Skip to main content
The Journal of Clinical Investigation logoLink to The Journal of Clinical Investigation
. 2010 Sep 27;120(10):3508–3519. doi: 10.1172/JCI43621

A βIV-spectrin/CaMKII signaling complex is essential for membrane excitability in mice

Thomas J Hund 1, Olha M Koval 1, Jingdong Li 1, Patrick J Wright 1, Lan Qian 1, Jedidiah S Snyder 1, Hjalti Gudmundsson 1, Crystal F Kline 1, Nathan P Davidson 1, Natalia Cardona 1, Matthew N Rasband 2, Mark E Anderson 1,3, Peter J Mohler 1,3
PMCID: PMC2947241  PMID: 20877009

Abstract

Ion channel function is fundamental to the existence of life. In metazoans, the coordinate activities of voltage-gated Na+ channels underlie cellular excitability and control neuronal communication, cardiac excitation-contraction coupling, and skeletal muscle function. However, despite decades of research and linkage of Na+ channel dysfunction with arrhythmia, epilepsy, and myotonia, little progress has been made toward understanding the fundamental processes that regulate this family of proteins. Here, we have identified βIV-spectrin as a multifunctional regulatory platform for Na+ channels in mice. We found that βIV-spectrin targeted critical structural and regulatory proteins to excitable membranes in the heart and brain. Animal models harboring mutant βIV-spectrin alleles displayed aberrant cellular excitability and whole animal physiology. Moreover, we identified a regulatory mechanism for Na+ channels, via direct phosphorylation by βIV-spectrin–targeted calcium/calmodulin-dependent kinase II (CaMKII). Collectively, our data define an unexpected but indispensable molecular platform that determines membrane excitability in the mouse heart and brain.

Introduction

Membrane excitability requires the coordinate function of precisely synchronized signaling networks. Inherent to this precision are evolved molecular pathways that compartmentalize specific signaling molecules to ensure tight spatial and temporal coupling between plasma membrane and intracellular receptors, effector proteins, and target molecules. Members of the calcium/calmodulin-dependent protein kinase II (CaMKII) family — CaMKIIα, CaMKIIβ, CaMKIIγ, and CaMKIIδ — are multifunctional serine/threonine kinases with critical roles in both excitable and non-excitable cells. CaMKII regulates diverse cellular functions, including ion channel biophysics, organelle transport, metabolism, and transcription, to modulate synaptic plasticity, cardiac excitation-contraction coupling, and hormone secretion (15). Furthermore, CaMKII signaling has been linked to specific disease phenotypes (513), including human heart failure and cognitive defects (14, 15), through its effects on a host of diverse and spatially distinct target proteins including ion channels and transporters (i.e., voltage-gated Na+ channels; refs. 1618), transcription factors, and cell death pathways (1922). Finally, CaMKII inhibition has shown exciting promise for the treatment of excitable cell disease (5, 2326). Collectively, these data strongly support the notion that local CaMKII/effector signaling nodes represent key cellular rheostats to translate local alterations in the cellular environment to global changes in membrane excitability and organism function.

Here, we define what we believe to be a novel signaling platform for the regulation of membrane excitability. Specifically, we identify βIV-spectrin, a key structural component required for ion channel clustering (including voltage-gated Na+ channels) in the nervous system (2734), as a multifunctional regulatory stage for Na+ channel signaling in excitable cells. βIV-spectrin targets critical structural and regulatory proteins to excitable membranes in heart and brain, and animal models harboring mutant βIV-spectrin alleles display aberrant cellular excitability and whole-animal physiology. Here we show a fundamental, but unanticipated, requirement for βIV-spectrin-dependent targeting of CaMKII to a controlling phosphorylation site, S571, on the dominant cardiac Na+ channel (Nav1.5). Moreover, our findings provide evidence for a similar targeting and regulatory complex in neurons. Collectively, our data define an unexpected yet commanding molecular platform that determines vertebrate membrane excitability.

Results

Identification of novel CaMKII-binding proteins.

We screened the human genome for novel CaMKII-binding proteins using an algorithm derived from the CaMKII autoregulatory domain sequence and identified 32 candidates. Putative CaMKII-binding molecules included nuclear, cytosolic, and mitochondrial proteins with a host of disparate roles, including cell metabolism, cytoskeletal dynamics, and signaling (Figure 1, A and B). All CaMKII gene products (α, β, γ, and δ) were recognized by the screen; notably, only 1 known CaMKII-binding partner was identified (35). All candidates were cloned from human tissue (brain or heart), and CaMKII-binding activity was assessed by in vitro binding assays using radiolabeled target proteins and activated CaMKII (CaMKII T287D). Of 32 candidates containing the consensus CaMKII autoregulatory domain consensus motif, only 12 clones displayed remarkable CaMKII-binding activity in vitro (see Supplemental Figure 1 for examples; supplemental material available online with this article; doi: 10.1172/JCI43621DS1). Positive targets included critical signaling molecules (PKCε, phospholipase A2 zeta; ref. 36), cytoskeletal-associated proteins (SPTBN4 and NEDD1, ref. 37; girdin, ref. 38), and ion channel regulatory molecules. Notably, targets that shared the highest sequence identity with the CaMKII autoregulatory domain (i.e., SPTBN4) retained the most consistent kinase binding activity (Figure 1A and Supplemental Figure 1).

Figure 1. Identification of putative new CaMKII-binding proteins.

Figure 1

(A) Putative CaMKII-targeting proteins identified through a screen of the human genome using a sequence from the CaMKII autoregulatory domain as bait. Candidates with the highest homology to CaMKII autoregulatory domain are listed first. (B) Candidate targeting molecules included cytoskeletal, nuclear, cytoplasmic, and mitochondrial proteins with identified roles in cell metabolism, cytoskeletal dynamics, and signaling. All CaMKII gene products (α, β, γ, and δ) were recognized by the screen; notably, only 1 known CaMKII-binding partner was identified (β2a). Candidates were cloned from human cDNA, and CaMKII-binding activity was assessed by in vitro binding assays using radiolabeled target proteins and activated CaMKII (CaMKII T287D). Notably, only clones underlined in red in A showed robust binding for CaMKII.

βIV-spectrin associates with CaMKII.

SPTBN4 encodes βIV-spectrin (Figure 2A), an actin-associated protein with roles in nervous system membrane biogenesis and maintenance as well as in ion channel clustering (39, 40). Importantly, all βIV-spectrin orthologs harbored sequences nearly identical to the CaMKII autoregulatory motif (Figure 2B). Based on these characteristics, we hypothesized that βIV-spectrin may target CaMKII in excitable cells. We first verified the βIV-spectrin/CaMKII association using radiolabeled CaMKIIδ and a biotinylated peptide mimicking the putative kinase-binding domain in human βIV-spectrin (CTP-P; residues 2,292–2,317). Consistent with a specific interaction, CTP-P, but not scrambled peptide control (CTP-C), bound radiolabeled CaMKII (Figure 2C).

Figure 2. βIV-spectrin is a CaMKII-binding protein in heart.

Figure 2

(A) βIV-spectrin contains an N-terminal actin-binding domain (NTD), 17 spectrin repeats, and specific and C-terminal domains (SD/CTD). The putative CaMKII-binding site is denoted by an asterisk. (B) The putative CaMKII-binding domain in βIV-spectrin was homologous to a CaMKIIδ autoregulatory domain motif and conserved across orthologs. (C) CTP-P bound radiolabeled CaMKIIδ; CTP-C and GST beads alone lacked binding. (D) βIV-spectrin RNA levels in adult rat brain and heart. (E) βIV-spectrin (Σ1 and Σ6) in ventricular lysates from multiple species. βIV-spectrin was expressed approximately 8–10 fold higher in cerebellum than in heart. Cardiac βIV-spectrin migrated approximately 4 kDa larger than did cerebellar βIV-spectrin. (F) CTP-P, but not CTP-C, associated with CaMKIIδ from rat heart. (G) Endogenous CaMKIIδ and βIV-spectrin coimmunoprecipitated from adult heart lysate. (H and I) βIV-spectrin and N-cadherin in rat cardiomyocytes. Nuclei are shown by Topro-3 dye (blue). (J) CaMKIIδ localization in adult rat myocytes. CaMKIIδ localized to the intercalated disc (white arrows) here and to a second population at transverse-tubules (yellow arrows). Ventricular sections stained for (K) βIV-spectrin, (L) N-cadherin, and (M) CaMKIIδ showed coexpression of these proteins at the intercalated disc (white arrows). (N and P) Ankyrin-G and (O and Q) Nav1.5 were also found at the intercalated disc (white arrows) in rat myocytes and tissue sections. (RU) Coimmunoprecipitation studies demonstrate cardiac complex of βIV-spectrin, CaMKIIδ, ankyrin-G (AnkG), and Nav1.5. Scale bars: 10 μm (HQ).

βIV-spectrin is expressed in the heart and associates with CaMKII in vivo.

Based on the critical role of CaMKII in regulating cardiac membrane excitability (5), we tested human heart for βIV-spectrin expression. Notably, both brain and heart expressed βIV-spectrin mRNA (Figure 2D). Moreover, βIV-spectrin Σ1 and Σ6 isoforms were readily detected in the left ventricle of multiple species by immunoblot, albeit at lower levels than were observed in brain (Figure 2E). In agreement with peptide binding experiments using radiolabeled CaMKII, CTP-P, but not CTP-C, interacted with endogenous CaMKIIδ from heart (Figure 2F). Moreover, consistent with an in vivo interaction, endogenous βIV-spectrin associated with endogenous CaMKIIδ in coimmunoprecipitation experiments (Figure 2G).

At the cellular level, βIV-spectrin was concentrated at the intercalated disc of cardiomyocytes (colocalized with N-cadherin; Figure 2, H, I, K, and L, and Supplemental Figure 2), an excitable membrane domain critical for cardiac action potential (AP) initiation and impulse conduction. We identified 2 distinct subcellular distributions of CaMKII in myocytes. Consistent with a potential in vivo interaction, CaMKIIδ (the primary CaMKII isoform in heart; ref. 41) was coexpressed with βIV-spectrin at the intercalated disc (Figure 2, J and M). A second population of CaMKIIδ was localized to the myocyte transverse-tubule network, apparently independent of βIV-spectrin (Figure 2J, yellow arrowheads). Together, our data identify βIV-spectrin as a cardiac protein that associates with endogenous CaMKII.

βIV-spectrin organizes a macromolecular signaling complex in the heart.

In the central nervous system, βIV-spectrin and ankyrin-G are highly enriched at axon initial segments (AISs) and nodes of Ranvier, and the clustering of Na+ channels at these sites depends on their interaction with ankyrin-G (31, 42). Furthermore, mice deficient in βIV-spectrin display defects in neuronal Na+ channel clustering and function (29). In heart, ankyrin-G targets voltage-gated Na+ channels (Nav1.5) to the intercalated disc (43, 44), and cardiac Na+ channel membrane regulation has been associated with CaMKII activity (16). We therefore tested whether the βIV-spectrin/CaMKII complex formed a macromolecular complex with ankyrin-G and Nav1.5 in heart. Similar to βIV-spectrin and CaMKIIδ, ankyrin-G and Nav1.5 were enriched at the myocyte intercalated disc (Figure 2, N–Q). Moreover, ankyrin-G and Nav1.5 associated with both βIV-spectrin and CaMKIIδ by coimmunoprecipitation (Figure 2, R–U). Thus, βIV-spectrin coordinates an in vivo cardiac macromolecular complex, effectively linking a critical signaling protein (CaMKIIδ) with its target substrate (Nav1.5) at a specialized cellular domain.

βIV-spectrin targets CaMKII to cardiomyocyte intercalated disc.

We evaluated the role of βIV-spectrin for CaMKII targeting in vivo using primary cardiomyocytes and heart tissue from the qv3J mouse strain (45). qv3J mice harbor a nucleotide insertion (InsT6786) in the βIV-spectrin gene immediately proximal to the CaMKII-binding motif, resulting in a premature C-terminal truncation (Figure 3A and ref. 45). In contrast to WT βIV-spectrin, the mutant qv3J polypeptide lacked cardiac CaMKIIδ-binding activity (Figure 3, B–D), as predicted by our peptide binding studies (Figure 2, C and F). However, both WT and mutant proteins retained binding activity for ankyrin-G (Figure 3E), actin, and α-spectrin (data not shown). Consistent with the location of the qv3J mutation, we observed association of ankyrin-G with CaMKIIδ from WT but not qv3J heart (Figure 3F). Thus, the βIV-spectrin qv3J mouse represented an ideal in vivo system to directly test the role of βIV-spectrin for CaMKII targeting.

Figure 3. βIV-spectrin is required for CaMKIIδ targeting.

Figure 3

(A) βIV-spectrin organization in WT and qv3J animals. qv3J animals lacked a CaMKII-binding domain, but retained actin-, ankyrin-, and α-spectrin–binding domains. (B and C) Schematic of control GST–βIV-spectrin fusion protein encompassing spectrin repeats 13–17 and the CaMKII-binding domain (βIV-WT) and truncated mutant lacking CaMKII-binding domain (βIV-qv3J). (D) βIV-spectrin WT and qv3J GST fusion proteins were incubated with detergent-soluble rat heart lysate and analyzed by immunoblot (CaMKIIδ). L-type Ca2+ channel β2a subunit was also expressed as a GST fusion protein and used as positive control. CaMKIIδ bound to the β2a and WT GST fusion proteins, but not to the qv3J GST fusion protein. Lanes were run on the same gel but were noncontiguous (white line). (E) βIV-spectrin WT and qv3J GST fusion proteins retained binding activity for ankyrin-G. (F) Coimmunoprecipitation studies showing association of ankyrin-G and CaMKIIδ in WT, but not qv3J, hearts. (G) Expression of ankyrin-G, Nav1.5, CaMKIIδ, β-catenin, and N-cadherin in heart lysates from WT and qv3J animals. Actin is shown as loading control. (HQ) Permeabilized adult rat cardiomyocytes from (HL) WT and (MQ) qv3J hearts were immunostained for (H and M) N-cadherin, (I and N) total CaMKIIδ, (J and O) CaMKII-phospho-T287, (K and P) βIV-spectrin, and (L and Q) Nav1.5. Localization of CaMKIIδ (total and phospho-T287) to the intercalated disc (white arrows) was disrupted in qv3J cardiomyocytes. Scale bars: 10 μm (HQ).

Consistent with the location of the qv3J mutation, qv3J mouse cardiomyocytes displayed normal expression of βIV-spectrin, ankyrin-G, and Nav1.5 as well as normal levels of intercalated disc proteins N-cadherin and β-catenin (Figure 3, G, H, K, and L, and Supplemental Figures 3 and 4). However, CaMKIIδ expression levels were modestly reduced in qv3J hearts (Figure 3G and Supplemental Figure 3). Parallel immunostaining experiments revealed a striking absence of CaMKIIδ from the myocyte intercalated disc (Figure 3, I and N, and Supplemental Figure 5). In contrast, CaMKIIδ localization at transverse-tubule membranes was unaffected (Figure 3, I and N, and Supplemental Figure 5), indicative of a βIV-spectrin–independent pathway that localizes CaMKII to transverse-tubules. Notably, we also observed dramatic reduction of activated CaMKIIδ (identified by CaMKII T287 phosphorylation) at the intercalated disc of qv3J ventricular cardiomyocytes (Figure 3, J and O). Finally, consistent with the noted lack of effect of the qv3J allele on ankyrin-G binding, we observed no difference in Nav1.5 (Figure 3, L and Q) or ankyrin-G (data not shown) intercalated disc targeting. We believe our findings identify βIV-spectrin as a novel CaMKII targeting protein in excitable cardiac cells.

βIV-spectrin is required for local CaMKII signaling in cardiomyocytes.

We tested the requirement of the βIV-spectrin–dependent CaMKII targeting pathway for normal cardiomyocyte function by assessing the activity of the resident intercalated disc voltage-gated Na+ channel in primary qv3J ventricular cardiomyocytes. qv3J cardiomyocytes displayed aberrant CaMKII-dependent Na+ channel regulation at baseline and in response to the β-adrenergic receptor agonist isoproterenol, which activates CaMKII (5). Specifically, qv3J myocytes showed a significant increase in peak Na+ current (INa; Figure 4, A–C) and a depolarizing shift in steady-state inactivation (Figure 4, D and E) compared with WT mouse ventricular cardiomyocytes. Additionally, persistent INa (INa,p), measured as current amplitude 50 ms after the peak value, was significantly reduced in qv3J myocytes compared with littermate controls (Figure 4, F–I). Differences in INa,p between genotypes were eliminated by mexiletine (Figure 4, F–H), a cardiac Nav1.5 antagonist used to treat patients with type 3 long QT syndrome, a genetic syndrome that predisposes affected hearts to increased INa,p and arrhythmias. Isoproterenol (1 μM) produced a depolarizing shift in steady-state inactivation and increased INa,p in ventricular cardiomyocytes derived from WT mice, but not qv3J mice or AC3-I transgenic mice (Figure 4, E and I), which overexpress a CaMKII-inhibitory peptide (5). In fact, we verified that differences in intercalated disc Na+ channel properties were due to specific elimination of CaMKII by testing the identical parameters in ventricular cardiomyocytes derived from AC3-I mice. Notably, all tested INa parameters were identical between AC3-I and qv3J mice (Figure 4, C–E). These findings demonstrate that the CaMKII-binding function of βIV-spectrin is required for CaMKII actions on INa in excitable cardiac myocytes.

Figure 4. βIV-spectrin/CaMKII complex is required for myocyte Na+ channel function.

Figure 4

(A and B) Whole-cell patch clamp INa traces from WT and qv3J cardiomyocytes. Test pulse potential is listed next to each trace. (C) Current-voltage relationship for cardiomyocytes from WT, qv3J, and AC3-I mice (n = 8 per group). *P < 0.05 versus WT. (D and E) Voltage-gated Na+ channel steady-state inactivation measured from WT, qv3J, and AC3-I cardiomyocytes (n = 8 per group). *P < 0.05 versus WT. Pulse protocol is shown in the inset of D. (F and G) INa,p from WT and qv3J mice with or without 10 μM mexiletine. Persistent current was determined as amplitude 50 ms after time of peak. (H and I) INa,p in WT (n = 10) and qv3J (n = 17) cardiomyocytes at baseline and in the presence of 10 μm mexiletine or isoproterenol. *P < 0.05. (J) Whole-cell patch clamp Ca2+ current-voltage relationship for WT (n = 9) and qv3J (n = 8) cardiomyocytes. (K) Ca2+ current facilitation (peak current in response to a train of depolarizing voltage pulses to 0 mV) in WT (n = 9) and qv3J (n = 7) cardiomyocytes. Peak current is expressed as percent increase from first pulse. (L and M) Transverse-tubule CaMKII labeling was preserved in qv3J mouse myocytes. Scale bars: 5 μm (L and M).

Finally, past work has demonstrated a clear role of CaMKII in regulation of L-type Ca2+ channel (Cav1.2) Ca2+ current (ICa,L) facilitation (35). As this functional property is specific to the cardiomyocyte transverse-tubule membrane network (35), we tested whether loss of CaMKII targeting by βIV-spectrin in qv3J mice would display defects in Ca2+ current facilitation. Consistent with a highly selective role of βIV-spectrin in the organization of CaMKII at the intercalated disc, we observed no differences in peak Ca2+ current or Ca2+ current facilitation between qv3J and WT cardiomyocytes (Figure 4, J–M). Moreover, WT and qv3J cardiomyocytes showed no difference in either total myocyte K+ current or the transient outward K+ current (Ito; Supplemental Figure 6; n = 12 myocytes/genotype; P = NS), the primary murine repolarizing current (46) and a target for CaMKII in heart (47, 48). Collectively, these data provide the first description to our knowledge of the molecular mechanism underlying CaMKII targeting to the cardiomyocyte intercalated disc, and establish βIV-spectrin as a bona fide CaMKII targeting protein in vivo. Notably, although our data link the βIV-spectrin/CaMKII complex with Nav1.5 regulation (and not ICa,L or Ito regulation), they do not exclude other intercalated disc proteins (e.g., channels, receptors, and signaling molecules) as potential targets of the regulatory complex.

CaMKII regulates Nav1.5 via direct phosphorylation of Nav1.5 S571.

While CaMKII has previously been shown to regulate cardiac INa (16, 18), the molecular mechanism underlying this modulation has remained elusive. Based on our findings, we hypothesized that CaMKII directly regulates INa by direct phosphorylation of Nav1.5. To test this hypothesis in cells, we measured INa properties from a library of full-length human Nav1.5 (SCN5A) mutants engineered to harbor single serine-to-alanine mutations at all putative consensus CaMKII phosphorylation sites (RXX[S/T]; Figure 5A). Notably, with the exception of S571A (localized in the cytoplasmic domain loop DI–DII), WT Nav1.5 and all mutant Nav1.5 channels showed a significant depolarizing shift in Na+ channel steady-state inactivation in the presence of CaMKII T287D (Figure 5, B–D). Based on these findings, we tested an additional S571E (phosphomimetic) mutant and observed a significant leftward shift in steady-state inactivation at baseline, with no CaMKII response in INa (Figure 5, F and G). These biophysical data strongly suggest that Nav1.5 is directly phosphorylated by CaMKII, and that the CaMKII regulation is dependent on Nav1.5 S571. In vitro phosphorylation assays (using purified Nav1.5 cytoplasmic domains and activated CaMKII) validated our functional findings, showing that Nav1.5 was directly phosphorylated on DI–DII (Figure 5E) and that S571 was a primary site for this regulation (Supplemental Figure 7). Together, these data identify S571 in the DI–DII loop as a critical site for regulation of Nav1.5 activity by CaMKII.

Figure 5. βIV-spectrin/CaMKII regulates Nav1.5 phosphorylation.

Figure 5

(A) CaMKII consensus phosphorylation sites in intracellular domains of Nav1.5. (B and C) Na+ channel steady-state inactivation measured from WT and Nav1.5 mutant channels expressed in HEK cells. *P < 0.05 versus WT. (D) Summary data showing depolarizing shift in steady-state inactivation V1/2 following CaMKII activation for WT and all mutants except S571A. Black, before CaMKII activation; blue, after CaMKII activation. (E) CaMKII phosphorylation assay on intracellular domains of Nav1.5. β2a was used as positive control. Asterisks denote location of purified proteins. (F and G) Na+ channel steady-state inactivation measured from WT, S571A, and S571E (phosphomimetic) channels expressed in HEK cells. *P < 0.05 versus WT. (H) Nav1.5 S571 antibody recognized WT, but not Nav1.5 S571A mutant, channels. WT and mutant channels were expressed at equivalent levels in HEK293 cells overexpressing active CaMKIIδ. (I) Immunoblots showing reduced levels of phospho–Nav1.5 S571, but unchanged total Nav1.5 levels, in qv3J versus WT heart lysates. Actin is shown as loading control.

βIV-spectrin regulates direct phosphorylation of Nav1.5 by CaMKII.

We created a custom phosphospecific antibody against Nav1.5 S571 to test the role of the βIV-spectrin/CaMKII pathway for Nav1.5 regulation in heart (Figure 5, H and I). Consistent with (a) a role of βIV-spectrin for CaMKII regulation in vivo and (b) a role of CaMKII for direct regulation of Nav1.5 in vivo, we observed decreased levels of phospho-Nav1.5 S571 in qv3J compared with WT hearts (Figure 5I and Supplemental Figure 8). In contrast, total Nav1.5 levels were unchanged between WT and qv3J mice (Figure 5I). These data support the hypothesis that βIV-spectrin creates a platform for regulation of cardiac Na+ channel activity through direct phosphorylation by CaMKII.

βIV-spectrin is critical for cardiac membrane excitability.

We next tested the role of βIV-spectrin for CaMKII-dependent regulation of cardiac excitability. Voltage-gated Na+ channel current is required for the rapid upstroke of the cardiac AP and normal electrical conduction in myocardium (49). Furthermore, alterations in INa and AP prolongation contribute to congenital and acquired forms of heart disease and arrhythmia (50, 51). Based on our observed dysfunction of INa in qv3J mice, we measured APs from isolated WT and qv3J cardiomyocytes. As observed in other vertebrates, AP duration (APD) in WT mouse cardiomyocytes decreased with increased pacing frequency (APD adaptation, Figure 6, A and C). Notably, WT cardiomyocytes displayed a parallel reduction in INa,p with increased pacing frequency (Supplemental Figure 9). Consistent with the measured decrease in INa,p in qv3J myocytes, qv3J myocyte APs (90% and 70% repolarization; APD90 and APD75, respectively) were significantly shorter at all pacing frequencies compared with WT myocytes (P < 0.05; APD50, P = NS; Figure 6, A–C and Supplemental Figure 10). Furthermore, qv3J myocytes displayed diminished capacity for APD adaptation (Figure 6C). These differences, as well as differences in INa,p, were eliminated by mexiletine but increased by isoproterenol (Figure 4, F–I, Figure 6D, and Supplemental Figure 10). These data indicate that disruption of CaMKII signaling through the βIV-spectrin pathway results in aberrant cardiomyocyte electrical activity.

Figure 6. βIV spectrin/CaMKII signaling complex is critical for normal cardiac cell membrane excitability.

Figure 6

(A and B) Representative APs and (C) APD90 in WT (n = 14) and qv3J (n = 16) cardiomyocytes paced at 4, 2, 1, or 0.5 Hz. *P < 0.05. (D) Differences in APD90 were eliminated by application of 50 μM mexiletine, but exaggerated by 1 μM isoproterenol (pacing frequency, 0.5 Hz). n = 14 (WT); 16 (qv3J). *P < 0.05. (E) Representative electrocardiograms recorded from Langendorff-perfused WT and qv3J hearts and summary data for (F) QRS duration (G), QT interval at 90% repolarization (QT90), and (H) RR interval for qv3J versus WT. n = 10 (WT); 10 (qv3J). *P < 0.05.

We used whole-animal and isolated heart electrocardiograms to test the effect of the βIV-spectrin/CaMKII pathway on in vivo cardiac function. qv3J mice showed decreased QRS duration (a marker of more rapid intraventricular conduction; refs. 50, 52), QT interval, and RR interval (increased heart rate) on both surface and isolated heart electrocardiograms (Figure 6, E–I). These findings were consistent with our observations of increased INa availability, decreased INa,p, and shorter APs in qv3J cardiomyocytes (Figure 4 and Figure 6, A–D).

We next investigated the effect of targeted disruption in CaMKII signaling on isoproterenol-induced cellular afterdepolarizations, which serve as important triggers for life-threatening cardiac electrical disturbances (53). At baseline, afterdepolarizations were infrequent in both WT and qv3J ventricular cardiomyocytes (20% and 10% of cells, respectively; n = 10; P = NS; Figure 7, A, B, and G). However, 1 μM isoproterenol significantly increased afterdepolarizations in WT but not qv3J cardiomyocytes (WT, 70%, P < 0.05 versus control; qv3J, 10%, P = NS versus control; n = 10 per group; Figure 7, B, E, and G). Mexiletine completely eliminated isoproterenol-induced afterdepolarizations (0% in WT and qv3J; P = NS; n = 5 per group; Figure 7, C, F, and G). Collectively, these data provide compelling evidence for the role of the βIV-spectrin/CaMKII signaling complex in normal myocyte physiology and suggest that modulation of this complex may represent a novel target to suppress persistent component of INa for arrhythmia treatment (Figure 7H).

Figure 7. βIV-spectrin/CaMKII complex regulates afterdepolarization formation in response to isoproterenol treatment.

Figure 7

Representative APs recorded from WT and qv3J cardiomyocytes at baseline (A and D) and following application of 1 μM isoproterenol (B and E) or isoproterenol applied with 10 μM mexiletine (C and F). (G) Summary data showing percent cells displaying afterdepolarizations in WT and qv3J at baseline (n = 10 per genotype) and following application of isoproterenol (n = 10 per genotype) or isoproterenol applied with mexiletine (n = 5 per genotype). *P < 0.05. (H) βIV-spectrin–based complex targets CaMKII to effector proteins in excitable cells.

The βIV-spectrin–dependent targeting pathway is conserved in excitable neurons.

While implicated in membrane biogenesis and maintenance, as well as ion channel clustering in the nervous system, we believe a role for βIV-spectrin in neuronal signaling is unprecedented. Based on our findings from heart, we tested the requirement of βIV-spectrin for local CaMKII organization in brain. In cerebellar Purkinje neurons, βIV-spectrin was localized to AISs with ankyrin-G (39, 40, 54) and associated with ankyrin-G in detergent-soluble cerebellar lysates (Figure 8A). In agreement with our findings from heart, CaMKII associated with βIV-spectrin in brain lysates (Figure 8B). Moreover, CaMKII was coexpressed with βIV-spectrin in Purkinje neuron AISs, but also localized to the neuronal cell body (Figure 8C). Consistent with our findings from heart, qv3J mice displayed aberrant targeting of CaMKII to the AIS (Figure 8C). Notably, CaMKII expression in the neuronal cell body was unaffected in qv3J cerebellum. In contrast, both ankyrin-G and βIV-spectrin were normally localized at qv3J AIS, although βIV-spectrin staining levels at the AIS were modestly decreased compared with WT cerebellum (Figure 8, D and E). Importantly, these findings indicate that the βIV-spectrin pathway has evolved to target CaMKII gene products to specific membrane domains in functionally diverse excitable cells.

Figure 8. βIV-spectrin targets CaMKII in neurons.

Figure 8

(A and B) βIV-spectrin associated with both ankyrin-G and CaMKII from mouse cerebellar lysates. (C) CaMKII (red) was observed at AISs (white arrows) in WT, but not qv3J, cerebellum. (D) AIS integrity (shown by positive ankyrin-G staining) was not markedly affected in qv3J mice. (E) βIV-spectrin (red) was present at qv3J mouse AISs, but at reduced levels compared with WT cerebellum. Calbindin (blue) was used to label the Purkinje cell body in CE. Scale bars: 10 μm (CE).

Discussion

Little is known about the cellular pathways that determine membrane excitability. Using a computational approach, we identified more than 30 putative CaMKII-binding proteins, including mitochondrial, nuclear, cytoskeletal, and membrane proteins. We performed comprehensive analyses that authenticated 1 candidate, βIV-spectrin, as a CaMKII-anchoring protein at select excitable membrane domains, including the Purkinje cell AIS and the cardiomyocyte intercalated disc. Notably, mice harboring a mutant βIV-spectrin lacking CaMKII binding activity showed defective membrane-specific CaMKII targeting, resulting in aberrant substrate activity, dysfunctional cellular excitability, and abnormal cardiac electrophysiology. Interestingly, our data also support the concept that modulating this complex may serve as an unexpected molecular target for arrhythmia therapy through suppression of late INa (i.e., INa,p; refs. 55, 56).

From bacteria to humans, regulated membrane excitability is central to life. Membrane ion flux drives salt and water balance, neurotransmitter release, myocyte contraction, and hormone secretion. Importantly, as recognized by Walter Cannon nearly a century ago (57), membrane excitability is remarkably fluid, constantly evolving to balance the relationship between an organism and its environment. Loss of ion homeostasis leads to disordered membrane excitability, which is a major factor in cardiac and neurological diseases. However, despite remarkable inroads into animal, tissue, and cellular physiology over the past century, arguably, the last major therapeutic breakthroughs in membrane excitability for human health and disease were the beta-blocker (discovered in the late 1950s; ref. 58), the ACE inhibitor (in the 1970s; ref. 59), and selective serotonin reuptake inhibitors (in the 1970s; ref. 60). Identification of new therapeutic agents to treat disorders of membrane excitability depends upon a more complete understanding of the cellular pathways regulating cell membrane excitability and organ function.

Previously identified CaMKII-binding partners have also served as direct substrates. For example, CaMKII associates with and phosphorylates the voltage-gated Ca2+ channel β2a subunit to modulate transverse-tubule L-type Ca2+ channel facilitation in cardiomyocytes (35). Similarly, direct binding of CaMKII with NR2B regulates NMDA receptor activity in vitro (61). In contrast, we postulate that βIV-spectrin, the CaMKII-targeting protein identified herein, will behave more analogous to AKAP proteins to organize the local signaling environment for multiple target substrates (Figure 8 and ref. 62). In fact, we predict that the βIV-spectrin qv3J mice will serve as an excellent in vivo model to identify and/or validate other potential CaMKII substrates at the cardiac intercalated disc or neuronal AIS.

Initial mRNA studies did not identify significant levels of βIV-spectrin in multiple tissues including heart and pancreas (39, 45, 63). Thus, our discovery of βIV-spectrin in the heart was initially surprising. However, βIV-spectrin has subsequently been identified at select membranes of pancreatic β cells, where βIV-spectrin was also not identified by whole-tissue mRNA analysis (39). Thus, our present findings of a strategic concentration of βIV-spectrin at the ventricular cardiomyocyte intercalated disc are consistent with the notion that this protein pathway has evolved as a master regulator of local excitable cell signaling in select cells of the brain, heart, and pancreas (albeit at significantly lower expression levels in heart compared with brain; Figure 2E). Furthermore, we predict that βIV-spectrin may also be localized to key membrane domains in other tissues to regulate local signaling pathways. For example, βIV-spectrin is localized to the neuromuscular junction, where it interacts with skeletal muscle Na+ channels (Nav1.4) and ankyrin-G (M.N. Rasband, unpublished observations).

Our present study revealed that βIV-spectrin exerted spatial control over CaMKII-dependent regulation of intercalated disc voltage-gated Na+ channels (targeted by βIV-spectrin–dependent interactions with ankyrin-G) and cell excitability. Independent studies from several groups demonstrate that cardiac Nav1.5 is highly concentrated at the vertebrate myocyte intercalated disc (between 80% and 95%; refs. 43, 6466). An important future goal will be to determine the mechanisms underlying regulation of this minor nonintercalated disc (i.e., transverse-tubule and peripheral sarcolemmal) population of Nav1.5 channels.

Our findings identify the βIV-spectrin/CaMKII complex as an essential component for INa regulation in the cardiomyocyte. In fact, qv3J myocytes with targeted disruption of this complex show abnormal INa inactivation properties, decreased INa,p, and altered cell excitability. Additionally, our findings illustrate that loss of this complex in myocytes does not affect other major cardiac ion currents, including total K+ current, ICa,L, or Ito. Nevertheless, our findings do not rule out the possibility that other targets of the βIV-spectrin/CaMKII complex are important in cardiomyocytes. In fact, we predict that other resident intercalated disc proteins (including ion channels, transporters, receptors, cytoskeletal proteins, and signaling molecules) may also be modulated by the βIV-spectrin/CaMKII complex and contribute to observed differences in APD75 and APD90 in qv3J myocytes. An important immediate future goal will be to identify additional targets for the βIV-spectrin/CaMKII complex, in both small and large animal models, that may yield unique insights into the relationship between CaMKII signaling and pathophysiology. For example, although both large and small animals show APD adaptation in response to changes in pacing frequency (Supplemental Figure 9 and ref. 67), INa,p is an important determinant of this behavior in mice, but the potential role of INa,p is less clear in larger animals, where APD adaptation is primarily attributed to K+ and/or Ca2+ channel activity (reviewed in ref. 67). Thus, we anticipate that studies in different animal models will be useful for studying all aspects of the βIV-spectrin/CaMKII signaling complex.

While our findings identify βIV-spectrin as a component of vertebrate heart, they also raise many exciting questions regarding the regulation of this complex gene and resulting gene products in heart. For example, while we have identified 2 isoforms in heart (Σ1 and Σ6), 6 βIV-spectrin splice forms are expressed in brain (29, 39, 63). Future experiments must address how these gene products are differentially expressed throughout development and across specific cardiac excitable cells, and whether other βIV-spectrin isoforms are found in heart. Moreover, while it is likely that each isoform has a unique role in heart (e.g., βIVS6-spectrin localizes Nav channels to AISs in neurons, ref. 28), the specific functions of βIV-spectrin splice variants in heart remain unknown. Based on conservation of the CaMKII- and ankyrin-G–binding motifs in Σ1 and Σ6 isoforms, we anticipate that both isoforms will be important for regulation of Nav channel activity and cell excitability. Finally, it is notable that we observed a small shift in the mobility of cardiac βIV-spectrin isoforms compared with their cerebellar βIV-spectrin counterparts (Figure 2E). While we hypothesize that this small shift may represent tissue-specific modification of residues found in both isoforms, additional experiments will be necessary to identify whether βIV-spectrin polypeptides display differential posttranslational modifications as well as to define how these potential modifications may differentially regulate βIV-spectrin function.

In addition to heart, βIV-spectrin also targets CaMKII to AISs in cerebellar Purkinje neurons (Figure 8). In both central and peripheral nervous systems, βIV-spectrin is critical for the biogenesis and maintenance of local membrane domains (32, 40, 68). Moreover, ankyrin-G recruits voltage-gated Na+ and K+ channels to neuronal membrane domains (69). Therefore, we speculate that βIV-spectrin and ankyrin-G coordinate coupling between CaMKII and brain ion channels to regulate neuronal function and coordinate the cell’s response to external stimuli (e.g., transmitted by increased catecholamine levels; Figure 7H). It is notable that, similar to βIV-spectrin qv3J mice and mice lacking ankyrin-G in the cerebellum, CaMKIIβ-deficient mice display defects in Purkinje cell excitability and pronounced ataxia (14). Moreover, human mutations in spectrin have been identified in patients with spinocerebellar ataxia, while genome-wide association studies have linked variability in ankyrin-G expression with bipolar disorder (70). Thus, compelling data from humans and animal models strongly support the βIV-spectrin–based signaling platform as a nodal point for regulating excitable cell function in health and disease.

Methods

Computational analysis of putative CaMKII binding partners and in vitro binding.

Novel CaMKII binding partners were identified by screening the human genome for proteins containing an 8–amino acid stretch ([LMI]-x-[RHK]-Q-[ED]-[ST]-x-[ED]) found in the CaMKII autoregulatory domain (35) using ScanProsite. Candidates were cloned from human tissue, and CaMKII-binding activity was assessed by in vitro binding assays using radiolabeled target proteins and activated CaMKII expressed in HEK cells and immunoprecipitated using monoclonal CaMKIIδ antibody (Santa Cruz) and TrueBlot anti-mouse Ig IP beads (eBioscience). Protein complexes were washed 5 times (50 mM Tris, pH 7.4; 1 mM EDTA; 1 mM EGTA; 500 mM NaCl; and 0.1% Triton X-100), resuspended in SDS sample buffer, and boiled for 5 minutes. Products were separated by SDS-PAGE. To ensure equal loading, gels were stained with Coomassie Blue. [35S]-labeled in vitro translated target proteins bound to CaMKII were detected by phosphorimaging.

Animals.

qv3J mice and WT littermates were obtained from Jackson Laboratories. All experiments were performed in mice 8 weeks of age. The present studies were reviewed and approved by the IACUC of the University of Iowa.

Molecular biology.

To generate cDNA to use for engineering fusion proteins, PCR primers were designed to TOPO-clone the C-terminal region of βIV-spectrin from the human heart cDNA library (BD Biosciences) and the reverse transcribed cDNA from human heart tissue, based on a sequence obtained from GeneBank. To generate constructs for in vitro transcription/translation, fusion protein expression, and binding assay, cDNAs were engineered in-frame into pcDNA3.1+ (Invitrogen) and pGEX6P1 (GE Healthcare). Vectors were completely sequenced. RNA was isolated and processed for RT-PCR as described previously, using oligo-dT to prime for cDNA production (71). Primers to identify βIV-spectrin transcripts in heart and brain include primers directed against the N terminus to recognize full-length βIVΣ1-spectrin in rat (forward, 5′-GGGATGGCTTTGTCCTCACCC-3′; reverse, 5′-GCCAAGAGTTCCAACTTCTCC-3′; nucleotides 271–1,226 in rat [βIV-spectrin-Spnb4; XM_218364], product shown in Figure 2) as well as C-terminal primers that recognize βIVΣ1-spectrin and βIVΣ6-spectrin in mouse (βIV-spectrin-Spnb4; NM_032610) and rat (forward, 5′-CCACGATCGAGAAACTCAAGG-3′; reverse, 5′-GGTGTCCGTCGTGTCCAACG-3′; not shown).

Biochemistry and imaging.

Adult heart immunoprecipitations were performed as described previously (72). The following antibodies were used for immunoblotting, immunoprecipitation, or immunostaining: antibodies against ankyrin-G (43), βIV-spectrin (N-terminal domain, specific domain, and C-terminal domain; provided by M. Komada, Tokyo Institute of Technology, Yokohama, Japan; or obtained from Santa Cruz or LSBio; refs. 29, 40), Nav1.5, and CaMKIIδ, CaMKIIδ (Santa Cruz), CaMKII-phospho-T287 (ABR), connexin43 (Invitrogen), N-cadherin (Abcam), β-catenin (BD Biosciences), actin (Sigma-Aldrich). Affinity-purified polyclonal antibody against phospho–Nav1.5 S571 was developed and validated against WT and S571A mutant Nav1.5 channels expressed in HEK cells. Adult cardiomyocytes and heart sections were isolated, immunostained, and imaged as described previously (43, 73).

In vitro binding.

CaMKIIδ was in vitro translated using the TNT Coupled Reticulocyte Lysate Systems (Promega) and labeled with [35S]-methionine. [35S]-labeled CaMKIIδ was incubated with CTP-P, a biotinylated peptide mimicking the putative kinase-binding domain in human βIV-spectrin, or CTP-C, a scrambled peptide, in 500 μl binding buffer (50 mM Tris, pH 7.4; 1 mM EDTA; 1 mM EGTA; 150 mM NaCl; 0.1% Triton X-100; and 1:1,000 protease inhibitor [Sigma-Aldrich]) overnight at 4°C. Protein complexes were washed 5 times (50 mM Tris, pH 7.4; 1 mM EDTA; 1 mM EGTA; 500 mM NaCl; and 0.1% Triton X-100), resuspended in SDS sample buffer, and boiled for 5 minutes. Products were separated by SDS-PAGE. To ensure equal loading, gels were stained with Coomassie Blue. [35S]-labeled in vitro translated CaMKIIδ that bound to the peptide was detected by phosphorimaging.

APs.

APs were recorded using the perforated (amphotericin B) patch-clamp technique at 36°C ± 1°C in Tyrode solution (bath). The pipette contained 130 mM potassium aspartate, 10 mM NaCl, 10 mM HEPES, 0.04 mM CaCl2, 2.0 mM MgATP, 7.0 mM phosphocreatine, 0.1 mM NaGTP, and 240 μg/ml amphotericin B, with the pH adjusted to 7.2 with KOH. APs were evoked by brief current pulses 1.5–4 pA, 0.5–1 ms. APD was assessed as the time from the AP upstroke to 50%, 75%, and 90% repolarization to baseline (APD50, APD75, and APD90, respectively).

Electrophysiology and electrocardiograms.

See Supplemental Methods and Supplemental Figure 11.

Statistics.

Data are presented as mean ± SEM. P values were assessed with a paired Student’s t test (2-tailed) or ANOVA, as appropriate, for continuous data. The Bonferroni test was used for post-hoc testing. The null hypothesis was rejected for P < 0.05.

Supplementary Material

Supplemental data
JCI43621sd.pdf (2MB, pdf)

Acknowledgments

We acknowledge support from the NIH (HL084583 and HL083422 to P.J. Mohler; HL079031, HL62494, and HL70250 to M.E. Anderson; NS044916 to M.N. Rasband; HL096805 to T.J. Hund), Pew Scholars Trust (to P.J. Mohler), and Fondation Leducq Award to the Alliance for Calmodulin Kinase Signaling in Heart Disease (to P.J. Mohler and M.E. Anderson). We thank M. Komada for anti–βIV-spectrin polyclonal antibody.

Footnotes

Conflict of interest: The authors have declared that no conflict of interest exists.

Citation for this article: J Clin Invest. 2010;120(10):3508–3519. doi:10.1172/JCI43621.

See the related Commentary beginning on page 3434.

References

  • 1.Frankland PW, O’Brien C, Ohno M, Kirkwood A, Silva AJ. Alpha-CaMKII-dependent plasticity in the cortex is required for permanent memory. Nature. 2001;411(6835):309–313. doi: 10.1038/35077089. [DOI] [PubMed] [Google Scholar]
  • 2.Karcher RL, et al. Cell cycle regulation of myosin-V by calcium/calmodulin-dependent protein kinase II. Science. 2001;293(5533):1317–1320. doi: 10.1126/science.1061086. [DOI] [PubMed] [Google Scholar]
  • 3.Lisman J, Schulman H, Cline H. The molecular basis of CaMKII function in synaptic and behavioural memory. Nat Rev Neurosci. 2002;3(3):175–190. doi: 10.1038/nrn753. [DOI] [PubMed] [Google Scholar]
  • 4.Nutt LK, et al. Metabolic regulation of oocyte cell death through the CaMKII-mediated phosphorylation of caspase-2. Cell. 2005;123(1):89–103. doi: 10.1016/j.cell.2005.07.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Zhang R, et al. Calmodulin kinase II inhibition protects against structural heart disease. Nat Med. 2005;11(4):409–417. doi: 10.1038/nm1215. [DOI] [PubMed] [Google Scholar]
  • 6.Boknik P, et al. Enhanced protein phosphorylation in hypertensive hypertrophy. Cardiovasc Res. 2001;51(4):717–728. doi: 10.1016/S0008-6363(01)00346-7. [DOI] [PubMed] [Google Scholar]
  • 7.Colomer JM, Mao L, Rockman HA, Means AR. Pressure overload selectively up-regulates Ca2+/calmodulin-dependent protein kinase II in vivo. Mol Endocrinol. 2003;17(2):183–192. doi: 10.1210/me.2002-0350. [DOI] [PubMed] [Google Scholar]
  • 8.Currie S, Smith GL. Enhanced phosphorylation of phospholamban and downregulation of sarco/endoplasmic reticulum Ca2+ ATPase type 2 (SERCA 2) in cardiac sarcoplasmic reticulum from rabbits with heart failure. Cardiovasc Res. 1999;41(1):135–146. doi: 10.1016/S0008-6363(98)00241-7. [DOI] [PubMed] [Google Scholar]
  • 9.Hagemann D, Bohlender J, Hoch B, Krause EG, Karczewski P. Expression of Ca2+/calmodulin-dependent protein kinase II delta-subunit isoforms in rats with hypertensive cardiac hypertrophy. Mol Cell Biochem. 2001;220(1-2):69–76. doi: 10.1023/A:1010899724222. [DOI] [PubMed] [Google Scholar]
  • 10.Hempel P, Hoch B, Bartel S, Karczewski P. Hypertrophic phenotype of cardiac calcium/calmodulin-dependent protein kinase II is reversed by angiotensin converting enzyme inhibition. Basic Res Cardiol. 2002;97 suppl 1:I96–I101. doi: 10.1007/s003950200037. [DOI] [PubMed] [Google Scholar]
  • 11.Hoch B, Meyer R, Hetzer R, Krause EG, Karczewski P. Identification and expression of delta-isoforms of the multifunctional Ca2+/calmodulin-dependent protein kinase in failing and nonfailing human myocardium. Circ Res. 1999;84(6):713–721. doi: 10.1161/01.res.84.6.713. [DOI] [PubMed] [Google Scholar]
  • 12.Mishra S, Sabbah HN, Jain JC, Gupta RC. Reduced Ca2+-calmodulin-dependent protein kinase activity and expression in LV myocardium of dogs with heart failure. Am J Physiol Heart Circ Physiol. 2003;284(3):H876–883. doi: 10.1152/ajpheart.00266.2002. [DOI] [PubMed] [Google Scholar]
  • 13.Netticadan T, Temsah RM, Kawabata K, Dhalla NS. Sarcoplasmic reticulum Ca(2+)/Calmodulin-dependent protein kinase is altered in heart failure. Circ Res. 2000;86(5):596–605. doi: 10.1161/01.res.86.5.596. [DOI] [PubMed] [Google Scholar]
  • 14.van Woerden GM, et al. betaCaMKII controls the direction of plasticity at parallel fiber-Purkinje cell synapses. Nat Neurosci. 2009;12(7):823–825. doi: 10.1038/nn.2329. [DOI] [PubMed] [Google Scholar]
  • 15.Silva AJ, Stevens CF, Tonegawa S, Wang Y. Deficient hippocampal long-term potentiation in alpha-calcium-calmodulin kinase II mutant mice. Science. 1992;257(5067):201–206. doi: 10.1126/science.1378648. [DOI] [PubMed] [Google Scholar]
  • 16.Wagner S, et al. Ca2+/calmodulin-dependent protein kinase II regulates cardiac Na+ channels. J Clin Invest. 2006;116(12):3127–3138. doi: 10.1172/JCI26620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Maier LS, Bers DM. Role of Ca2+/calmodulin-dependent protein kinase (CaMK) in excitation-contraction coupling in the heart. Cardiovasc Res. 2007;73(4):631–640. doi: 10.1016/j.cardiores.2006.11.005. [DOI] [PubMed] [Google Scholar]
  • 18.Aiba T, et al. Na+ channel regulation by Ca2+/calmodulin and Ca2+/calmodulin-dependent protein kinase II in guinea-pig ventricular myocytes. Cardiovasc Res. 2010;85(3):454–463. doi: 10.1093/cvr/cvp324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Bossuyt J, et al. Ca2+/calmodulin-dependent protein kinase IIdelta and protein kinase D overexpression reinforce the histone deacetylase 5 redistribution in heart failure. Circ Res. 2008;102(6):695–702. doi: 10.1161/CIRCRESAHA.107.169755. [DOI] [PubMed] [Google Scholar]
  • 20.Passier R, et al. CaM kinase signaling induces cardiac hypertrophy and activates the MEF2 transcription factor in vivo. J Clin Invest. 2000;105(10):1395–1406. doi: 10.1172/JCI8551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Zhu W, Woo AY, Yang D, Cheng H, Crow MT, Xiao RP. Activation of CaMKIIdeltaC is a common intermediate of diverse death stimuli-induced heart muscle cell apoptosis. J Biol Chem. 2007;282(14):10833–10839. doi: 10.1074/jbc.M611507200. [DOI] [PubMed] [Google Scholar]
  • 22.Ramirez MT, Zhao XL, Schulman H, Brown JH. The nuclear deltaB isoform of Ca2+/calmodulin-dependent protein kinase II regulates atrial natriuretic factor gene expression in ventricular myocytes. J Biol Chem. 1997;272(49):31203–31208. doi: 10.1074/jbc.272.49.31203. [DOI] [PubMed] [Google Scholar]
  • 23.Thiel WH, et al. Proarrhythmic Defects in Timothy Syndrome Require Calmodulin Kinase II. Circulation. 2008;118(22):2225–2234. doi: 10.1161/CIRCULATIONAHA.108.788067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Wu Y, et al. Suppression of dynamic Ca(2+) transient responses to pacing in ventricular myocytes from mice with genetic calmodulin kinase II inhibition. J Mol Cell Cardiol. 2006;40(2):213–223. doi: 10.1016/j.yjmcc.2005.11.005. [DOI] [PubMed] [Google Scholar]
  • 25.Khoo MS, et al. Death, cardiac dysfunction, and arrhythmias are increased by calmodulin kinase II in calcineurin cardiomyopathy. Circulation. 2006;114(13):1352–1359. doi: 10.1161/CIRCULATIONAHA.106.644583. [DOI] [PubMed] [Google Scholar]
  • 26.Chelu MG, et al. Calmodulin kinase II-mediated sarcoplasmic reticulum Ca2+ leak promotes atrial fibrillation in mice. J Clin Invest. 2009;119(7):1940–1951. doi: 10.1172/JCI37059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Kosaka T, Komada M, Kosaka K. Sodium channel cluster, betaIV-spectrin and ankyrinG positive “hot spots” on dendritic segments of parvalbumin-containing neurons and some other neurons in the mouse and rat main olfactory bulbs. Neurosci Res. 2008;62(3):176–186. doi: 10.1016/j.neures.2008.08.002. [DOI] [PubMed] [Google Scholar]
  • 28.Uemoto Y, et al. Specific role of the truncated betaIV-spectrin Sigma6 in sodium channel clustering at axon initial segments and nodes of ranvier. J Biol Chem. 2007;282(9):6548–6555. doi: 10.1074/jbc.M609223200. [DOI] [PubMed] [Google Scholar]
  • 29.Komada M, Soriano P. [Beta]IV-spectrin regulates sodium channel clustering through ankyrin-G at axon initial segments and nodes of Ranvier. J Cell Biol. 2002;156(2):337–348. doi: 10.1083/jcb.200110003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Koticha D, et al. Neurofascin interactions play a critical role in clustering sodium channels, ankyrin G and beta IV spectrin at peripheral nodes of Ranvier. Dev Biol. 2006;293(1):1–12. doi: 10.1016/j.ydbio.2005.05.028. [DOI] [PubMed] [Google Scholar]
  • 31.Dzhashiashvili Y, Zhang Y, Galinska J, Lam I, Grumet M, Salzer JL. Nodes of Ranvier and axon initial segments are ankyrin G-dependent domains that assemble by distinct mechanisms. J Cell Biol. 2007;177(5):857–870. doi: 10.1083/jcb.200612012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Yang Y, Lacas-Gervais S, Morest DK, Solimena M, Rasband MN. BetaIV spectrins are essential for membrane stability and the molecular organization of nodes of Ranvier. J Neurosci. 2004;24(33):7230–7240. doi: 10.1523/JNEUROSCI.2125-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Jenkins SM, Bennett V. Ankyrin-G coordinates assembly of the spectrin-based membrane skeleton, voltage-gated sodium channels, and L1 CAMs at Purkinje neuron initial segments. J Cell Biol. 2001;155(5):739–746. doi: 10.1083/jcb.200109026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Srinivasan Y, Elmer L, Davis J, Bennett V, Angelides K. Ankyrin and spectrin associate with voltage-dependent sodium channels in brain. Nature. 1988;333(6169):177–180. doi: 10.1038/333177a0. [DOI] [PubMed] [Google Scholar]
  • 35.Grueter CE, et al. L-Type Ca(2+) Channel Facilitation Mediated by Phosphorylation of the beta Subunit by CaMKII. Mol Cell. 2006;23(5):641–650. doi: 10.1016/j.molcel.2006.07.006. [DOI] [PubMed] [Google Scholar]
  • 36.Saurin AT, Durgan J, Cameron AJ, Faisal A, Marber MS, Parker PJ. The regulated assembly of a PKCepsilon complex controls the completion of cytokinesis. Nat Cell Biol. 2008;10(8):891–901. doi: 10.1038/ncb1749. [DOI] [PubMed] [Google Scholar]
  • 37.Luders J, Patel UK, Stearns T. GCP-WD is a gamma-tubulin targeting factor required for centrosomal and chromatin-mediated microtubule nucleation. Nat Cell Biol. 2006;8(2):137–147. doi: 10.1038/ncb1349. [DOI] [PubMed] [Google Scholar]
  • 38.Enomoto A, et al. Akt/PKB regulates actin organization and cell motility via Girdin/APE. Dev Cell. 2005;9(3):389–402. doi: 10.1016/j.devcel.2005.08.001. [DOI] [PubMed] [Google Scholar]
  • 39.Berghs S, et al. betaIV spectrin, a new spectrin localized at axon initial segments and nodes of ranvier in the central and peripheral nervous system. J Cell Biol. 2000;151(5):985–1002. doi: 10.1083/jcb.151.5.985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Lacas-Gervais S, et al. BetaIVSigma1 spectrin stabilizes the nodes of Ranvier and axon initial segments. J Cell Biol. 2004;166(7):983–990. doi: 10.1083/jcb.200408007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Backs J, et al. The delta isoform of CaM kinase II is required for pathological cardiac hypertrophy and remodeling after pressure overload. Proc Natl Acad Sci U S A. 2009;106(7):2342–2347. doi: 10.1073/pnas.0813013106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Zhou D, Lambert S, Malen PL, Carpenter S, Boland LM, Bennett V. AnkyrinG is required for clustering of voltage-gated Na channels at axon initial segments and for normal action potential firing. J Cell Biol. 1998;143(5):1295–1304. doi: 10.1083/jcb.143.5.1295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Mohler PJ, et al. Nav1.5 E1053K mutation causing Brugada syndrome blocks binding to ankyrin-G and expression of Nav1.5 on the surface of cardiomyocytes. Proc Natl Acad Sci U S A. 2004;101(50):17533–17538. doi: 10.1073/pnas.0403711101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Lowe JS, et al. Voltage-gated Nav channel targeting in the heart requires an ankyrin-G dependent cellular pathway. J Cell Biol. 2008;180(1):173–186. doi: 10.1083/jcb.200710107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Parkinson NJ, et al. Mutant beta-spectrin 4 causes auditory and motor neuropathies in quivering mice. Nat Genet. 2001;29(1):61–65. doi: 10.1038/ng710. [DOI] [PubMed] [Google Scholar]
  • 46.Salama G, London B. Mouse models of long QT syndrome. J Physiol. 2007;578(pt 1):43–53. doi: 10.1113/jphysiol.2006.118745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Wagner S, et al. Ca/calmodulin kinase II differentially modulates potassium currents. Circ Arrhythm Electrophysiol. 2009;2(3):285–294. doi: 10.1161/CIRCEP.108.842799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Li J, et al. Calmodulin kinase II inhibition shortens action potential duration by upregulation of K+ currents. Circ Res. 2006;99(10):1092–1099. doi: 10.1161/01.RES.0000249369.71709.5c. [DOI] [PubMed] [Google Scholar]
  • 49.Papadatos GA, et al. Slowed conduction and ventricular tachycardia after targeted disruption of the cardiac sodium channel gene Scn5a. Proc Natl Acad Sci U S A. 2002;99(9):6210–6215. doi: 10.1073/pnas.082121299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Bennett PB, Yazawa K, Makita N, George AL., Jr Molecular mechanism for an inherited cardiac arrhythmia. Nature. 1995;376(6542):683–685. doi: 10.1038/376683a0. [DOI] [PubMed] [Google Scholar]
  • 51.Tan HL, et al. A sodium-channel mutation causes isolated cardiac conduction disease. Nature. 2001;409(6823):1043–1047. doi: 10.1038/35059090. [DOI] [PubMed] [Google Scholar]
  • 52.Kanda M, et al. Electrophysiologic characteristics and implications of induced ventricular fibrillation in symptomatic patients with Brugada syndrome. J Am Coll Cardiol. 2002;39(11):1799–1805. doi: 10.1016/S0735-1097(02)01867-3. [DOI] [PubMed] [Google Scholar]
  • 53.Roden DM, Lazzara R, Rosen M, Schwartz PJ, Towbin J, Vincent GM. Multiple mechanisms in the long-QT syndrome. Current knowledge, gaps, and future directions. The SADS Foundation Task Force on LQTS. Circulation. 1996;94(8):1996–2012. doi: 10.1161/01.cir.94.8.1996. [DOI] [PubMed] [Google Scholar]
  • 54.Jenkins SM, Bennett V. Developing nodes of Ranvier are defined by ankyrin-G clustering and are independent of paranodal axoglial adhesion. Proc Natl Acad Sci U S A. 2002;99(4):2303–2308. doi: 10.1073/pnas.042601799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Jarecki BW, Piekarz AD, Jackson JO 2nd, Cummins TR. Human voltage-gated sodium channel mutations that cause inherited neuronal and muscle channelopathies increase resurgent sodium currents. J Clin Invest. 2010;120(1):369–378. doi: 10.1172/JCI40801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Koopmann TT, Bezzina CR, Wilde AA. Voltage-gated sodium channels: action players with many faces. Ann Med. 2006;38(7):472–482. doi: 10.1080/07853890600969072. [DOI] [PubMed] [Google Scholar]
  • 57.Cannon WB. Bodily Changes in Pain, Hunger, Fear, and Rage: An Account of Recent Researches into the Function of Emotional Excitement. New York, New York, USA: D. Appleton. & Co.; 1915. [Google Scholar]
  • 58.Black JW, Stephenson JS. Pharmacology of a new adrenergic beta-receptor-blocking compound (Nethalide). Lancet. 1962;2(7251):311–314. doi: 10.1016/s0140-6736(62)90103-4. [DOI] [PubMed] [Google Scholar]
  • 59.Haber E, Barger AC. Experimental and clinical studies with converting enzyme inhibitor. Prog Biochem Pharmacol. 1976;12:16–32. [PubMed] [Google Scholar]
  • 60.Wong DT, Horng JS, Bymaster FP, Hauser KL, Molloy BB. A selective inhibitor of serotonin uptake: Lilly 110140, 3-(p-trifluoromethylphenoxy)-N-methyl-3-phenylpropylamine. Life Sci. 1974;15(3):471–479. doi: 10.1016/0024-3205(74)90345-2. [DOI] [PubMed] [Google Scholar]
  • 61.Barria A, Malinow R. NMDA receptor subunit composition controls synaptic plasticity by regulating binding to CaMKII. Neuron. 2005;48(2):289–301. doi: 10.1016/j.neuron.2005.08.034. [DOI] [PubMed] [Google Scholar]
  • 62.Wong W, Scott JD. AKAP signalling complexes: focal points in space and time. Nat Rev Mol Cell Biol. 2004;5(12):959–970. doi: 10.1038/nrm1527. [DOI] [PubMed] [Google Scholar]
  • 63.Tse WT, et al. A new spectrin, beta IV, has a major truncated isoform that associates with promyelocytic leukemia protein nuclear bodies and the nuclear matrix. J Biol Chem. 2001;276(26):23974–23985. doi: 10.1074/jbc.M009307200. [DOI] [PubMed] [Google Scholar]
  • 64.Allouis M, et al. 14-3-3 is a regulator of the cardiac voltage-gated sodium channel Nav1.5. Circ Res. 2006;98(12):1538–1546. doi: 10.1161/01.RES.0000229244.97497.2c. [DOI] [PubMed] [Google Scholar]
  • 65.Maier SK, Westenbroek RE, McCormick KA, Curtis R, Scheuer T, Catterall WA. Distinct subcellular localization of different sodium channel alpha and beta subunits in single ventricular myocytes from mouse heart. Circulation. 2004;109(11):1421–1427. doi: 10.1161/01.CIR.0000121421.61896.24. [DOI] [PubMed] [Google Scholar]
  • 66.Maier SK, et al. An unexpected requirement for brain-type sodium channels for control of heart rate in the mouse sinoatrial node. Proc Natl Acad Sci U S A. 2003;100(6):3507–3512. doi: 10.1073/pnas.2627986100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Eisner DA, Dibb KM, Trafford AW. The mechanism and significance of the slow changes of ventricular action potential duration following a change of heart rate. Exp Physiol. 2009;94(5):520–528. doi: 10.1113/expphysiol.2008.044008. [DOI] [PubMed] [Google Scholar]
  • 68.Yang Y, Ogawa Y, Hedstrom KL, Rasband MN. betaIV spectrin is recruited to axon initial segments and nodes of Ranvier by ankyrinG. J Cell Biol. 2007;176(4):509–519. doi: 10.1083/jcb.200610128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Hill AS, et al. Ion channel clustering at the axon initial segment and node of ranvier evolved sequentially in early chordates. PLoS Genet. 2008;4(12):e1000317. doi: 10.1371/journal.pgen.1000317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Ferreira MA, et al. Collaborative genome-wide association analysis supports a role for ANK3 and CACNA1C in bipolar disorder. Nat Genet. 2008;40(9):1056–1058. doi: 10.1038/ng.209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Gudmundsson H, et al. EH domain proteins regulate cardiac membrane protein targeting. Circ Res. 2010;107(1):84–95. doi: 10.1161/CIRCRESAHA.110.216713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Mohler PJ, Davis JQ, Bennett V. Ankyrin-B Coordinates the Na/K ATPase, Na/Ca Exchanger, and InsP(3) Receptor in a Cardiac T-Tubule/SR Microdomain. PLoS Biol. 2005;3(12):e423. doi: 10.1371/journal.pbio.0030423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Mohler PJ, et al. Ankyrin-B mutation causes type 4 long-QT cardiac arrhythmia and sudden cardiac death. Nature. 2003;421(6923):634–639. doi: 10.1038/nature01335. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental data
JCI43621sd.pdf (2MB, pdf)

Articles from The Journal of Clinical Investigation are provided here courtesy of American Society for Clinical Investigation

RESOURCES