurosci 19: RC43, 1999.
 44. de Graffenried CL, Bertozzi CR. The roles of enzyme localisation and complex formation in glycan assembly within the Golgi apparatus. Curr Opin Cell Biol 16: 356‐363, 2004.
 45. Dib‐Hajj SD, Tyrrell L, Cummins TR, Black JA, Wood PM, Waxman SG. Two tetrodotoxin‐resistant sodium channels in human dorsal root ganglion neurons. FEBS Lett 462: 117‐120, 1999.
 46. Dorrscheidt‐Kafer M. The action of Ca2+, Mg2+ and H+ on the contraction threshold of frog skeletal muscle: Evidence for surface charges controlling electro‐mechanical coupling. Pflugers Arch 362: 33‐41, 1976.
 47. Eckhardt M, Gotza B, Gerardy‐Schahn R. Mutants of the CMP‐sialic acid transporter causing the Lec2 phenotype. J Biol Chem 273: 20189‐20195, 1998.
 48. Elinder F, Arhem P. The functional surface charge density of a fast K channel in the myelinated axon of Xenopus laevis. J Membr Biol 165: 175‐181, 1998.
 49. Elinder F, Arhem P. Role of individual surface charges of voltage‐gated K channels. Biophys J 77: 1358‐1362, 1999.
 50. Elinder F, Liu Y, Arhem P. Divalent cation effects on the Shaker K channel suggest a pentapeptide sequence as determinant of functional surface charge density. J Membr Biol 165: 183‐189, 1998.
 51. Elinder F, Madeja M, Arhem P. Surface Charges of K channels. Effects of strontium on five cloned channels expressed in Xenopus oocytes. J Gen Physiol 108: 325‐332, 1996.
 52. Esko JD, Stanley P. Glycosylation mutants of cultured cells. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, editors. Essentials of Glycobiology (2nd ed). Cold Spring Harbor Laboratory Press, 2009, chap 46.
 53. Esposito G, Santana LF, Dilly K, Cruz JD, Mao L, Lederer WJ, Rockman HA. Cellular and functional defects in a mouse model of heart failure. Am J Physiol Heart Circ Physiol 279: H3101‐H3112, 2000.
 54. Fedida D, Wible B, Wang Z, Fermini B, Faust F, Nattel S, Brown AM. Identity of a novel delayed rectifier current from human heart with a cloned K+ channel current. Circ Res 73: 210‐216, 1993.
 55. Feng J, Wible B, Li GR, Wang Z, Nattel S. Antisense oligodeoxynucleotides directed against Kv1.5 mRNA specifically inhibit ultrarapid delayed rectifier K+ current in cultured adult human atrial myocytes. Circ Res 80: 572‐579, 1997.
 56. Fermini B, Nathan RD. Sialic acid and the surface charge associated with hyperpolarization‐activated, inward rectifying channels. J Membr Biol 114: 61‐69, 1990.
 57. Fermini B, Nathan RD. Removal of sialic acid alters both T‐ and L‐type calcium currents in cardiac myocytes. Am J Physiol 260: H735‐H743, 1991.
 58. Footitt EJ, Karimova A, Burch M, Yayeh T, Dupre T, Vuillaumier‐Barrot S, Chantret I, Moore SE, Seta N, Grunewald S. Cardiomyopathy in the congenital disorders of glycosylation (CDG): A case of late presentation and literature review. J Inherit Metab Dis 2009. DOI 10.1007/510545‐009‐1262‐1
 59. Frankenhaeuser B, Hodgkin AL. The action of calcium on the electrical properties of squid axons. J Physiol 137: 218‐244, 1957.
 60. Freeman LC, Lippold JJ, Mitchell KE. Glycosylation influences gating and pH sensitivity of I(sK). J Membr Biol 177: 65‐79, 2000.
 61. Freeze HH, Aebi M. Altered glycan structures: The molecular basis of congenital disorders of glycosylation. Curr Opin Struct Biol 15: 490‐498, 2005.
 62. Gan L, Kaczmarek LK. When, where, and how much? Expression of the Kv3.1 potassium channel in high‐frequency firing neurons. J Neurobiol 37: 69‐79, 1998.
 63. Gaudry JP, Arod C, Sauvage C, Busso S, Dupraz P, Pankiewicz R, Antonsson B. Purification of the extracellular domain of the membrane protein GlialCAM expressed in HEK and CHO cells and comparison of the glycosylation. Protein Expr Purif 58: 94‐102, 2008.
 64. Gellens ME, George AL Jr, Chen LQ, Chahine M, Horn R, Barchi RL, Kallen RG. Primary structure and functional expression of the human cardiac tetrodotoxin‐insensitive voltage‐dependent sodium channel. Proc Natl Acad Sci U S A 89: 554‐558, 1992.
 65. Gilbert DL, Ehrenstein G. Effect of divalent cations on potassium conductance of squid axons: determination of surface charge. Biophys J 9: 447‐463, 1969.
 66. Goldin AL. Evolution of voltage‐gated Na(+) channels. J Exp Biol 205: 575‐584, 2002.
 67. Gong Q, Anderson CL, January CT, Zhou Z. Role of glycosylation in cell surface expression and stability of HERG potassium channels. Am J Physiol Heart Circ Physiol 283: H77‐H84, 2002.
 68. Grissmer S, Ghanshani S, Dethlefs B, McPherson JD, Wasmuth JJ, Gutman GA, Cahalan MD, Chandy KG. The Shaw‐related potassium channel gene, Kv3.1, on human chromosome 11, encodes the type l K+ channel in T cells. J Biol Chem 267: 20971‐20979, 1992.
 69. Gross HJ, Merling A, Moldenhauer G, Schwartz‐Albiez R. Ecto‐sialyltransferase of human B lymphocytes reconstitutes differentiation markers in the presence of exogenous CMP‐N‐acetyl neuraminic acid. Blood 87: 5113‐5126, 1996.
 70. Grudnik P, Bange G, Sinning I. Protein targeting by the signal recognition particle. Biol Chem 390: 775‐782, 2009.
 71. Grundfest H, Shanes AM, Freygang W. The effect of sodium and potassium ions on the impedance change accompanying the spike in the squid giant axon. J Gen Physiol 37: 25‐37, 1953.
 72. Guo W, Li H, Aimond F, Johns DC, Rhodes KJ, Trimmer JS, Nerbonne JM. Role of heteromultimers in the generation of myocardial transient outward K+ currents. Circ Res 90: 586‐593, 2002.
 73. Hagiwara S, Watanabe A. The effect of tetraethylammonium chloride on the muscle membrane examined with an intracellular microelectrode. J Physiol 129: 513‐527, 1955.
 74. Hahin R, Campbell DT. Simple shifts in the voltage dependence of sodium channel gating caused by divalent cations. J Gen Physiol 82: 785‐805, 1983.
 75. Hall MK, Cartwright TA, Fleming CM, Schwalbe RA. Importance of glycosylation on function of a potassium channel in neuroblastoma cells. PLoS One 6: e19317, 2011.
 76. Hanlon MR, Wallace BA. Structure and function of voltage‐dependent ion channel regulatory beta subunits. Biochemistry 41: 2886‐2894, 2002.
 77. Haris PI, Ramesh B, Sansom MS, Kerr ID, Srai KS, Chapman D. Studies of the pore‐forming domain of a voltage‐gated potassium channel protein. Protein Eng 7: 255‐262, 1994.
 78. Hatem SN, Coulombe A, Balse E. Specificities of atrial electrophysiology: Clues to a better understanding of cardiac function and the mechanisms of arrhythmias. J Mol Cell Cardiol 48: 90‐95, 2010.
 79. Helenius A, Aebi M. Intracellular functions of N‐linked glycans. Science 291: 2364‐2369, 2001.
 80. Henrissat B, Surolia A, Stanley P. A Genomic view of glycobiology. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, editors. Essentials of Glycobiology (2nd ed). Cold Spring Harbor Laboratory Press, 2009, chap 7.
 81. Herzog RI, Cummins TR, Waxman SG. Persistent TTX‐resistant Na+ current affects resting potential and response to depolarization in simulated spinal sensory neurons. J Neurophysiol 86: 1351‐1364, 2001.
 82. Hille B. Charges and potentials at the nerve surface. Divalent ions and pH. J Gen Physiol 51: 221‐236, 1968.
 83. Hille B. Ionic channels: Molecular pores of excitable membranes. Harvey Lect 82: 47‐69, 1986.
 84. Hille B. Ion Channels of Excitable Membranes. Sunderland, Mass.: Sinauer, 2001, 3rd edition, pp. 646–660.
 85. Hille B, Ritchie JM, Strichartz GR. The effect of surface charge on the nerve membrane on the action of tetrodotoxin and saxitoxin in frog myelinated nerve. J Physiol 250: 34P‐35P, 1975a.
 86. Hille B, Woodhull AM, Shapiro BI. Negative surface charge near sodium channels of nerve: Divalent ions, monovalent ions, and pH. Philos Trans R Soc Lond B Biol Sci 270: 301‐318, 1975b.
 87. Hodgkin AL, Huxley AF. The components of membrane conductance in the giant axon of Loligo. J Physiol 116: 473‐496, 1952a.
 88. Hodgkin AL, Huxley AF. Currents carried by sodium and potassium ions through the membrane of the giant axon of Loligo. J Physiol 116: 449‐472, 1952b.
 89. Hodgkin AL, Huxley AF. The dual effect of membrane potential on sodium conductance in the giant axon of Loligo. J Physiol 116: 497‐506, 1952c.
 90. Hodgkin AL, Huxley AF. Movement of sodium and potassium ions during nervous activity. Cold Spring Harb Symp Quant Biol 17: 43‐52, 1952d.
 91. Hodgkin AL, Huxley AF. A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 117: 500‐544, 1952e.
 92. Hodgkin AL, Huxley AF, Katz B. Measurement of current‐voltage relations in the membrane of the giant axon of Loligo. J Physiol 116: 424‐448, 1952.
 93. Huxley AF, Stampfli R. Effect of potassium and sodium on resting and action potentials of single myelinated nerve fibers. J Physiol 112: 496‐508, 1951.
 94. Imura A, Iwano A, Tohyama O, Tsuji Y, Nozaki K, Hashimoto N, Fujimori T, Nabeshima Y. Secreted Klotho protein in sera and CSF: Implication for post‐translational cleavage in release of Klotho protein from cell membrane. FEBS Lett 565: 143‐147, 2004.
 95. Isaev D, Isaeva E, Khazipov R, Holmes GL. Anticonvulsant action of GABA in the high potassium‐low magnesium model of ictogenesis in the neonatal rat hippocampus in vivo and in vitro. J Neurophysiol 94: 2987‐2992, 2005.
 96. Isaev D, Isaeva E, Shatskih T, Zhao Q, Smits NC, Shworak NW, Khazipov R, Holmes GL. Role of extracellular sialic acid in regulation of neuronal and network excitability in the rat hippocampus. J Neurosci 27: 11587‐11594, 2007.
 97. Isaev D, Zhao Q, Kleen JK, Lenck‐Santini PP, Adstamongkonkul D, Isaeva E, Holmes GL. Neuroaminidase reduces interictal spikes in a rat temporal lobe epilepsy model. Epilepsia 52: e12‐e15, 2011.
 98. Isaeva E, Lushnikova I, Savrasova A, Skibo G, Holmes GL, Isaev D. Blockade of endogenous neuraminidase leads to an increase of neuronal excitability and activity‐dependent synaptogenesis in the rat hippocampus. Eur J Neurosci 32: 1889‐1896, 2010.
 99. Isom LL. Sodium channel beta subunits: anything but auxiliary. Neuroscientist 7: 42‐54, 2001.
 100. Jaeken J. Congenital disorders of glycosylation. Ann N Y Acad Sci 1214: 190‐198, 2010.
 101. James WM, Agnew WS. Multiple oligosaccharide chains in the voltage‐sensitive Na channel from electrophorus electricus: Evidence for alpha‐2,8‐linked polysialic acid. Biochem Biophys Res Commun 148: 817‐826, 1987.
 102. Jegla T, Salkoff L. Molecular evolution of K+ channels in primitive eukaryotes. Soc Gen Physiol Ser 49: 213‐222, 1994.
 103. Jiang Y, Lee A, Chen J, Ruta V, Cadene M, Chait BT, MacKinnon R. X‐ray structure of a voltage‐dependent K+ channel. Nature 423: 33‐41, 2003.
 104. Jiang Y, Ruta V, Chen J, Lee A, MacKinnon R. The principle of gating charge movement in a voltage‐dependent K+ channel. Nature 423: 42‐48, 2003.
 105. Johnson D, Bennett ES. Isoform‐specific effects of the beta2 subunit on voltage‐gated sodium channel gating. J Biol Chem 281: 25875‐25881, 2006.
 106. Johnson D, Bennett ES. Gating of the shaker potassium channel is modulated differentially by N‐glycosylation and sialic acids. Pflugers Arch 456: 393‐405, 2008.
 107. Johnson D, Montpetit ML, Stocker PJ, Bennett ES. The sialic acid component of the beta1 subunit modulates voltage‐gated sodium channel function. J Biol Chem 279: 44303‐44310, 2004.
 108. Katz B, Miledi R. Propagation of electric activity in motor nerve terminals. Proc R Soc Lond B Biol Sci 161: 453‐482, 1965.
 109. Katz B, Miledi R. Spontaneous and evoked activity of motor nerve endings in calcium Ringer. J Physiol 203: 689‐706, 1969a.
 110. Katz B, Miledi R. Tetrodotoxin‐resistant electric activity in presynaptic terminals. J Physiol 203: 459‐487, 1969b.
 111. Keynes RD, Rojas E. Kinetics and steady‐state properties of the charged system controlling sodium conductance in the squid giant axon. J Physiol 239: 393‐434, 1974.
 112. Kuro‐o M. Klotho. Pflugers Arch 459: 333‐343, 2010.
 113. Kurosu H, Yamamoto M, Clark JD, Pastor JV, Nandi A, Gurnani P, McGuinness OP, Chikuda H, Yamaguchi M, Kawaguchi H, Shimomura I, Takayama Y, Herz J, Kahn CR, Rosenblatt KP, Kuro‐o M. Suppression of aging in mice by the hormone Klotho. Science 309: 1829‐1833, 2005.
 114. Lairson LL, Henrissat B, Davies GJ, Withers SG. Glycosyltransferases: Structures, functions, and mechanisms. Annu Rev Biochem 77: 521‐555, 2008.
 115. Lakshminarayanaiah N. Surface charges on membranes. J Membr Biol 29: 243‐253, 1976.
 116. Lakshminarayanaiah N, Murayama K. Estimation of surface charges in some biological membranes. J Membr Biol 23: 279‐292, 1975.
 117. Latorre R, Labarca P, Naranjo D. Surface charge effects on ion conduction in ion channels. Methods Enzymol 207: 471‐501, 1992.
 118. Lau A, McLaughlin A, McLaughlin S. The adsorption of divalent cations to phosphatidylglycerol bilayer membranes. Biochim Biophys Acta 645: 279‐292, 1981.
 119. Laughlin ST, Agard NJ, Baskin JM, Carrico IS, Chang PV, Ganguli AS, Hangauer MJ, Lo A, Prescher JA, Bertozzi CR. Metabolic labeling of glycans with azido sugars for visualization and glycoproteomics. Methods Enzymol 415: 230‐250, 2006.
 120. Levinson SR, Thornhill WB, Duch DS, Recio‐Pinto E, Urban BW. The role of nonprotein domains in the function and synthesis of voltage‐gated sodium channels. Ion Channels 2: 33‐64, 1990.
 121. Liu J, Kim KH, London B, Morales MJ, Backx PH. Dissection of the voltage‐activated potassium outward currents in adult mouse ventricular myocytes: I(to,f), I(to,s), I(K,slow1), I(K,slow2), and I(ss). Basic Res Cardiol 106: 189‐204, 2011.
 122. Long SB, Campbell EB, Mackinnon R. Crystal structure of a mammalian voltage‐dependent Shaker family K+ channel. Science 309: 897‐903, 2005.
 123. Long SB, Tao X, Campbell EB, MacKinnon R. Atomic structure of a voltage‐dependent K+ channel in a lipid membrane‐like environment. Nature 450: 376‐382, 2007.
 124. Lopreato GF, Lu Y, Southwell A, Atkinson NS, Hillis DM, Wilcox TP, Zakon HH. Evolution and divergence of sodium channel genes in vertebrates. Proc Natl Acad Sci U S A 98: 7588‐7592, 2001.
 125. Lu Z, Abe J, Taunton J, Lu Y, Shishido T, McClain C, Yan C, Xu SP, Spangenberg TM, Xu H. Reactive oxygen species‐induced activation of p90 ribosomal S6 kinase prolongs cardiac repolarization through inhibiting outward K+ channel activity. Circ Res 103: 269‐278, 2008.
 126. Mackinnon R. Structural biology. Voltage sensor meets lipid membrane. Science 306: 1304‐1305, 2004.
 127. Malin SA, Nerbonne JM. Elimination of the fast transient in superior cervical ganglion neurons with expression of KV4.2W362F: molecular dissection of IA. J Neurosci 20: 5191‐5199, 2000.
 128. Malin SA, Nerbonne JM. Molecular heterogeneity of the voltage‐gated fast transient outward K+ current, I(Af), in mammalian neurons. J Neurosci 21: 8004‐8014, 2001.
 129. Marchal I, Jarvis DL, Cacan R, Verbert A. Glycoproteins from insect cells: Sialylated or not? Biol Chem 382: 151‐159, 2001.
 130. Marengo FD, Wang SY, Wang B, Langer GA. Dependence of cardiac cell Ca2+ permeability on sialic acid‐containing sarcolemmal gangliosides. J Mol Cell Cardiol 30: 127‐137, 1998.
 131. Marionneau C, Brunet S, Flagg TP, Pilgram TK, Demolombe S, Nerbonne JM. Distinct cellular and molecular mechanisms underlie functional remodeling of repolarizing K+ currents with left ventricular hypertrophy. Circ Res 102: 1406‐1415, 2008.
 132. McDonagh JC, Nathan RD. Sialic acid and the surface charge of delayed rectifier potassium channels. J Mol Cell Cardiol 22: 1305‐1316, 1990.
 133. McDowell W, Schwarz RT. Dissecting glycoprotein biosynthesis by the use of specific inhibitors. Biochimie 70: 1535‐1549, 1988.
 134. McLaughlin S. The electrostatic properties of membranes. Annu Rev Biophys Biophys Chem 18: 113‐136, 1989.
 135. McLaughlin S, Mulrine N, Gresalfi T, Vaio G, McLaughlin A. Adsorption of divalent cations to bilayer membranes containing phosphatidylserine. J Gen Physiol 77: 445‐473, 1981.
 136. McLaughlin SG, Szabo G, Eisenman G. Divalent ions and the surface potential of charged phospholipid membranes. J Gen Physiol 58: 667‐687, 1971.
 137. McLaughlin SG, Szabo G, Eisenman G, Ciani SM. Surface charge and the conductance of phospholipid membranes. Proc Natl Acad Sci U S A 67: 1268‐1275, 1970.
 138. Miller JA, Agnew WS, Levinson SR. Principal glycopeptide of the tetrodotoxin/saxitoxin binding protein from Electrophorus electricus: Isolation and partial chemical and physical characterization. Biochemistry 22: 462‐470, 1983.
 139. Mitcheson JS, Sanguinetti MC. Biophysical properties and molecular basis of cardiac rapid and slow delayed rectifier potassium channels. Cell Physiol Biochem 9: 201‐216, 1999.
 140. Miyake A, Mochizuki S, Yokoi H, Kohda M, Furuichi K. New ether‐a‐go‐go K(+) channel family members localized in human telencephalon. J Biol Chem 274: 25018‐25025, 1999.
 141. Monti E, Bonten E, D'Azzo A, Bresciani R, Venerando B, Borsani G, Schauer R, Tettamanti G. Sialidases in vertebrates: A family of enzymes tailored for several cell functions. Adv Carbohydr Chem Biochem 64: 403‐479, 2010.
 142. Montpetit ML, Stocker PJ, Schwetz TA, Harper JM, Norring SA, Schaffer L, North SJ, Jang‐Lee J, Gilmartin T, Head SR, Haslam SM, Dell A, Marth JD, Bennett ES. Regulated and aberrant glycosylation modulate cardiac electrical signaling. Proc Natl Acad Sci U S A 106: 16517‐16522, 2009.
 143. Morimoto K, Fahnestock M, Racine RJ. Kindling and status epilepticus models of epilepsy: Rewiring the brain. Prog Neurobiol 73: 1‐60, 2004.
 144. Mozhayeva GN, Naumov AP. Effect of surface charge on the steady‐state potassium conductance of nodal membrane. Nature 228: 164‐165, 1970.
 145. Nakano M, Kakehi K, Tsai MH, Lee YC. Detailed structural features of glycan chains derived from alpha1‐acid glycoproteins of several different animals: The presence of hypersialylated, O‐acetylated sialic acids but not disialyl residues. Glycobiology 14: 431‐441, 2004.
 146. Neumcke B, Stampfli R. Heterogeneity of external surface charges near sodium channels in the nodal membrane of frog nerve. Pflugers Arch 401: 125‐131, 1984.
 147. Niwa N, Nerbonne JM. Molecular determinants of cardiac transient outward potassium current (I(to)) expression and regulation. J Mol Cell Cardiol 48: 12‐25.
 148. Noda M, Ikeda T, Suzuki H, Takeshima H, Takahashi T, Kuno M, Numa S. Expression of functional sodium channels from cloned cDNA. Nature 322: 826‐828, 1986.
 149. Noma K, Kimura K, Minatohara K, Nakashima H, Nagao Y, Mizoguchi A, Fujiyoshi Y. Triple N‐glycosylation in the long S5‐P loop regulates the activation and trafficking of the Kv12.2 potassium channel. J Biol Chem 284: 33139‐33150, 2009.
 150. Ohtsubo K, Marth JD. Glycosylation in cellular mechanisms of health and disease. Cell 126: 855‐867, 2006.
 151. Ohtsubo K, Takamatsu S, Minowa MT, Yoshida A, Takeuchi M, Marth JD. Dietary and genetic control of glucose transporter 2 glycosylation promotes insulin secretion in suppressing diabetes. Cell 123: 1307‐1321, 2005.
 152. Pabon A, Chan KW, Sui JL, Wu X, Logothetis DE, Thornhill WB. Glycosylation of GIRK1 at Asn119 and ROMK1 at Asn117 has different consequences in potassium channel function. J Biol Chem 275: 30677‐30682, 2000.
 153. Parker RB, Kohler JJ. Regulation of intracellular signaling by extracellular glycan remodeling. ACS Chem Biol 5: 35‐46, 2010.
 154. Patnaik SK, Stanley P. Lectin‐resistant CHO glycosylation mutants. Methods Enzymol 416: 159‐182, 2006.
 155. Petrecca K, Atanasiu R, Akhavan A, Shrier A. N‐linked glycosylation sites determine HERG channel surface membrane expression. J Physiol 515(Pt 1): 41‐48, 1999.
 156. Phartiyal P, Jones EM, Robertson GA. Heteromeric assembly of human ether‐a‐go‐go‐related gene (hERG) 1a/1b channels occurs cotranslationally via N‐terminal interactions. J Biol Chem 282: 9874‐9882, 2007.
 157. Pongs O, Schwarz JR. Ancillary subunits associated with voltage‐dependent K+ channels. Physiol Rev 90: 755‐796, 2010.
 158. Post JA. Removal of sarcolemmal sialic acid residues results in a loss of sarcolemmal functioning and integrity. Am J Physiol 263: H147‐H152, 1992.
 159. Puskin JS, Coene MT. Na+ and H+ dependent Mn2+ binding to phosphatidylserine vesicles as a test of the Gouy‐Chapman‐Stern theory. J Membr Biol 52: 69‐74, 1980.
 160. Recio‐Pinto E, Thornhill WB, Duch DS, Levinson SR, Urban BW. Neuraminidase treatment modifies the function of electroplax sodium channels in planar lipid bilayers. Neuron 5: 675‐684, 1990.
 161. Remme CA, Bezzina CR. Sodium channel (dys)function and cardiac arrhythmias. Cardiovasc Ther 28: 287‐294, 2010.
 162. Repnikova E, Koles K, Nakamura M, Pitts J, Li H, Ambavane A, Zoran MJ, Panin VM. Sialyltransferase regulates nervous system function in Drosophila. J Neurosci 30: 6466‐6476, 2010.
 163. Rockman HA, Chien KR, Choi DJ, Iaccarino G, Hunter JJ, Ross J Jr, Lefkowitz RJ, Koch WJ. Expression of a beta‐adrenergic receptor kinase 1 inhibitor prevents the development of myocardial failure in gene‐targeted mice. Proc Natl Acad Sci U S A 95: 7000‐7005, 1998.
 164. Rogart RB, Cribbs LL, Muglia LK, Kephart DD, Kaiser MW. Molecular cloning of a putative tetrodotoxin‐resistant rat heart Na+ channel isoform. Proc Natl Acad Sci U S A 86: 8170‐8174, 1989.
 165. Rugiero F, Mistry M, Sage D, Black JA, Waxman SG, Crest M, Clerc N, Delmas P, Gola M. Selective expression of a persistent tetrodotoxin‐resistant Na+ current and NaV1.9 subunit in myenteric sensory neurons. J Neurosci 23: 2715‐2725, 2003.
 166. Sato C, Sato M, Iwasaki A, Doi T, Engel A. The sodium channel has four domains surrounding a central pore. J Struct Biol 121: 314‐325, 1998.
 167. Savrasova AV, Lushnikova IV, Isaeva EV, Skibo GG, Isaev DS, Kostyuk PG. The effect of neuraminidase blocker on gabazine‐induced seizures in rat hippocampus. Fiziol Zh 56: 14‐18, 2010.
 168. Schwalbe RA, Bianchi L, Brown AM. Mapping the kidney potassium channel ROMK1. Glycosylation of the pore signature sequence and the COOH terminus. J Biol Chem 272: 25217‐25223, 1997.
 169. Schwalbe RA, Corey MJ, Cartwright TA. Novel Kv3 glycoforms differentially expressed in adult mammalian brain contain sialylated N‐glycans. Biochem Cell Biol 86: 21‐30, 2008.
 170. Schwalbe RA, Wang Z, Bianchi L, Brown AM. Novel sites of N‐glycosylation in ROMK1 reveal the putative pore‐forming segment H5 as extracellular. J Biol Chem 271: 24201‐24206, 1996.
 171. Schwalbe RA, Wang Z, Wible BA, Brown AM. Potassium channel structure and function as reported by a single glycosylation sequon. J Biol Chem 270: 15336‐15340, 1995.
 172. Schwartz‐Albiez R, Merling A, Martin S, Haas R, Gross HJ. Cell surface sialylation and ecto‐sialyltransferase activity of human CD34 progenitors from peripheral blood and bone marrow. Glycoconj J 21: 451‐459, 2004.
 173. Schwetz TA, Norring SA, Bennett ES. N‐glycans modulate K(v)1.5 gating but have no effect on K(v)1.4 gating. Biochim Biophys Acta 1798: 367‐375, 2010.
 174. Schwetz TA, Norring SA, Ednie AR, Bennett ES. Sialic acids attached to O‐glycans modulate voltage‐gated potassium channel gating. J Biol Chem 286: 4123‐4132, 2011.
 175. Shi G, Trimmer JS. Differential asparagine‐linked glycosylation of voltage‐gated K+ channels in mammalian brain and in transfected cells. J Membr Biol 168: 265‐273, 1999.
 176. Snider MD, Rogers OC. Membrane traffic in animal cells: Cellular glycoproteins return to the site of Golgi mannosidase I. J Cell Biol 103: 265‐275, 1986.
 177. Sparks SE, Krasnewich DM. Congenital disorders of glycosylation overview. In: Pagon RA, Bird TD, Dolan CR, Stephens K, editors. Gene Reviews. Seattle: University of Washington, Seattle, 2011.
 178. Spires S, Begenisich T. Modulation of potassium channel gating by external divalent cations. J Gen Physiol 104: 675‐692, 1994.
 179. Stanley P, Schachter H, Taniguchi N. N‐Glycans. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, editors. Essentials of Glycobiology (2nd ed). Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press, 2009, Chap 8.
 180. Stocker PJ, Bennett ES. Differential sialylation modulates voltage‐gated Na+ channel gating throughout the developing myocardium. J Gen Physiol 127: 253‐265, 2006.
 181. Suen KF, Turner MS, Gao F, Liu B, Althage A, Slavin A, Ou W, Zuo E, Eckart M, Ogawa T, Yamada M, Tuntland T, Harris JL, Trauger JW. Transient expression of an IL‐23R extracellular domain Fc fusion protein in CHO vs. HEK cells results in improved plasma exposure. Protein Expr Purif 71: 96‐102, 2010.
 182. Sutachan JJ, Watanabe I, Zhu J, Gottschalk A, Recio‐Pinto E, Thornhill WB. Effects of Kv1.1 channel glycosylation on C‐type inactivation and simulated action potentials. Brain Res 1058: 30‐43, 2005.
 183. Sutula TP. Experimental models of temporal lobe epilepsy: New insights from the study of kindling and synaptic reorganization. Epilepsia 31(Suppl 3): S45‐S54, 1990.
 184. Suzuki H, Beckh S, Kubo H, Yahagi N, Ishida H, Kayano T, Noda M, Numa S. Functional expression of cloned cDNA encoding sodium channel III. FEBS Lett 228: 195‐200, 1988.
 185. Suzuki T, Kitajima K, Inoue S, Inoue Y. N‐glycosylation/deglycosylation as a mechanism for the post‐translational modification/remodification of proteins. Glycoconj J 12: 183‐193, 1995.
 186. Suzuki T, Kitajima K, Inoue S, Inoue Y. Occurrence and biological roles of ‘proximal glycanases’ in animal cells. Glycobiology 4: 777‐789, 1994.
 187. Sweeley CC. Extracellular sialidases. Adv Lipid Res 26: 235‐252, 1993.
 188. Taatjes DJ, Roth J, Weinstein J, Paulson JC. Post‐Golgi apparatus localization and regional expression of rat intestinal sialyltransferase detected by immunoelectron microscopy with polypeptide epitope‐purified antibody. J Biol Chem 263: 6302‐6309, 1988.
 189. Takamatsu S, Antonopoulos A, Ohtsubo K, Ditto D, Chiba Y, Le DT, Morris HR, Haslam SM, Dell A, Marth JD, Taniguchi N. Physiological and glycomic characterization of N‐acetylglucosaminyltransferase‐IVa and ‐IVb double deficient mice. Glycobiology 20: 485‐497, 2010.
 190. Takashima S. Characterization of mouse sialyltransferase genes: Their evolution and diversity. Biosci Biotechnol Biochem 72: 1155‐1167, 2008.
 191. Tang K, Li X, Zheng MQ, Rozanski GJ. Role of apoptosis signal‐regulating kinase‐1‐c‐Jun NH2‐terminal kinase‐p38 signaling in voltage‐gated K+ channel remodeling of the failing heart: Regulation by thioredoxin. Antioxid Redox Signal 14: 25‐35, 2011.
 192. Tenno M, Ohtsubo K, Hagen FK, Ditto D, Zarbock A, Schaerli P, von Andrian UH, Ley K, Le D, Tabak LA, Marth JD. Initiation of protein O glycosylation by the polypeptide GalNAcT‐1 in vascular biology and humoral immunity. Mol Cell Biol 27: 8783‐8796, 2007.
 193. Thornhill WB, Watanabe I, Sutachan JJ, Wu MB, Wu X, Zhu J, Recio‐Pinto E. Molecular cloning and expression of a Kv1.1‐like potassium channel from the electric organ of Electrophorus electricus. J Membr Biol 196: 1‐8, 2003.
 194. Thornhill WB, Wu MB, Jiang X, Wu X, Morgan PT, Margiotta JF. Expression of Kv1.1 delayed rectifier potassium channels in Lec mutant Chinese hamster ovary cell lines reveals a role for sialidation in channel function. J Biol Chem 271: 19093‐19098, 1996.
 195. Tompkins‐Macdonald GJ, Gallin WJ, Sakarya O, Degnan B, Leys SP, Boland LM. Expression of a poriferan potassium channel: Insights into the evolution of ion channels in metazoans. J Exp Biol 212: 761‐767, 2009.
 196. Trepanier‐Boulay V, St‐Michel C, Tremblay A, Fiset C. Gender‐based differences in cardiac repolarization in mouse ventricle. Circ Res 89: 437‐444, 2001.
 197. Trudeau MC, Titus SA, Branchaw JL, Ganetzky B, Robertson GA. Functional analysis of a mouse brain Elk‐type K+ channel. J Neurosci 19: 2906‐2918, 1999.
 198. Tyrrell L, Renganathan M, Dib‐Hajj SD, Waxman SG. Glycosylation alters steady‐state inactivation of sodium channel Nav1.9/NaN in dorsal root ganglion neurons and is developmentally regulated. J Neurosci 21: 9629‐9637, 2001.
 199. Ufret‐Vincenty CA, Baro DJ, Lederer WJ, Rockman HA, Quinones LE, Santana LF. Role of sodium channel deglycosylation in the genesis of cardiac arrhythmias in heart failure. J Biol Chem 276: 28197‐28203, 2001a.
 200. Ufret‐Vincenty CA, Baro DJ, Santana LF. Differential contribution of sialic acid to the function of repolarizing K(+) currents in ventricular myocytes. Am J Physiol Cell Physiol 281: C464‐C474, 2001b.
 201. Ungar D. Golgi linked protein glycosylation and associated diseases. Semin Cell Dev Biol 20: 762‐769, 2009.
 202. Utsunomiya I, Tanabe S, Terashi T, Ikeno S, Miyatake T, Hoshi K, Taguchi K. Identification of amino acids in the pore region of Kv1.2 potassium channel that regulate its glycosylation and cell surface expression. J Neurochem 112: 913‐923, 2010.
 203. Van den Steen P, Rudd PM, Dwek RA, Opdenakker G. Concepts and principles of O‐linked glycosylation. Crit Rev Biochem Mol Biol 33: 151‐208, 1998.
 204. Varki A, Esko JD, Colley KJ. Cellular organization of glycosylation. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, editors. Essentials of Glycobiology (2nd ed). Cold Spring Harbor Laboratory Press, 2009, chap 3.
 205. Varki A, Schauer R. Sialic Acids. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, Hart GW, Etzler ME, . Essentials of Glycobiology (2nd ed). Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press, 2009, Chap 14.
 206. Waite A, Tinsley CL, Locke M, Blake DJ. The neurobiology of the dystrophin‐associated glycoprotein complex. Ann Med 41: 344‐359, 2009.
 207. Wang Y, Sun Z. Current understanding of klotho. Ageing Res Rev 8: 43‐51, 2009.
 208. Warmke JW, Ganetzky B. A family of potassium channel genes related to eag in Drosophila and mammals. Proc Natl Acad Sci U S A 91: 3438‐3442, 1994.
 209. Watanabe I, Wang HG, Sutachan JJ, Zhu J, Recio‐Pinto E, Thornhill WB. Glycosylation affects rat Kv1.1 potassium channel gating by a combined surface potential and cooperative subunit interaction mechanism. J Physiol 550: 51‐66, 2003.
 210. Watanabe I, Zhu J, Recio‐Pinto E, Thornhill WB. Glycosylation affects the protein stability and cell surface expression of Kv1.4 but Not Kv1.1 potassium channels. A pore region determinant dictates the effect of glycosylation on trafficking. J Biol Chem 279: 8879‐8885, 2004.
 211. Watanabe I, Zhu J, Sutachan JJ, Gottschalk A, Recio‐Pinto E, Thornhill WB. The glycosylation state of Kv1.2 potassium channels affects trafficking, gating, and simulated action potentials. Brain Res 1144: 1‐18, 2007.
 212. Watanabe S, Kokuho T, Takahashi H, Takahashi M, Kubota T, Inumaru S. Sialylation of N‐glycans on the recombinant proteins expressed by a baculovirus‐insect cell system under beta‐N‐acetylglucosaminidase inhibition. J Biol Chem 277: 5090‐5093, 2002.
 213. Weidmann S. Effects of calcium ions and local anesthetics on electrical properties of Purkinje fibres. J Physiol 129: 568‐582, 1955.
 214. Yee HF Jr, Kuwata JH, Langer GA. Effects of neuraminidase on cellular calcium and contraction in cultured cardiac myocytes. J Mol Cell Cardiol 23: 175‐185, 1991.
 215. Yee HF Jr, Weiss JN, Langer GA. Neuraminidase selectively enhances transient Ca2+ current in cardiac myocytes. Am J Physiol 256: C1267‐C1272, 1989.
 216. Yohe HC, Rosenberg A. Action of intrinsic sialidase of rat brain synaptic membranes on membrane sialolipid and sialoprotein components in situ. J Biol Chem 252: 2412‐2418, 1977.
 217. Zhang Y, Hartmann HA, Satin J. Glycosylation influences voltage‐dependent gating of cardiac and skeletal muscle sodium channels. J Membr Biol 171: 195‐207, 1999.
 218. Zhou W, Jones SW. Surface charge and calcium channel saturation in bullfrog sympathetic neurons. J Gen Physiol 105: 441‐462, 1995.
 219. Zhu J, Recio‐Pinto E, Hartwig T, Sellers W, Yan J, Thornhill WB. The Kv1.2 potassium channel: the position of an N‐glycan on the extracellular linkers affects its protein expression and function. Brain Res 1251: 16‐29, 2009.
 220. Zhu J, Watanabe I, Gomez B, Thornhill WB. Determinants involved in Kv1 potassium channel folding in the endoplasmic reticulum, glycosylation in the Golgi, and cell surface expression. J Biol Chem 276: 39419‐39427, 2001.
 221. Zhu J, Watanabe I, Gomez B, Thornhill WB. Heteromeric Kv1 potassium channel expression: amino acid determinants involved in processing and trafficking to the cell surface. J Biol Chem 278: 25558‐25567, 2003.
 222. Zolk O, Caroni P, Bohm M. Decreased expression of the cardiac LIM domain protein MLP in chronic human heart failure. Circulation 101: 2674‐2677, 2000.
 223. Zuber C, Lackie PM, Catterall WA, Roth J. Polysialic acid is associated with sodium channels and the neural cell adhesion molecule N‐CAM in adult rat brain. J Biol Chem 267: 9965‐9971, 1992.

Related Articles:

Ionic Basis of Resting and Action Potentials
K+ Channels: Function‐Structural Overview
The Cardiac Ventricular Action Potential

Contact Editor

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

* Required Field

How to Cite

Andrew R. Ednie, Eric S. Bennett. Modulation of Voltage‐Gated Ion Channels by Sialylation. Compr Physiol 2012, 2: 1269-1301. doi: 10.1002/cphy.c110044