Abstract
David T. Yue was a renowned biophysicist who dedicated his life to the study of Ca2+ signaling in cells. In the wake of his passing, we are left not only with a feeling of great loss, but with a tremendous and impactful body of work contributed by a remarkable man. David's research spanned the spectrum from atomic structure to organ systems, with a quantitative rigor aimed at understanding the fundamental mechanisms underlying biological function. Along the way he developed new tools and approaches, enabling not only his own research but that of his contemporaries and those who will come after him. While we cannot hope to replicate the eloquence and style we are accustomed to in David's writing, we nonetheless undertake a review of David's chosen field of study with a focus on many of his contributions to the calcium channel field.
Keywords: biophysics, calcium channel, calmodulin, protein kinase A, single channels, sodium channel
Dream + Wonder
It is 2:00 AM on a typical Tuesday night in Baltimore, Maryland. While many are enjoying a peaceful night of sleep, the lights are still on in Room 713 of the Richard Starr Ross Research Building at the Johns Hopkins University School of Medicine. One can distinctly hear the drumming of the hard drive on an antiquated computer recording digitized electric currents as it has faithfully done so since 1988. The nearby oscilloscope vividly displays a neon green trace – the picoampere flickers of a single Ca2+ ion channel gating open and shut. What ensues is a moment of silence, then anticipation, and suddenly a distinct joyous laughter erupts. It is as though the clouds that had shrouded a mystery of nature had parted and the blue sky beyond became apparent – Dr. David T. Yue had a deep insight on the mechanism by which Ca2+ channels functioned. He would soon walk over to the blackboard with you and derive a mathematically precise explanation of the observed phenomenon. To his students, trainees, and colleagues at the Calcium Signals Laboratory, such moments of unbridled joy and wonder were a constant inspiration to march forward in the “pursuit of truth, ” and a glimmer of hope that they may yet be able to complete their training.
David T. Yue graduated magna cum laude from Harvard University in 1979 with a Bachelor of Arts in Biochemistry. At Harvard, he was influenced by many prominent scientists including Edward Purcell, who taught him the beauty and elegance of Maxwell's equations. He became captivated by physics and the power of quantitative sciences to explain nature and its many mysteries. For his senior thesis, he worked with J.W. Hastings, a pioneer in bioluminescence. Perhaps as a result of these early influences, he found himself in doubt as he transitioned to become an M.D.-Ph.D student at Johns Hopkins University School of Medicine. “I'm not quite sure I'll ever find the right field. I'll always want to be a bit mathematical, a bit physical – and have the same relevancy of a biologist, ” he wrote to his best friend from high school at the end of his first year of medical school. He continues, “What I hope to do after acquainting myself with the field of battle is to develop a biophysics angle on the situation – a tall order, a definite maybe, and probably a naïve dream. But I'm banking on it.” Herein began his ultimate quest to find “a molecular problem with physics + math ∼ biophysics.”
The first leg of this journey was in the laboratory of Kiichi Sagawa, an eminent cardiovascular physiologist and a Professor of Biomedical Engineering at Johns Hopkins University. The Sagawa laboratory was a scientific environment that encouraged independent thinking. David with his close friends Dan Burkhoff and Joe Alexander soon conjured up their own research projects. They began characterizing force-interval relationships in the heart in order to understand alterations in the strength of cardiac contraction as a result of brief changes in stimulation patterns.1-3 Through elegant mathematical formulation of these initial macroscopic studies it became apparent that there may be a common underlying feedback loop that related to fundamental properties of excitation-contraction coupling and the dynamics of Ca2+ ions in a single myocyte. This negative feedback loop seemed to suggest that the Ca2+ influx through the Ca2+ channel in the cell depended on the cytosolic Ca2+ in the cell.3 The biophysicist in him became ecstatic and translating the tissue-level phenomena and mathematical concepts into the ground molecular reality became his next mission. The vertically integrated approach to interpret systems level physiology through a molecular lens became a recurring theme throughout his career.
With encouragement from his mentor Kiichi Sagawa to pursue his molecular dream, David moonlighted in Gil Wier's Laboratory a few blocks across town at the University of Maryland, Baltimore to measure steady-state [Ca2+]-tension relations in intact cardiac muscle – a fundamental parameter in cardiac excitation-contraction coupling.4-6 This was the early days of intracellular Ca2+ measurements before the popular Ca2+-sensing dyes from the Roger Tsien laboratory7 became widely available and the possibility of genetically engineered Ca2+-sensing proteins was more like science fiction. Instead, Yue with Eduardo Marban and Gil Wier utilized pressure injection of aequorin, a Ca2+-sensitive bioluminescent protein derived from jelly fish, into a superficial layer of papillary muscle, and measuring changes in luminescence using a photomultiplier tube (Fig. 1).5 Aequorin possessing a relatively low affinity for Ca2+ and fast on-kinetics was particularly apt for such state-of-the-art measurements.8 By integrating mathematical modeling and meticulous luminescence measurements, Yue and colleagues demonstrated an unusually steep relationship between tension and intracellular [Ca2+] concentrations (Fig. 1C)5 – a result fundamental to our current understanding of cardiac excitation-contraction coupling.9
Now fully orbed into molecular level physiology, David Yue faced yet another predicament – a challenge that nearly all of us encounter at some juncture in our careers. He felt conflicted about what direction to steer his life toward even though he had sent in his residency applications. He loved research and teaching, but knew it was hard to make it in science. The Director of Biomedical Engineering, Dr. Richard Johns, had smiled when David asked for a faculty position while still a student, telling him to try again after residency. Nonetheless, after much introspection, David decided that he needed to trust in his growing faith and move in the direction of his dreams, against the advice of his parents and his advisors. To our great benefit, he withdrew his residency applications and instead dedicated himself to quantitative biophysical research. When Dr. Johns saw that David had dedicated himself solely to research, he offered him the faculty position in Biomedical Engineering that David had requested only a few months earlier. David accepted, deferring for a year so that he could join the Marban lab as a postdoc in order to bootstrap his research without faculty obligations.
In the late 1980's, single channel recording of ion channels had become an increasingly popular methodology to obtain high-resolution biophysical understanding of these molecules-nature's version of a transistor.10 Richard Aldrich, David Corey, and Charles Stevens had utilized such single-channel recordings from Na channels in neuroblastoma cell lines to revise the canonical Hodgkin-Huxley model of channel gating (ACS model)11 – a study that deeply influenced David's thinking. David thus began his rendezvous with ion channels as a post-doctoral fellow in the Marban laboratory, identifying a novel cardiac potassium channel12 as well as studying the permeation13, 14 and gating of voltage-gated sodium and calcium channels.15, 16 These studies invariably incorporated mathematical models in the form of discreet and continuous-time Markov processes to explain the biophysics of these channels. Over the next quarter century, David would devote significant time and effort toward incorporating this mode of thinking into a popular graduate-level course entitled “Ion Channels of Excitable Membranes.”
We reminisce in this review an unapologetically biased account of nearly 3 decades of research by David T. Yue, and his Calcium Signals Laboratory that blends atomic structure, single-channel biophysics, cellular physiology, and systems neuroscience to understand the innumerable consequences that a single ion, Ca2+, has on human physiology, pathophysiology, and life as we know it.
The Importance of Ca2+
Calcium constitutes a vital signaling system critical for normal physiological function. Beyond its importance in making stronger bones, the biological roles of Ca2+ remained largely unknown until late 19th century when British physiologist Sidney Ringer characterized Ca2+ as an essential component required for the contractility of frog ventricles.17 More than a century of research since has confirmed the role of Ca2+ as ubiquitous second messenger critical for a vast number of signal transduction pathways. For David Yue, the precise regulation of these Ca2+ signals in a cell was akin to the feedback control of electrical circuits (Fig. 2). Thus as cardiac, immune, motor and neuronal systems each rely on Ca2+ signaling; understanding the detailed components of this biological circuit would have enormous physiological relevance. Here we mention just a few key roles for Ca2+, illustrative of its importance across multiple systems.
Figure 1.
David Yue and his early work in the role of Ca2+ in excitation-contraction coupling. (A, B) Stress and Ca2+ transient measurements in cardiac muscle. (C) The relationship between intracellular Ca2+ and relative stress in cardiac muscle. Reproduced from Yue et al., 1986. © Rockefeller University Press. Reproduced by permission of Rockefeller University Press. Permission to reuse must be obtained from the rightsholder.
Figure 2.

Ca2+ signals as an electrical circuit. View of cellular Ca2+ signaling as a feedback control pathway in an electrical circuit (modified from a slide created by David T. Yue, unpublished).
The importance of Ca2+ in the heart is clear and thus constitutes the first system upon which David Yue focused his attention.12, 15, 16, 18, 19 Upon depolarization, Ca2+ enters the cardiac myocyte through voltage-gated L-type Ca2+ channels in the cardiac dyad. This Ca2+ influx triggers Ca2+ induced Ca2+ release from the sarcoplasmic reticulum (SR) via Ca2+ binding to the ryanodine receptor.20 This release of Ca2+ from intracellular stores is in turn responsible for cardiac contraction via Ca2+ binding to troponin C.21 Such dependence of cardiac excitation-contraction coupling on Ca2+ illustrates the critical nature of Ca2+ signaling, but it does not represent the full set of pathways dependent on Ca2+ within the heart. Ca2+ entry through L-type channels is also a major factor in setting the action potential duration of the heart, 22 making physiological regulation of these channels a necessity16 as well as an important substrate for pathology.22, 23 Additionally, Ca2+ activates a number of kinases and phosphatases including CaMKII and calcineurin, both of which may play an important role in cardiac hypertrophy and heart failure.24, 25
Excitation-contraction coupling is not only a vital component of cardiac function, but is a fundamental mechanism of muscle contraction. In skeletal muscle, the role of the voltage gated calcium channel in contraction begins at the neuromuscular junction, where Ca2+ entry at the presynaptic terminal is required for neurotransmitter release and subsequent depolarization of the muscle fiber.26 The subsequent opening of L-type Ca2+ channels directly couples to ryanodine receptors, inducing a release of Ca2+ from the SR leading to muscle contraction.27
Lastly, Ca2+ signaling is a major player in neuronal function. Synaptic transmission is dependent on Ca2+ entry through voltage gated Ca2+ channels in the presynaptiC-terminal. Ca2+ entry through P/Q, N, and R-type channels mediate synaptic vesicle fusion, placing neurotransmitter release under the control of Ca2+ signaling.28 Furthermore, Ca2+ oscillations underlie the rhythmic activity of pacemaker neurons found in various neuronal networks, driving diverse processes from circadian rhythms to precise control of motor movement.29, 30 Finally, Ca2+ is a critical requirement of both short and long term forms of synaptic plasticity, 31, 32 making Ca2+ signaling an important mediator for learning and memory.
With such diverse and widespread roles, Ca2+ signaling is clearly a critical component of a wide variety of biological functions and a requirement of life itself. It is no wonder that David Yue, a scientist filled with passion and wonder, devoted his career to studying the precise regulation of the Ca2+ signals within cells. A deeper biophysical understanding of voltage gated Ca2+ channels and their regulatory mechanisms informs both normal and pathological states of the immune, cardiac, neuronal, and motor systems. The generation of cloned CaV channels in the 1980's33 fostered the growth of the calcium channel field. Thus when David Yue began his lab in 1987, a myriad of rich biophysical questions about Ca2+ and its entry through voltage gated calcium channels lay before him. The wonder of these biological transistors10 captured his attention and motivated the next decades of his research.
Ca2+ Regulatory Mechanisms
Among all the mysteries of voltage-gated Ca2+ channels (CaV), David Yue was most fascinated by one peculiar phenomenon – Ca2+-dependent inactivation (CDI). First described by Brehm and Eckert in 1978 in Paramecium, 34 CaV channels were found to inactivate in response to elevated intracellular Ca2+ concentration. The discovery of this feedback inhibition led to the hypothesis that CDI played a prominent role in the homeostatic control of Ca2+ levels within the cell. Further investigations revealed that CDI was not restricted to Paramecium, but extended to a myriad of systems including mammalian cardiac and neuronal cell types.35 While this phenomenon was replicated by several groups using whole cell patch clamp recordings, 35, 36, 37 the existence of CDI at the single channel level remained elusive.38, 39 Driven by the desire to reconcile whole cell and single channel results, David Yue and John Imredy, set out to record single CaV channel currents in 1990 in search of CDI on the molecular level.40
Before the Ca2+ dependence of CDI could be proven, single channel records needed to be obtained in the absence of Ca2+. Ba2+ was therefore used as a charge carrier since it fluxes well through CaV channels without eliciting Ca2+ dependent processes. Single channel records of Ba2+ currents through the cardiac L-type calcium channel demonstrated robust openings in response to a voltage step (Fig. 3A, left). Ensemble average currents displayed only a modest decline in current during the step, indicative of a voltage dependent inactivation (VDI) process (Fig. 3A, bottom left). However, when Ca2+ was used as the charge carrier in these same L-type channels, a significant decrease in channel gating was evident (Fig. 3A, right). Channel openings were not only reduced, but the ensemble average current displayed a marked decay in Ca2+ current during the voltage step indicating robust CDI at the single channel level (Fig. 3A, bottom right). This change in channel gating was later shown to be the result of a true shift in channel gating modes, whereby Ca2+ entry through the channel promotes a transition from a high open probability mode of gating (mode 1) to a low probability mode (mode Ca).41
Figure 3.

CaV regulatory mechanisms. (A) Single channel recordings of LTCCs unequivocally demonstrate CDI, a reduction in open probability as reflected in the ensemble average in the bottom row of the right column (Ca2+ as a charge carrier) as opposed to the left column (Ba2+ as a charge carrier). Reproduced from Yue, Backx and Imredy, 1990.40 © AAAS. Reproduced by permission of AAAS. Permission to reuse must be obtained from the rightsholder. (B, C) Two forms of calmodulation in CaV2.1. The N-lobe of CaM orchestrates CDI (B) while the C-lobe coordinates CDF (C). Reproduced from DeMaria et al., 2001.
At the whole cell level, this CDI manifests simply as the faster decay of the Ca2+ current when recorded with Ca2+ as the charge carrier, as compared to Ba2+, as demonstrated in recombinant CaV 2.1 channels42 (Fig. 3B). The extent of CDI can be quantified by plotting the fraction of peak current remaining at the end of the test pulse as a function of voltage (Fig. 3B, right). The U-shape of this curve provides a hallmark of Ca2+ dependent regulation.35, 43 Remarkably, these same channels also exhibit a second type of Ca2+ dependent regulation. When shorter voltage pulses, which initiate a negligible amount of CDI, are used to probe the channel, a facilitation of the Ca2+ current can be seen (Fig. 3C). This can be quantified by applying a pre-pulse which allows the channel to prefacilitate prior to the test pulse. This duel regulation of the CaV2.1 channel reveals a growing theme of Ca2+ dependent regulation of CaV channels whereby Ca2+ entry into a cell can induce disparate modes of channel regulation.
Identifying the sensor for CDI became a subject of intense study within the field. Multiple theories were proposed including Ca2+ induced channel (de)phosphoylation, 44 direct binding of Ca2+ to the channel α subunit, 43, 45 or calmodulin46 induced channel regulation.47 Chimeric channel analyses and site-directed mutagenesis identified critical components of CDI within the carboxyl-tail of CaV1.2, 43, 48 but identifying the key component of CDI would require a new tool. In 1998, the Adelman lab determined that CaM was the Ca2+ sensor for small conductance potassium channels49 by utilizing a form of CaM in which all 4 EF-hand domains were mutated (CaM1234). This novel tool helped the Yue lab to definitively identify CaM as the Ca2+ sensor for CDI in the L-type Ca2+ channel, 50 a result which extends to CaV2.1 channels as well.51, 52 In CaV2.1channels, both CDI and CDF were completely eliminated by the overexpression of CaM1234 in HEK cells (Fig. 4A), validating CaM as the Ca2+ sensor and indicating a pre-association of the channel with apoCaM.42 Further, expression of mutant CaMs in which only the C lobe is able to bind Ca2+ (CaM12) specifically abolishes CDI in CaV2.1, while a mutant CaM with N lobe Ca2+ binding eliminates Ca2+-dependent facilitation (CDF) (Fig. 4A). Thus the dual regulation of CaM arises as distinct regulatory mechanisms triggered independently by the 2 lobes of CaM.42, 52-56 This functional bipartition of Ca2+ regulation by CaM (calmodulation) appears to be a general rule across channel subtypes (Fig. 4C).42, 50, 53, 54, 57-61
Figure 4.
CaM is the Ca2+ sensor for CaV channels. (A) Simple yet elegant tools, mutations of Ca2+-binding residues in one or both lobes of CaM, used to dissect Ca2+ regulation by each lobe of CaM. Reproduced from DeMaria etal., 2001.42 (B) Usage of various Ca2+ buffers differentiates global and local Ca2+ signals. Modified from Dick etal, 2008.64 (C) Summary of calmodulation in various ion channels.
In addition to bifurcating the Ca2+ response, CaM has a remarkable capability to discriminate the spatial origin of a Ca2+ signal. Such spatial selectivity is revealed by varying the intracellular Ca2+ buffering conditions. Strong buffering restricts Ca2+ fluctuations to the nanodomain at the mouth of the CaV channel62, 63 (Fig. 4B). This local Ca2+ signal is sufficient to allow C-lobe calmodulation, while the N-lobe often requires a global elevation in Ca2+, observed only in the presence of low intracellular Ca2+ buffering.42, 50, 53, 59, 64, 65 Thus C-lobe calmodulation permits signaling within a single CaM/channel complex without interference from neighboring Ca2+ sources, while N-lobe calmodulation allows coordination of Ca2+ regulation across distances within a cell. Such spatial selectivity rationalizes the early challenges in studying CDI in single channels where only a local Ca2+ signal exists. Indeed, CaV2.1 channels, described as having no CDI at the single channel level (an N-lobe mediated effect), have been shown to exhibit robust C-lobe triggered CDF at the single channel level.66
The general principle of the C-lobe as a local Ca2+ sensor and the N-lobe as a global Ca2+ sensor remained unquestioned until the Yue lab focused more closely on the CDI of CaV1.2/1.3. Remarkably, both the N and C-lobe of CaM functioned as local Ca2+ sensors within each of these channels.64, 65 This switch in the spatial selectivity of the N-lobe of CaM was shown to be the result of an additional CaM binding site within the amino-terminus of CaV1.2/1.3 channels called NSCaTE (N-terminal spatial Ca2+-transforming element). The presence of NSCaTE tuned the spatial selectivity of CaM, augmenting the repertoire of Ca2+ decoding available to channels. Further, the modularity of NSCaTE allowed for careful deduction of the mechanisms underlying the spatial selectivity of CaM lobes leading to a unified mechanism for Ca2+/CaM decoding across the CaV channel family (Fig. 4C).
Calmodulation of CaV channels has emerged as prominent regulatory scheme with extraordinary modes of behavior. But such CaM mediated regulation is not limited to CaV channels. Both the SK and KCNQ potassium channels associate with apoCaM and are Ca2+ regulated49, 67-69 Cyclic nucleotide-gated channels, 70, 71 NMDA receptors72, 73 and transient receptor potential (TRP) channels add to this growing list of channels proposed to undergo CaM regulation.74-78 Finally, the Ca2+ dependence of voltage gated sodium channels has long been proposed, but the mechanism and extent of such modulation has been controversial.79-85 Nonetheless, robust CDI was confirmed in NaV1.4 channels by the rapid delivery of Ca2+ to the channels either via photo-uncaging or influx though nearby engineered Ca2+ channels.86 Furthermore, CaM was shown to mediate this NaV CDI, primarily though Ca2+ binding to its N-lobe. Importantly, the structural similarities between the carboxyl tail of CaV and NaV channels proved to provide functional overlap, unifying the mechanism of calmodulation across the 2 channel families. Thus the principles deduced for CaM regulation of CaV channels is paralleled in the NaV channel family, and may well generalize to a multitude of channels, many yet to be discovered.
Structural Perception Changes the Meaning of Calcium Channel Regulation
David loved to describe voltage-gated Ca2+ channels as “biological transistors.” However, these “biological transistors” differ from silicon transistors in one key manner: while the components and details of silicon transistors are known to their human designers, the inner-workings of voltage-gated Ca2+ channels are shrouded in mystery. As such, unveiling the molecular machineries which govern Ca2+ regulation of voltage-gated Ca2+ channels was of intense interest to David, and a large body of his work was dedicated to deducing the mechanism by which parts of the channel interact to induce Ca2+ regulation. Utilizing biochemical, optical, and crystallographic means, David contributed significantly to the structural understanding of the voltage-gated Ca2+ channels.
Calcium channels are multi-subunit complexes with 3 main components: a pore-forming α1 subunit, a cytoplasmic β subunit, and a membrane anchored α2δ subunit. These subunits assemble to form the molecular machinery for Ca2+ entry into the cell while CaM serves as the actual Ca2+ sensor for feedback regulation. The pore forming α1 subunit is central for understanding the gating mechanisms of the channel. In particular, the carboxyl terminus of the α1 subunit was found to harbor a number of critical elements required for Ca2+ regulation of CaV channels. Two vestigial EF hand motifs were shown to be important components of CDI and CDF via analyses of mutant, chimeric, and alternatively-spliced channels.43, 48, 87 Further, the carboxyl-terminus also contains an IQ domain, which has long been a recognized apoCaM binding motif, 88 a result confirmed for the IQ domain of CaV channels.89-92 Additionally, the amino-terminus of a subset of CaV channels contains the Ca2+/CaM binding site, NSCaTE, which is capable of modulating the inactivation profile of CaV channels.64, 65
How do these key channel segments interact with CaM so as to produce Ca2+ regulation of the channel? To answer this, we begin by considering the typical resting state of the channels. In 2001, the Yue lab developed a novel quantitative FRET 2-hybrid method known as 33 FRET and used it to demonstrate that in the Ca2+ free state, apoCaM pre-associates with CaV channels.89, 90 The C-lobe of apoCaM has been shown to bind to the channel IQ domain, while the N-lobe is now known to bind to the channel vestigial EF-hand segments (Fig. 5A).90, 91, 93, 94 Although definitive crystal structures of apoCaM/channel interactions remain elusive, ab-initio modeling based on functional data provides a structural basis for a mechanistic model (Fig. 5). Traditionally, the structural basis for Ca2+ regulation of Ca2+ channels was explained by the IQ-centric model. In this model, pre-association of apoCaM on the channel carboxyl-terminus positions CaM as a resident Ca2+ sensor and primes the channel for Ca2+ regulation. As cytosolic Ca2+ rises, the pre-associated CaM binds Ca2+ and repositions on the channel inducing CDI or CDF. Although the channels featured a different gating mode corresponding to CDI or CDF, CaM was thought to remain in contact with the IQ region. This IQ centric model was supported by the high affinity of Ca2+/CaM with the IQ domain42, 50, 54, 61, 91-96 as well as atomic structures of Ca2+/CaM in complex with the IQ domain.97-101 However when the Yue lab presented their crystal structures of Ca2+/CaM in complex with the IQ domain of CaV2.1 and CaV2.3, they noted that the structure was not fully reconciled with the functional data, suggesting that Ca2+ /CaM may in fact depart from the IQ region during Ca2+ regulation.99 This contrary hypothesis was eventually confirmed using exhaustive alanine scanning and a novel analysis utilizing individually transformed Langmuir (iTL) relations.93, 94 In this model, the Ca2+ bound C-lobe of CaM forms a tri-partite complex with the IQ domain and the vestigial EF-hands within the CI region of the channel (Fig. 5, state I).93 While departing from a pure IQ centric model, this tri-partite arrangement is in agreement with early experiments indicating the importance of the IQ domain and vestigial EF-hand segments.43, 50, 102 The Ca2+ bound N-lobe of CaM also appears to depart from the IQ domain. For channels containing an NSCaTE motif on the N-terminus, it is likely that the N-lobe of Ca2+ CaM interacts with this domain, perhaps bridging the amino and carboxyl termini of the channel (Fig. 5).64, 65, 93
Figure 5.
Structural elements of CaV channels. Top: homology model demonstrating dynamic switching of CaM interaction with CaV1.3. Bottom: conceptual framework representing the 3 states of CaM regulation of the channel. Modified from Ben Johny et al.93 and Adams et al.109
Thus a coherent story of the structural rearrangements underlying Ca2+ regulation has emerged, but the Yue lab continued to delve deeper into the mystery, revealing yet another unexpected result. Previously, the high affinity of apoCaM for the channel and the ability of overexpressed mutant CaM to change the Ca2+ regulatory profile of the channel led to the conclusion that at rest, all channels must contain a preassociated apoCaM.89, 91 In 2010, however, the Yue lab provided evidence that this may not, in fact, be the case. By combining patch clamp and optical recordings with quantitative biochemical analysis, Yue and colleagues determined that some variants of CaV1.3/1.4 possess a module (ICDI – inhibitor of CDI) within their carboxyl tail which acts as a competitive inhibitor for apoCaM, thus introducing a channel state in which apoCaM is no longer preassociated (Fig. 5, state E) and therefore cannot undergo CDI.57, 93, Moreover, RNA-editing of CaV1.3 channels in the brain naturally modulates apoCaM binding affinity, 57, 58, 94, 103, 104 tuning the amount of CDI supported by the channel variant.94, 105-108 As increased CaM levels restore these CaM-less channels to configuration A by mass action, natural fluctuations in the levels of CaM may determine the distribution of channels among the E and A states, thus conferring physiological variations in CDI.94, 103, 108, 109 Finally, apoCaM itself has been found to have a profound effect on CaV channels. Recent studies using single channel recordings and chemical dimerization methods demonstrate that configuration E represents a gating mode with a low open probability (PO).109 Hence a single mechanistic scheme is proposed to underlie natural variations in channel PO and CDI: 1) channels in the E state lack a CaM and have low PO, 2) channels transition to the A state upon binding of apoCaM and exhibit a high PO and 3) with the addition of Ca2+, structural rearrangement of Ca2+ /CaM drives the channel into a low PO state producing CDI (Fig. 5, bottom).
There is significant rearrangement in the CaV/CaM interaction upon Ca2+ entry into the cell affecting channel gating. Unfortunately, the lack of structural data for many of the binding states presents a challenging mechanistic problem. However, additional insight into these structural aspects may be gained by looking outside the CaV channel family. Knowledge of CaV Ca2+ regulation has already informed on the function of NaV channel Ca2+ regulation as described previously.86 Given the homology between the CaV and NaV channel carboxyl termini, it is reasonable to apply the structural insights gained in the NaV channel field to the CaV channel.110-112 Such comparisons of function and structure between the two channel families have created a synergistic relationship in which both fields have prospered.
Ca2+ in Systems
A fine mechanistic understanding of the role of Ca2+ signaling in cellular processes hints at its broader impact in biological systems and disease states. Indeed, the highly coordinated and complex nature of Ca2+ signaling and its involvement in a wide array of biological functions – spanning the gamut from neural excitability and synaptic transmission to excitation-contraction coupling – argues for the necessity of its study. A firmer grasp of the biological impact of Ca2+ signaling requires linking single-molecule mechanisms described previously in this review with their impact on higher order functions of whole systems.
In neurons, Ca2+ orchestrates a number of functions, such as vesicle fusion113 and synaptic plasticity.114 Moreover, transient increases in intracellular Ca2+ are directly related to neural activity.115 Thus the development of robust genetically encoded calcium indicators (GECIs) is enabling for the study of neural activity. While much work has focused on the study of single neuron activity, the study of cortical activity on larger scales is useful for addressing the question of cortical organization such as the tonotopic arrangement of auditory cortex. To address this question, the Yue lab has pioneered a multiscale approach that takes advantage of transgenic mice expressing GCaMP3, a GECI, in neurons.116 These mice allowed for large-scale transcranial mapping of auditory cortex followed by fine-scale 2-photon imaging of Ca2+ activity in individual neurons within single trials (Fig. 6A–C). Frequency tuning curves of individual neurons in the primary auditory cortex, as seen by 2-photon imaging, adhered tightly to tuning predicted by their location in the large-scale maps (Fig. 6D–F), confirming the presence of a tonotopic axis in mouse cortex.117 These experiments brought the CSL full circle from David Yue's early experiments as a graduate student injecting aequorin into cardiac cells.18
Figure 6.
Multiscale Ca2+ imaging of mouse auditory cortex. (A) Baseline fluorescence image obtained via widefield transcranial imaging of GCaMP3 transgenic mice. (B) Ca2+ transients in response to tones as obtained from widefield imaging. Responses taken from region delineated by red square in A. (C) Tonotopic map obtained from transcranial imaging. ‘L’ and ‘H’ highlight low- and high-frequency landmarks, respectively. The red square indicates region scrutinized in D-E under 2-photon microscopy. AI: primary auditory cortex; AII: secondary auditory cortex; AAF: anterior auditory field; UF: ultrasonic field. (D) Two-photon imaging field showing individual neurons. (E) Ca2+ transients from a single neuron (indicated by black square in D). Average response to tones (dark trace) and individual responses (gray traces) are shown, indicating low-frequency tuning in this particular neuron. (F) Population data for tonotopic axis in AI. Each square indicates the average best frequency (BF) for a field of view and each dot indicates BF for an individual neuron. The red square is the same as in (C–E). (0 = ‘L’ pole and 1 = ‘H’ pole). Reproduced from Issa et al.117 © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the rightsholder.
Beyond an understanding of healthy cardiac and neural systems, Ca2+ signaling also affords a means of studying disease states, which not only helps in understanding the disease itself but can also aid in recognizing the general underlying biological principles and in deciphering which biophysical features are most relevant in the disease states. For example, owing to their diverse physiological roles, it is perhaps not surprising that disruption of CaV function has been implicated in numerous human pathologies. Diseases associated with CaVs are referred to as Ca2+ channelopathies and to date there have been 10 human channelopathies associated with 5 of the 10 CaV α-subunit genes. The consequences of disease causing mutations on CaV channel function varies widely and includes many loss-of-function and gain-of-function effects on voltage dependency and kinetics of channel gating.118 However much of the pioneering work of the Yue lab on calmodulin regulation of CaV channels has inspired subsequent work exploring the effects of human disease mutations on calcium regulation of CaV channels.
Several mutations in the CACNA1A gene encoding the P/Q-type channel α-subunit (CaV2.1) enriched within central nervous system cause channelopathies with severe neurological pathology. Such disorders include Familial Hemiplegic Migraine (FHM1), Episodic Ataxia type 2 (EA2)119 and Spinocerebellar Ataxia type 6 (SCA6).120 In fact, mutations associated with FHM1and EA2 have been shown to modify CDF and CDI of CaV2.1 channels and alter short-term synaptic plasticity.121 Furthermore, the L-type channel α1 subunit (CaV1.4) is localized within retinal cells and mutations in the CACNA1F gene encoding this channel are implicated in the vision related disorders incomplete X-linked congenital stationary night blindness (IXLCSNB) and X-linked cone-rod dystrophy (CORDX).122, 123 One IXLCSNB mutation within the carboxyl tail of the CaV1.4 enables CaM binding, resulting in robust CDI in these channels124 and a boost in channel open probability109 Furthermore, mutations in the CACNA1C gene encoding the α1 subunit CaV1.2 enriched in cardiac tissue have been associated with a severe arrhythmic disorder called Timothy syndrome, 125 and implicated in autism spectrum disorder (ASD).126 These mutations have been shown to have significant effect on CaV1.2 CDI (Dick et al, BPS abstract).
The recent work by the Yue lab demonstrating the conservation of CaM regulation between calcium and sodium channels has already shown that channelopathic disease mutations in sodium channels127, 128 alter sodium channel CDI, 86 and likely channel open probability.109 The consequences could be extensive, as the NaV1 superfamily governs excitability in brain, heart, and skeletal muscle, 129 and related diseases encompass epilepsy, autism, pathological pain, cardiac arrhythmias, and skeletal muscle myotonias.127, 128 Specifically, a number of mutations in NaV1.4, an isoform commonly found in skeletal muscle, are associated with myotonia, difficulty voluntarily relaxing skeletal muscle. Genetic screening revealed a variety of mutations in NaV1.4 corresponding to a wide spectrum of the severity of myotonia.130–132 Since a subset of these mutations lie in the carboxyl-tail of NaV1.4, 133, 134 one might postulate that the disruption of CDI could be one of the mechanisms for myotonia in this subgroup as CaM orchestrates CDI of NaV1.4 through interacting with the channel's carboxyl tail.86
Beyond skeletal NaV1.4 channels, mutations in cardiac NaV1.5 channels have also been implicated in several prominent cardiac channelopathies such as Brugada syndrome and congenital long QT syndrome (LQTS).135 While the mechanism behind Brugada syndrome has yet to be firmly established, decreased Na current density is generally accepted as the functional consequence.135-137 In contrast, in LQTS associated with mutations in SCN5A, a gene encoding the NaV1.5 channel, an increase in persistent inward sodium current is often implicated as the mechanism of action potential lengthening in LQTS.135, 138-140 Interestingly, many Brugada syndrome and LQTS mutations occur within the carboxyl-terminus of NaV1.5 channels, 141-145 thus calmodulation and Ca2+ regulation dysfunction of NaV1.5 present as promising potential mechanisms behind the 2 prominent cardiac channelopathies with disparate functional effects on Na channel function. For example, a A1924T mutation associated with Brugada syndrome leads to decreased CaM binding affinity to the NaV1.5 carboxyl-terminus and may alter NaV1.5 channel gating kinetics.80, 85, 146 Taken together, recent insights and elucidation in calmodulation and Ca2+ regulation of NaV1.5 channels leads to exciting frontiers in understanding cardiac channelopathy mechanisms.
Not only do disruptions in CaV channels themselves lead to disease, but alterations to CaM itself can underlie pathophysiology. For instance, the open probability of the long variant of CaV1.3 is strongly dependent on the concentration of CaM within cells (Fig. 7A, middle). When CaM levels were increased within dopaminergic neurons (where the long variant of CaV1.3 is found), the action potential duration of these pacemaking cells were dramatically increased (Fig. 7A, right). With increased Ca2+ exposure linked to neuronal toxicity, 147 the finding that the concentration of CaM can significantly alter Ca2+ influx raises the possibility of CaM's involvement in the pathogenesis of Parkinson's disease.
Figure 7.

Biophysical and physiological impact of altered CaM. (A) CaM boosts a channel open probability (red). In neurons, overexpression of CaM increases the magnitude of Ca2+ current leading to depolarized resting membrane potential and the lengthening of action potential duration. Reproduced from Adams et al., 2014.109 (B) CaM orchestrates CDI of the L-type Ca2+ channel. LQTS-associated mutant CaM has diminished Ca2+ binding affinity slows current inactivation (red), thus prolonging cardiac APD, a cellular correlate of LQTS. Reproduced from Limpitikul et al., 2014. © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the rightsholder.
Additionally, given the importance of CaM in triggering the CDI of L-type Ca2+ channels in the heart, mutations within CaM might disrupt CDI, prolonging the cardiac action potential via an increase in the inward current during the plateau phase. Indeed, myocytes overexpressing CaM1234 exhibit markedly prolonged action potentials.148, 149 These studies highlighted the importance of CaM in the heart, adding credence to the widespread belief that naturally-occurring mutant CaMs would likely be so deleterious to be incompatible with life. However, it was not until a decade later, that a rare group of diseases resulting from heterozygous missense mutations within one of the 3 genes (CALM1-3) encoding identical CaM proteins was discovered. These calmodulinopathies result in catecholaminergic polymorphic ventricular tachycardia (CPVT) and severe long-QT syndrome (LQTS) with recurrent cardiac arrest.150
Expressing these CaM mutants in myocytes (Fig. 7B, left), the Yue lab found that the CDI of cardiac L-type Ca2+ channels was attenuated (Fig. 7B, middle). Although a diminished CDI could explain the QT interval prolongation, the severity of the disease in human patients was puzzling, given that the number of wild-type CaM molecules is thought to greatly outnumber the number of mutant CaMs. Using FRET-based binding assays, the Yue lab demonstrated that the mutant CaMs had preserved binding with the L-type channel, suggesting that a fraction of L-type channels were occupied by a mutant CaM. Through modeling and further electrophysiology, they showed that the large effect that mutant CaM has on a small fraction of channels was sufficient to produce a significant reduction of CDI, 22 leading to a large prolongation of action potential duration (Fig. 7B, right).
Remembering a Great Scientist
From his early childhood as a ham radio operator, David T. Yue was fascinated by electrical signals and delighted by transistors. It is no wonder then, that as he began his scientific career he gravitated to the biological equivalent of a transistor, the voltage gated Ca2+ channel. He brought with him a desire to apply detailed quantitative methods to biological function, a dream he made a reality. He sought to understand Ca2+ regulation at its most fundamental level, and yet he never lost sight of the biological implications of these mechanistic discoveries. As a result, David's scientific contributions spanned the spectrum all the way from atomic structure and single molecule recordings to live brain imaging. He was able to gain mechanistic insight within a single channel which could then be applied not only across channel subtypes, but across families of channels, ultimately linking 2 channel fields.
The story of Ca2+ regulation presented here is but a glimpse into the world of David T. Yue. He was a seeker of truth who pursued science with a sense of wonder which was inspirational to all who knew him. He was passionate about his research, relating experimental discovery with “a syllable that God spoke.” All who passed through the Calcium Signals Lab have fond memories of late nights of exciting research, punctuated with humorous and loving stories of David's three exceptional sons, Michael, Daniel and Jonathan. David's wife Nancy was such a strong and supportive presence in the CSL that she earned the nickname ‘Saint Nancy’ from the lab members. So now, as David's Calcium Signals Lab family looks to the future, we remember the lessons he taught us and strive to carry his legacy forward.
“As for you, you mustn't let what could have been destroy the dream and wonder of what is and what will be. Keep your bow pointed – Good luck.” – David T. Yue, 1980
Disclosure of Potential Conflicts of Interest
No potential conflicts of interest were disclosed.
Acknowledgments
We would like to thank David T. Yue for his years of dedicated and enthusiastic support, guidance and friendship. He was a brilliant scientist, caring mentor and inspiring teacher who will be greatly missed.
Funding
This work was supported by grants from the NHLBI (R37HL076795), NIMH (R01MH065531), NINDS (R01NS073874) and NINDS (R01NS08074)to David T. Yue; NHLBI(HL50411) to GT, AHA(13PRE16500029) to WL, NSF(DGE-1232825) to JN and MSTP fellowships to JBI and SRL.
References
- 1.Burkhoff D, Yue DT, Franz MR, Hunter WC, Sagawa K. Mechanical restitution of isolated perfused canine left ventricles. Am J Physiol 1984; 246:H8-16; PMID:6696092 [DOI] [PubMed] [Google Scholar]
- 2.Burkhoff D, Yue DT, Franz MR, Hunter WC, Sunagawa K, Maughan WL, Sagawa K. Quantitative comparison of the force-interval relationships of the canine right and left ventricles. Circ Res 1984; 54:468-73; PMID:6201298; http://dx.doi.org/ 10.1161/01.RES.54.4.468 [DOI] [PubMed] [Google Scholar]
- 3.Yue DT, Burkhoff D, Franz MR, Hunter WC, Sagawa K. Postextrasystolic potentiation of the isolated canine left ventricle. Relationship to mechanical restitution. Circ Res 1985; 56:340-50; PMID:2578901; http://dx.doi.org/ 10.1161/01.RES.56.3.340 [DOI] [PubMed] [Google Scholar]
- 4.Wier WG, Yue DT. Intracellular calcium transients underlying the short-term force-interval relationship in ferret ventricular myocardium. J Physiol 1986; 376:507-30; PMID:2432238; http://dx.doi.org/ 10.1113/jphysiol.1986.sp016167 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Yue DT, Marban E, Wier WG. Relationship between force and intracellular [Ca2+] in tetanized mammalian heart muscle. J Gen Physiol 1986; 87:223-42; PMID:2419483; http://dx.doi.org/ 10.1085/jgp.87.2.223 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Yue DT, Wier WG. Estimation of intracellular [Ca2+] by nonlinear indicators. A quantitative analysis. Biophys J 1985; 48:533-7; PMID:4041543; http://dx.doi.org/ 10.1016/S0006-3495(85)83810-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Grynkiewicz G, Poenie M, Tsien RY. A new generation of Ca2+ indicators with greatly improved fluorescence properties. J Biol Chem 1985; 260:3440-50; PMID:3838314 [PubMed] [Google Scholar]
- 8.Hastings JW, Mitchell G, Mattingly PH, Blinks JR, Van Leeuwen M. Response of aequorin bioluminescence to rapid changes in calcium concentration. Nature 1969; 222:1047-50; PMID:4389183; http://dx.doi.org/ 10.1038/2221047a0 [DOI] [PubMed] [Google Scholar]
- 9.Bers DM. Myofilaments. Excitation-Contraction Coupling and Cardiac Contractile Force. Dordrecht, The Netherlands: Kluwer Academic Publishers, 2001:28-35. [Google Scholar]
- 10.Sigworth FJ. Structural biology: Life's transistors. Nature 2003; 423:21-2; PMID:12721605; http://dx.doi.org/ 10.1038/423021a [DOI] [PubMed] [Google Scholar]
- 11.Aldrich RW, Corey DP, Stevens CF. A reinterpretation of mammalian sodium channel gating based on single channel recording. Nature 1983; 306:436-41; PMID:6316158; http://dx.doi.org/ 10.1038/306436a0 [DOI] [PubMed] [Google Scholar]
- 12.Yue DT, Marban E. A novel cardiac potassium channel that is active and conductive at depolarized potentials. Pflugers Archiv 1988; 413:127-33; PMID:3217234; http://dx.doi.org/ 10.1007/BF00582522 [DOI] [PubMed] [Google Scholar]
- 13.Yue DT, Marban E. Permeation in the dihydropyridine-sensitive calcium channel. Multi-ion occupancy but no anomalous mole-fraction effect between Ba2+ and Ca2+. J Gen Physiol 1990; 95:911-39; PMID:2163433; http://dx.doi.org/ 10.1085/jgp.95.5.911 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Backx PH, Yue DT, Lawrence JH, Marban E, Tomaselli GF. Molecular localization of an ion-binding site within the pore of mammalian sodium channels. Science 1992; 257:248-51; PMID:1321496; http://dx.doi.org/ 10.1126/science.1321496 [DOI] [PubMed] [Google Scholar]
- 15.Yue DT, Lawrence JH, Marban E. Two molecular transitions influence cardiac sodium channel gating. Science 1989; 244:349-52; PMID:2540529; http://dx.doi.org/ 10.1126/science.2540529 [DOI] [PubMed] [Google Scholar]
- 16.Yue DT, Herzig S, Marban E. Beta-adrenergic stimulation of calcium channels occurs by potentiation of high-activity gating modes. Proc Natl Acad Sci U S A 1990; 87:753-7; PMID:1689051; http://dx.doi.org/ 10.1073/pnas.87.2.753 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Ringer S, Buxton DW. Concerning the action of small quantities of calcium, sodium, and potassium salts upon the vitality and function of contractile tissue and the cuticular cells of fishes. J Physiol 1885; 6:154-61; PMID:16991412; http://dx.doi.org/ 10.1113/jphysiol.1885.sp000193 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Yue DT. Intracellular [Ca2+] related to rate of force development in twitch contraction of heart. Am J Physiol 1987; 252:H760-70; PMID:3565592 [DOI] [PubMed] [Google Scholar]
- 19.Burkhoff D, Yue DT, Oikawa RY, Franz MR, Schaefer J, Sagawa K. Influence of ventricular contractility on non-work-related myocardial oxygen consumption. Heart Vessels 1987; 3:66-72; PMID:3693257; http://dx.doi.org/ 10.1007/BF02058521 [DOI] [PubMed] [Google Scholar]
- 20.Bers DM. Sarcoplasmic reticulum Ca release in intact ventricular myocytes. Front Biosci 2002; 7:d1697-711; PMID:12086924; http://dx.doi.org/ 10.2741/bers [DOI] [PubMed] [Google Scholar]
- 21.Cheng H, Lederer WJ, Cannell MB. Calcium sparks: elementary events underlying excitation-contraction coupling in heart muscle. Science 1993; 262:740-4; PMID:8235594; http://dx.doi.org/ 10.1126/science.8235594 [DOI] [PubMed] [Google Scholar]
- 22.Limpitikul WB, Dick IE, Joshi-Mukherjee R, Overgaard MT, George AL Jr., Yue DT. Calmodulin mutations associated with long QT syndrome prevent inactivation of cardiac L-type Ca(2+) currents and promote proarrhythmic behavior in ventricular myocytes. J Mol Cell Cardiol 2014; 74:115-24; PMID:24816216; http://dx.doi.org/ 10.1016/j.yjmcc.2014.04.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Splawski I, Timothy KW, Decher N, Kumar P, Sachse FB, Beggs AH, Sanguinetti MC, Keating MT. Severe arrhythmia disorder caused by cardiac L-type calcium channel mutations. Proc Natl Acad Sci U S A 2005; 102:8089-96; discussion 6–8; PMID:15863612; http://dx.doi.org/ 10.1073/pnas.0502506102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Anderson ME. Calmodulin kinase and L-type calcium channels; a recipe for arrhythmias? Trends Cardiovasc Med 2004; 14:152-61; PMID:15177266; http://dx.doi.org/ 10.1016/j.tcm.2004.02.005 [DOI] [PubMed] [Google Scholar]
- 25.Molkentin JD. Calcineurin-NFAT signaling regulates the cardiac hypertrophic response in coordination with the MAPKs. Cardiovasc Res 2004; 63:467-75; PMID:15276472; http://dx.doi.org/ 10.1016/j.cardiores.2004.01.021 [DOI] [PubMed] [Google Scholar]
- 26.Bean BP. Neurotransmitter inhibition of neuronal calcium currents by changes in channel voltage dependence. Nature 1989; 340:153-6; PMID:2567963; http://dx.doi.org/ 10.1038/340153a0 [DOI] [PubMed] [Google Scholar]
- 27.Sandow A. Excitation-contraction coupling in muscular response. Yale J Biol Med 1952; 25:176-201; PMID:13015950 [PMC free article] [PubMed] [Google Scholar]
- 28.Dunlap K, Luebke JI, Turner TJ. Exocytotic Ca2+ channels in mammalian central neurons. Trends Neurosci 1995; 18:89-98; PMID:7537420; http://dx.doi.org/ 10.1016/0166-2236(95)93882-X [DOI] [PubMed] [Google Scholar]
- 29.Dragicevic E, Schiemann J, Liss B. Dopamine midbrain neurons in health and Parkinson's disease: emerging roles of voltage-gated calcium channels and ATP-sensitive potassium channels. Neuroscience 2015; 284:798-814; PMID:25450964; http://dx.doi.org/ 10.1016/j.neuroscience.2014.10.037 [DOI] [PubMed] [Google Scholar]
- 30.Pennartz CM, de Jeu MT, Bos NP, Schaap J, Geurtsen AM. Diurnal modulation of pacemaker potentials and calcium current in the mammalian circadian clock. Nature 2002; 416:286-90; PMID:11875398; http://dx.doi.org/ 10.1038/nature728 [DOI] [PubMed] [Google Scholar]
- 31.Zucker RS, Regehr WG. Short-term synaptic plasticity. Annu Rev Physiol 2002; 64:355-405; PMID:11826273; http://dx.doi.org/ 10.1146/annurev.physiol.64.092501.114547 [DOI] [PubMed] [Google Scholar]
- 32.Lisman J, Yasuda R, Raghavachari S. Mechanisms of CaMKII action in long-term potentiation. Nat Rev Neurosci 2012; 13:169-82; PMID:22334212 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Takahashi M, Seagar MJ, Jones JF, Reber BF, Catterall WA. Subunit structure of dihydropyridine-sensitive calcium channels from skeletal muscle. Proc Natl Acad Sci U S A 1987; 84:5478-82; PMID:2440051; http://dx.doi.org/ 10.1073/pnas.84.15.5478 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Brehm P, Eckert R. Calcium entry leads to inactivation of calcium channel in Paramecium. Science 1978; 202:1203-6; PMID:103199; http://dx.doi.org/ 10.1126/science.103199 [DOI] [PubMed] [Google Scholar]
- 35.Eckert R, Chad JE. Inactivation of Ca channels. Prog Biophys Mol Biol 1984; 44:215-67; PMID:6095365; http://dx.doi.org/ 10.1016/0079-6107(84)90009-9 [DOI] [PubMed] [Google Scholar]
- 36.Lee KS, Marban E, Tsien RW. Inactivation of calcium channels in mammalian heart cells: joint dependence on membrane potential and intracellular calcium. J Physiol 1985; 364:395-411; PMID:2411919; http://dx.doi.org/ 10.1113/jphysiol.1985.sp015752 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Kass RS, Sanguinetti MC. Inactivation of calcium channel current in the calf cardiac Purkinje fiber. Evidence for voltage- and calcium-mediated mechanisms. J Gen Physiol 1984; 84:705-26; PMID:6096480; http://dx.doi.org/ 10.1085/jgp.84.5.705 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Lux HD, Brown AM. Single channel studies on inactivation of calcium currents. Science 1984; 225:432-4; PMID:6330896; http://dx.doi.org/ 10.1126/science.6330896 [DOI] [PubMed] [Google Scholar]
- 39.Rosenberg RL, Hess P, Tsien RW. Cardiac calcium channels in planar lipid bilayers. L-type channels and calcium-permeable channels open at negative membrane potentials. J Gen Physiol 1988; 92:27-54; PMID:2844956; http://dx.doi.org/ 10.1085/jgp.92.1.27 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Yue DT, Backx PH, Imredy JP. Calcium-sensitive inactivation in the gating of single calcium channels. Science 1990; 250:1735-8; PMID:2176745; http://dx.doi.org/ 10.1126/science.2176745 [DOI] [PubMed] [Google Scholar]
- 41.Imredy JP, Yue DT. Mechanism of Ca(2+)-sensitive inactivation of L-type Ca2+ channels. Neuron 1994; 12:1301-18; PMID:8011340; http://dx.doi.org/ 10.1016/0896-6273(94)90446-4 [DOI] [PubMed] [Google Scholar]
- 42.DeMaria CD, Soong TW, Alseikhan BA, Alvania RS, Yue DT. Calmodulin bifurcates the local Ca2+ signal that modulates P/Q-type Ca2+ channels. Nature 2001; 411:484-9; PMID:11373682; http://dx.doi.org/ 10.1038/35078091 [DOI] [PubMed] [Google Scholar]
- 43.de Leon M, Wang Y, Jones L, Perez-Reyes E, Wei X, Soong TW, Snutch TP, Yue DT. Essential Ca(2+)-binding motif for Ca(2+)-sensitive inactivation of L-type Ca2+ channels. Science 1995; 270:1502-6; PMID:7491499; http://dx.doi.org/ 10.1126/science.270.5241.1502 [DOI] [PubMed] [Google Scholar]
- 44.Chad JE, Eckert R. An enzymatic mechanism for calcium current inactivation in dialysed Helix neurones. J Physiol 1986; 378:31-51; PMID:2432251; http://dx.doi.org/ 10.1113/jphysiol.1986.sp016206 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Sherman A, Keizer J, Rinzel J. Domain model for Ca2(+)-inactivation of Ca2+ channels at low channel density. Biophys J 1990; 58:985-95; PMID:2174274; http://dx.doi.org/ 10.1016/S0006-3495(90)82443-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Nyegaard M, Overgaard MT, Sondergaard MT, Vranas M, Behr ER, Hildebrandt LL, Lund J, Hedley PL, Camm AJ, Wettrell G, et al.. Mutations in calmodulin cause ventricular tachycardia and sudden cardiac death. Am J Hum Genet 2012; 91:703-12; PMID:23040497; http://dx.doi.org/ 10.1016/j.ajhg.2012.08.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Armstrong DL. Calcium channel regulation by calcineurin, a Ca2+-activated phosphatase in mammalian brain. Trends Neurosci 1989; 12:117-22; PMID:2469218; http://dx.doi.org/ 10.1016/0166-2236(89)90168-9 [DOI] [PubMed] [Google Scholar]
- 48.Peterson BZ, Lee JS, Mulle JG, Wang Y, de Leon M, Yue DT. Critical determinants of Ca(2+)-dependent inactivation within an EF-hand motif of L-type Ca(2+) channels. Biophys J 2000; 78:1906-20; PMID:10733970; http://dx.doi.org/ 10.1016/S0006-3495(00)76739-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Xia XM, Fakler B, Rivard A, Wayman G, Johnson-Pais T, Keen JE, Lund J, Hedley PL, Camm AJ, Wettrell G, et al.. Mechanism of calcium gating in small-conductance calcium-activated potassium channels. Nature 1998; 395:503-7; PMID:9774106; http://dx.doi.org/ 10.1038/26758 [DOI] [PubMed] [Google Scholar]
- 50.Peterson BZ, DeMaria CD, Adelman JP, Yue DT. Calmodulin is the Ca2+ sensor for Ca2+ -dependent inactivation of L-type calcium channels. Neuron 1999; 22:549-58; PMID:10197534; http://dx.doi.org/ 10.1016/S0896-6273(00)80709-6 [DOI] [PubMed] [Google Scholar]
- 51.Lee A, Wong ST, Gallagher D, Li B, Storm DR, Scheuer T, Catterall WA. Ca2+/calmodulin binds to and modulates P/Q-type calcium channels. Nature 1999; 399:155-9; PMID:10335845; http://dx.doi.org/ 10.1038/399a007 [DOI] [PubMed] [Google Scholar]
- 52.Lee A, Scheuer T, Catterall WA. Ca2+/calmodulin-dependent facilitation and inactivation of P/Q-type Ca2+ channels. J Neurosci 2000; 20:6830-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Yang PS, Alseikhan BA, Hiel H, Grant L, Mori MX, Yang W, Fuchs PA, Yue DT. Switching of Ca2+-dependent inactivation of Ca(v)1.3 channels by calcium binding proteins of auditory hair cells. J Neurosci 2006; 26:10677-89; PMID:17050707; http://dx.doi.org/ 10.1523/JNEUROSCI.3236-06.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Liang H, DeMaria CD, Erickson MG, Mori MX, Alseikhan BA, Yue DT. Unified mechanisms of Ca2+ regulation across the Ca2+ channel family. Neuron 2003; 39:951-60; PMID:12971895; http://dx.doi.org/ 10.1016/S0896-6273(03)00560-9 [DOI] [PubMed] [Google Scholar]
- 55.Bootman MD, Collins TJ, Peppiatt CM, Prothero LS, MacKenzie L, De Smet P, Travers M, Tovey SC, Seo JT, Berridge MJ, et al.. Calcium signalling–an overview. Semin Cell Devel Biol 2001; 12:3-10; PMID:11162741; http://dx.doi.org/ 10.1006/scdb.2000.0211 [DOI] [PubMed] [Google Scholar]
- 56.Dunlap K. Calcium channels are models of self-control. J Gen Physiol 2007; 129:379-83; PMID:17438121; http://dx.doi.org/ 10.1085/jgp.200709786 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Singh A, Hamedinger D, Hoda JC, Gebhart M, Koschak A, Romanin C, Striessnig J. C-terminal modulator controls Ca2+-dependent gating of CaV1.4 L-type Ca2+ channels. Nat Neurosci 2006; 9:1108-16; PMID:16921373; http://dx.doi.org/ 10.1038/nn1751 [DOI] [PubMed] [Google Scholar]
- 58.Wahl-Schott C, Baumann L, Cuny H, Eckert C, Griessmeier K, Biel M. Switching off calcium-dependent inactivation in L-type calcium channels by an autoinhibitory domain. Proc Natl Acad Sci U S A 2006; 103:15657-62; PMID:17028172; http://dx.doi.org/ 10.1073/pnas.0604621103 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Lee A, Zhou H, Scheuer T, Catterall WA. Molecular determinants of Ca(2+)/calmodulin-dependent regulation of Ca(v)2.1 channels. Proc Natl Acad Sci U S A 2003; 100:16059-64; PMID:14673106; http://dx.doi.org/ 10.1073/pnas.2237000100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Li W, Halling DB, Hall AW, Aldrich RW. EF hands at the N-lobe of calmodulin are required for both SK channel gating and stable SK-calmodulin interaction. J General Physiol 2009; 134:281-93; PMID:19752189; http://dx.doi.org/ 10.1085/jgp.200910295 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Zuhlke RD, Pitt GS, Deisseroth K, Tsien RW, Reuter H. Calmodulin supports both inactivation and facilitation of L-type calcium channels. Nat 1999; 399:159-62; PMID:10335846; http://dx.doi.org/ 10.1038/20200 [DOI] [PubMed] [Google Scholar]
- 62.Stern MD. Buffering of calcium in the vicinity of a channel pore. Cell Calcium 1992; 13:183-92; PMID:1315621; http://dx.doi.org/ 10.1016/0143-4160(92)90046-U [DOI] [PubMed] [Google Scholar]
- 63.Neher E, Almers W. Fast calcium transients in rat peritoneal mast cells are not sufficient to trigger exocytosis. EMBO J 1986; 5:51-3; PMID:3082623 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Dick IE, Tadross MR, Liang H, Tay LH, Yang W, Yue DT. A modular switch for spatial Ca2+ selectivity in the calmodulin regulation of CaV channels. Nature 2008; 451:830-4; PMID:18235447; http://dx.doi.org/ 10.1038/nature06529 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Tadross MR, Dick IE, Yue DT. Mechanism of local and global Ca2+ sensing by calmodulin in complex with a Ca2+ channel. Cell 2008; 133:1228-40; PMID:18585356; http://dx.doi.org/ 10.1016/j.cell.2008.05.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Chaudhuri D, Issa JB, Yue DT. Elementary mechanisms producing facilitation of Cav2.1 (P/Q-type) channels. J Gen Physiol 2007; 129:385-401; PMID:17438119; http://dx.doi.org/ 10.1085/jgp.200709749 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Gamper N, Li Y, Shapiro MS. Structural requirements for differential sensitivity of KCNQ K+ channels to modulation by Ca2+/calmodulin. Mol Biol Cell 2005; 16:3538-51; PMID:15901836; http://dx.doi.org/ 10.1091/mbc.E04-09-0849 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Schumacher MA, Crum M, Miller MC. Crystal structures of apocalmodulin and an apocalmodulin/SK potassium channel gating domain complex. Structure 2004; 12:849-60; PMID:15130477; http://dx.doi.org/ 10.1016/j.str.2004.03.017 [DOI] [PubMed] [Google Scholar]
- 69.Ghosh S, Nunziato DA, Pitt GS. KCNQ1 assembly and function is blocked by long-QT syndrome mutations that disrupt interaction with calmodulin. Circul Res 2006; 98:1048-54; PMID:16556866; http://dx.doi.org/ 10.1161/01.RES.0000218863.44140.f2 [DOI] [PubMed] [Google Scholar]
- 70.Bradley J, Reisert J, Frings S. Regulation of cyclic nucleotide-gated channels. Curr Opin Neurobiol 2005; 15:343-9; PMID:15922582; http://dx.doi.org/ 10.1016/j.conb.2005.05.014 [DOI] [PubMed] [Google Scholar]
- 71.Trudeau MC, Zagotta WN. Dynamics of Ca2+-calmodulin-dependent inhibition of rod cyclic nucleotide-gated channels measured by patch-clamp fluorometry. J Gen Physiol 2004; 124:211-23; PMID:15314069; http://dx.doi.org/ 10.1085/jgp.200409101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Ehlers MD, Zhang S, Bernhadt JP, Huganir RL. Inactivation of NMDA receptors by direct interaction of calmodulin with the NR1 subunit. Cell 1996; 84:745-55; PMID:8625412; http://dx.doi.org/ 10.1016/S0092-8674(00)81052-1 [DOI] [PubMed] [Google Scholar]
- 73.Zhang S, Ehlers MD, Bernhardt JP, Su CT, Huganir RL. Calmodulin mediates calcium-dependent inactivation of N-methyl-D-aspartate receptors. Neuron 1998; 21:443-53; PMID:9728925; http://dx.doi.org/ 10.1016/S0896-6273(00)80553-X [DOI] [PubMed] [Google Scholar]
- 74.Numazaki M, Tominaga T, Takeuchi K, Murayama N, Toyooka H, Tominaga M. Structural determinant of TRPV1 desensitization interacts with calmodulin. Proc Nat Acad Sci U S A 2003; 100:8002-6; PMID:12808128; http://dx.doi.org/ 10.1073/pnas.1337252100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Rosenbaum T, Gordon-Shaag A, Munari M, Gordon SE. Ca2+/calmodulin modulates TRPV1 activation by capsaicin. J Gen Physiol 2004; 123:53-62; PMID:14699077; http://dx.doi.org/ 10.1085/jgp.200308906 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Gordon-Shaag A, Zagotta WN, Gordon SE. Mechanism of Ca(2+)-dependent desensitization in TRP channels. Channels 2008; 2:125-9; PMID:18849652; http://dx.doi.org/ 10.4161/chan.2.2.6026 [DOI] [PubMed] [Google Scholar]
- 77.Mercado J, Gordon-Shaag A, Zagotta WN, Gordon SE. Ca2+-dependent desensitization of TRPV2 channels is mediated by hydrolysis of phosphatidylinositol 4, 5-bisphosphate. J Neurosci 2010; 30:13338-47; PMID:20926660; http://dx.doi.org/ 10.1523/JNEUROSCI.2108-10.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Lau SY, Procko E, Gaudet R. Distinct properties of Ca2+-calmodulin binding to N- and C-terminal regulatory regions of the TRPV1 channel. J Gen Physiol 2012; 140:541-55; PMID:23109716; http://dx.doi.org/ 10.1085/jgp.201210810 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Deschenes I, Neyroud N, DiSilvestre D, Marban E, Yue DT, Tomaselli GF. Isoform-specific modulation of voltage-gated Na(+) channels by calmodulin. Circul Res 2002; 90:E49-57; PMID:11884381; http://dx.doi.org/ 10.1161/01.RES.0000012502.92751.E6 [DOI] [PubMed] [Google Scholar]
- 80.Tan HL, Kupershmidt S, Zhang R, Stepanovic S, Roden DM, Wilde AA, Anderson ME, Balser JR. A calcium sensor in the sodium channel modulates cardiac excitability. Nature 2002; 415:442-7; PMID:11807557; http://dx.doi.org/ 10.1038/415442a [DOI] [PubMed] [Google Scholar]
- 81.Herzog RI, Liu C, Waxman SG, Cummins TR. Calmodulin binds to the C terminus of sodium channels Nav1.4 and Nav1.6 and differentially modulates their functional properties. J Neurosci 2003; 23:8261-70; PMID:12967988 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Van Petegem F, Lobo PA, Ahern CA. Seeing the forest through the trees: towards a unified view on physiological calcium regulation of voltage-gated sodium channels. Biophys J 2012; 103:2243-51; PMID:23283222; http://dx.doi.org/ 10.1016/j.bpj.2012.10.020 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Biswas S, DiSilvestre D, Tian Y, Halperin VL, Tomaselli GF. Calcium-mediated dual-mode regulation of cardiac sodium channel gating. Circul Res 2009; 104:870-8; PMID:19265034; http://dx.doi.org/ 10.1161/CIRCRESAHA.108.193565 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Kim J, Ghosh S, Liu H, Tateyama M, Kass RS, Pitt GS. Calmodulin mediates Ca2+ sensitivity of sodium channels. J Biol Chem 2004; 279:45004-12; PMID:15316014; http://dx.doi.org/ 10.1074/jbc.M407286200 [DOI] [PubMed] [Google Scholar]
- 85.Shah VN, Wingo TL, Weiss KL, Williams CK, Balser JR, Chazin WJ. Calcium-dependent regulation of the voltage-gated sodium channel hH1: intrinsic and extrinsic sensors use a common molecular switch. Proc Natl Acad Sci U S A 2006; 103:3592-7; PMID:16505387; http://dx.doi.org/ 10.1073/pnas.0507397103 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Ben-Johny M, Yang PS, Niu J, Yang W, Joshi-Mukherjee R, Yue DT. Conservation of Ca2+/calmodulin regulation across Na and Ca2+ channels. Cell 2014; 157:1657-70; PMID:24949975; http://dx.doi.org/ 10.1016/j.cell.2014.04.035 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Chaudhuri D, Chang SY, DeMaria CD, Alvania RS, Soong TW, Yue DT. Alternative splicing as a molecular switch for Ca2+/calmodulin-dependent facilitation of P/Q-type Ca2+ channels. J Neurosci 2004; 24:6334-42; http://dx.doi.org/ 10.1523/JNEUROSCI.1712-04.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Jurado LA, Chockalingam PS, Jarrett HW. Apocalmodulin. Physiol Rev 1999; 79:661-82; PMID:10390515 [DOI] [PubMed] [Google Scholar]
- 89.Erickson MG, Alseikhan BA, Peterson BZ, Yue DT. Preassociation of calmodulin with voltage-gated Ca(2+) channels revealed by FRET in single living cells. Neuron 2001; 31:973-85; PMID:11580897; http://dx.doi.org/ 10.1016/S0896-6273(01)00438-X [DOI] [PubMed] [Google Scholar]
- 90.Erickson MG, Liang H, Mori MX, Yue DT. FRET two-hybrid mapping reveals function and location of L-type Ca2+ channel CaM preassociation. Neuron 2003; 39:97-107; PMID:12848935; http://dx.doi.org/ 10.1016/S0896-6273(03)00395-7 [DOI] [PubMed] [Google Scholar]
- 91.Pitt GS, Zuhlke RD, Hudmon A, Schulman H, Reuter H, Tsien RW. Molecular basis of calmodulin tethering and Ca2+-dependent inactivation of L-type Ca2+ channels. J Biol Chem 2001; 276:30794-802; PMID:11408490; http://dx.doi.org/ 10.1074/jbc.M104959200 [DOI] [PubMed] [Google Scholar]
- 92.Black DJ, Halling DB, Mandich DV, Pedersen SE, Altschuld RA, Hamilton SL. Calmodulin interactions with IQ peptides from voltage-dependent calcium channels. Am J Physiol Cell Physiol 2005; 288:C669-76; PMID:15496482; http://dx.doi.org/ 10.1152/ajpcell.00191.2004 [DOI] [PubMed] [Google Scholar]
- 93.Ben-Johny M, Yang PS, Bazzazi HX, Yue DT. Dynamic switching of calmodulin interactions underlies Ca2+ regulation of CaV1.3 channels. Nat Commun 2013; 4:1717; PMID:23591884; http://dx.doi.org/ 10.1038/ncomms2727 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Bazzazi H, Ben Johny M, Adams PJ, Soong TW, Yue DT. Continuously tunable Ca(2+) regulation of RNA-edited CaV1.3 channels. Cell Rep 2013; 5:367-77; PMID:24120865; http://dx.doi.org/ 10.1016/j.celrep.2013.09.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Findeisen F, Rumpf CH, Minor DL Jr. Apo states of calmodulin and CaBP1 control CaV1 voltage-gated calcium channel function through direct competition for the IQ domain. J Mol Biol 2013; 425:3217-34; PMID:23811053; http://dx.doi.org/ 10.1016/j.jmb.2013.06.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Evans TI, Hell JW, Shea MA. Thermodynamic linkage between calmodulin domains binding calcium and contiguous sites in the C-terminal tail of Ca(V)1.2. Biophys Chem 2012; 159:172-87; http://dx.doi.org/ 10.1016/j.bpc.2011.06.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Fallon JL, Halling DB, Hamilton SL, Quiocho FA. Structure of calmodulin bound to the hydrophobic IQ domain of the cardiac Ca(v)1.2 calcium channel. Structure 2005; 13:1881-6; PMID:16338416; http://dx.doi.org/ 10.1016/j.str.2005.09.021 [DOI] [PubMed] [Google Scholar]
- 98.Van Petegem F, Chatelain FC, Minor DL Jr. Insights into voltage-gated calcium channel regulation from the structure of the CaV1.2 IQ domain-Ca2+/calmodulin complex. Nat Struct Mol Biol 2005; 12:1108-15; PMID:16299511; http://dx.doi.org/ 10.1038/nsmb1027 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Mori MX, Vander Kooi CW, Leahy DJ, Yue DT. Crystal structure of the CaV2 IQ domain in complex with Ca2+/calmodulin: high-resolution mechanistic implications for channel regulation by Ca2+. Structure 2008; 16:607-20; PMID:18400181; http://dx.doi.org/ 10.1016/j.str.2008.01.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Kim EY, Rumpf CH, Fujiwara Y, Cooley ES, Van Petegem F, Minor DL Jr. Structures of CaV2 Ca2+/CaM-IQ domain complexes reveal binding modes that underlie calcium-dependent inactivation and facilitation. Structure 2008; 16:1455-67; PMID:18940602; http://dx.doi.org/ 10.1016/j.str.2008.07.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Ben-Johny M, Yue DT. Calmodulin regulation (calmodulation) of voltage-gated calcium channels. J Gen Physiol 2014; 143:679-92; PMID:24863929; http://dx.doi.org/ 10.1085/jgp.201311153 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Zuhlke RD, Reuter H. Ca2+-sensitive inactivation of L-type Ca2+ channels depends on multiple cytoplasmic amino acid sequences of the a1c subunit. Proc Natl Acad Sci USA 1998; 95:3287-94; http://dx.doi.org/ 10.1073/pnas.95.6.3287 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Liu X, Yang PS, Yang W, Yue DT. Enzyme-inhibitor-like tuning of Ca(2+) channel connectivity with calmodulin. Nature 2010; 463:968-72; PMID:20139964; http://dx.doi.org/ 10.1038/nature08766 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Griessmeier K, Cuny H, Rotzer K, Griesbeck O, Harz H, Biel M, Wahl-Schott C. Calmodulin is a functional regulator of Cav1.4 L-type Ca2+ channels. J Biol Chem 2009; 284:29809-16; PMID:19717559; http://dx.doi.org/ 10.1074/jbc.M109.048082 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Bock G, Gebhart M, Scharinger A, Jangsangthong W, Busquet P, Poggiani C, Sartori S, Mangoni ME, Sinnegger-Brauns MJ, Herzig S, et al.. Functional properties of a newly identified C-terminal splice variant of Cav1.3 L-type Ca2+ channels. J Biol Chem 2011; 286:42736-48; PMID:21998310; http://dx.doi.org/ 10.1074/jbc.M111.269951 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Huang H, Tan BZ, Shen Y, Tao J, Jiang F, Sung YY, Ng CK, Raida M, Köhr G, Higuchi M, et al.. RNA editing of the IQ domain of CaV1.3 channels modulates their Ca2+-dependent inactivation. Neuron 2012; 73:304-16; PMID:22284185; http://dx.doi.org/ 10.1016/j.neuron.2011.11.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Tan BZ, Jiang F, Tan MY, Yu D, Huang H, Shen Y, Soong TW. Functional characterization of alternative splicing in the C terminus of L-type CaV1.3 channels. J Biol Chem 2011; 286:42725-35; PMID:21998309; http://dx.doi.org/ 10.1074/jbc.M111.265207 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Ben Johny M, Yang PS, Bazzazi H, Yue DT. Dynamic switching of calmodulin interactions underlies Ca2+ regulation of CaV1.3 channels. Nat Commun 2013; 4:1717; PMID:23591884; http://dx.doi.org/ 10.1038/ncomms2727 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Adams PJ, Ben-Johny M, Dick IE, Inoue T, Yue DT. Apocalmodulin itself promotes ion channel opening and Ca(2+) regulation. Cell 2014; 159:608-22; PMID:25417111; http://dx.doi.org/ 10.1016/j.cell.2014.09.047 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Chagot B, Chazin WJ. Solution NMR structure of Apo-calmodulin in complex with the IQ motif of human cardiac sodium channel NaV1.5. J Mol Biol 2011; 406:106-19; PMID:21167176; http://dx.doi.org/ 10.1016/j.jmb.2010.11.046 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Sarhan MF, Tung CC, Van Petegem F, Ahern CA. Crystallographic basis for calcium regulation of sodium channels. Proc Natl Acad Sci U S A 2012; 109:3558-63; PMID:22331908; http://dx.doi.org/ 10.1073/pnas.1114748109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Gabelli SB, Boto A, Kuhns VH, Bianchet MA, Farinelli F, Aripirala S, Yoder J, Jakoncic J, Tomaselli GF, Amzel LM, et al.. Regulation of the NaV1.5 cytoplasmic domain by calmodulin. Nat Commun 2014; 5:5126; PMID:25370050; http://dx.doi.org/ 10.1038/ncomms6126 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Schneggenburger R, Neher E. Presynaptic calcium and control of vesicle fusion. Curr Opin Neurobiol 2005; 15:266-74; PMID:15919191; http://dx.doi.org/ 10.1016/j.conb.2005.05.006 [DOI] [PubMed] [Google Scholar]
- 114.Nishiyama M, Hong K, Mikoshiba K, Poo MM, Kato K. Calcium stores regulate the polarity and input specificity of synaptic modification. Nat 2000; 408:584-8; PMID:11117745; http://dx.doi.org/ 10.1038/35046067 [DOI] [PubMed] [Google Scholar]
- 115.Helmchen F, Imoto K, Sakmann B. Ca2+ buffering and action potential-evoked Ca2+ signaling in dendrites of pyramidal neurons. Biophys J 1996; 70:1069-81; PMID:8789126; http://dx.doi.org/ 10.1016/S0006-3495(96)79653-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Zariwala HA, Borghuis BG, Hoogland TM, Madisen L, Tian L, De Zeeuw CI, Zeng H, Looger LL, Svoboda K, Chen TW. A Cre-dependent GCaMP3 reporter mouse for neuronal imaging in vivo. J Neurosci 2012; 32:3131-41; PMID:22378886; http://dx.doi.org/ 10.1523/JNEUROSCI.4469-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Issa JB, Haeffele BD, Agarwal A, Bergles DE, Young ED, Yue DT. Multiscale optical Ca2+ imaging of tonal organization in mouse auditory cortex. Neuron 2014; 83:944-59; PMID:25088366; http://dx.doi.org/ 10.1016/j.neuron.2014.07.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Adams PJ, Snutch TP. Calcium channelopathies: voltage-gated calcium channels. Sub-Cell Biochem 2007; 45:215-51; PMID:18193639 [DOI] [PubMed] [Google Scholar]
- 119.Ophoff RA, Terwindt GM, Vergouwe MN, van Eijk R, Oefner PJ, Hoffman SM, Lamerdin JE, Mohrenweiser HW, Bulman DE, Ferrari M, et al.. Familial hemiplegic migraine and episodic ataxia type-2 are caused by mutations in the Ca2+ channel gene CACNL1A4. Cell 1996; 87:543-52; PMID:8898206; http://dx.doi.org/ 10.1016/S0092-8674(00)81373-2 [DOI] [PubMed] [Google Scholar]
- 120.Zhuchenko O, Bailey J, Bonnen P, Ashizawa T, Stockton DW, Amos C, Dobyns WB, Subramony SH, Zoghbi HY, Lee CC. Autosomal dominant cerebellar ataxia (SCA6) associated with small polyglutamine expansions in the alpha 1A-voltage-dependent calcium channel. Nat Genet 1997; 15:62-9; PMID:8988170; http://dx.doi.org/ 10.1038/ng0197-62 [DOI] [PubMed] [Google Scholar]
- 121.Adams PJ, Rungta RL, Garcia E, van den Maagdenberg AM, MacVicar BA, Snutch TP. Contribution of calcium-dependent facilitation to synaptic plasticity revealed by migraine mutations in the P/Q-type calcium channel. Proc Natl Acad Sci U S A 2010; 107:18694-9; PMID:20937883; http://dx.doi.org/ 10.1073/pnas.1009500107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Nachman-Clewner M, Jules R, Townes-Anderson E. L-type calcium channels in the photoreceptor ribbon synapse: localization and role in plasticity. J Comp Neurol 1999; 415:1-16; PMID:10540354; http://dx.doi.org/ 10.1002/(SICI)1096-9861(19991206)415:1%3c1::AID-CNE1%3e3.0.CO;2-G [DOI] [PubMed] [Google Scholar]
- 123.Tachibana M, Okada T, Arimura T, Kobayashi K, Piccolino M. Dihydropyridine-sensitive calcium current mediates neurotransmitter release from bipolar cells of the goldfish retina. J Neurosci 1993; 13:2898-909; PMID:7687280 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Singh A, Hamedinger D, Hoda JC, Gebhart M, Koschak A, Romanin C, Striessnig J. C-terminal modulator controls Ca2+-dependent gating of Ca(v)1.4 L-type Ca2+ channels. Nat Neurosci 2006; 9:1108-16; PMID:16921373; http://dx.doi.org/ 10.1038/nn1751 [DOI] [PubMed] [Google Scholar]
- 125.Splawski I, Timothy KW, Sharpe LM, Decher N, Kumar P, Bloise R, Napolitano C, Schwartz PJ, Joseph RM, Condouris K, et al.. Ca(V)1.2 calcium channel dysfunction causes a multisystem disorder including arrhythmia and autism. Cell 2004; 119:19-31; PMID:15454078; http://dx.doi.org/ 10.1016/j.cell.2004.09.011 [DOI] [PubMed] [Google Scholar]
- 126.Pinggera A, Lieb A, Benedetti B, Lampert M, Monteleone S, Liedl KR, Tuluc P, Striessnig J.. CACNA1D de novo mutations in autism spectrum disorders activate Cav1.3 L-type calcium channels. Biol Psychiatry 2014; 77(9):816-22; PMID:25620733 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Lossin C. A catalog of SCN1A variants. Brain Dev 2009; 31:114-30; PMID:18804930; http://dx.doi.org/ 10.1016/j.braindev.2008.07.011 [DOI] [PubMed] [Google Scholar]
- 128.Schroeter A, Walzik S, Blechschmidt S, Haufe V, Benndorf K, Zimmer T. Structure and function of splice variants of the cardiac voltage-gated sodium channel NaV1.5. J Mol Cell Cardiol 2010; 49:16-24; PMID:20398673; http://dx.doi.org/ 10.1016/j.yjmcc.2010.04.004 [DOI] [PubMed] [Google Scholar]
- 129.Hille B. Ionic channels of excitable membranes (Sunderland, MA, Sinauer Associates; ). 1984 [Google Scholar]
- 130.Fontaine B, Khurana TS, Hoffman EP, Bruns GA, Haines JL, Trofatter JA, Hanson MP, Rich J, McFarlane H, Yasek DM, et al.. Hyperkalemic periodic paralysis and the adult muscle sodium channel alpha-subunit gene. Science 1990; 250:1000-2; PMID:2173143; http://dx.doi.org/ 10.1126/science.2173143 [DOI] [PubMed] [Google Scholar]
- 131.Heatwole CR, Moxley RT 3rd. The nondystrophic myotonias. Neurotherapeutics 2007; 4:238-51; PMID:17395134; http://dx.doi.org/ 10.1016/j.nurt.2007.01.012 [DOI] [PubMed] [Google Scholar]
- 132.Rudel R, Ricker K, Lehmann-Horn F. Genotype-phenotype correlations in human skeletal muscle sodium channel diseases. Arch Neurol 1993; 50:1241-8; PMID:8215982; http://dx.doi.org/ 10.1001/archneur.1993.00540110113011 [DOI] [PubMed] [Google Scholar]
- 133.Kubota T, Kinoshita M, Sasaki R, Aoike F, Takahashi MP, Sakoda S, Hirose K. New mutation of the Na channel in the severe form of potassium-aggravated myotonia. Muscle Nerve 2009; 39:666-73; PMID:19347921; http://dx.doi.org/ 10.1002/mus.21155 [DOI] [PubMed] [Google Scholar]
- 134.Wu FF, Gordon E, Hoffman EP, Cannon SC. A C-terminal skeletal muscle sodium channel mutation associated with myotonia disrupts fast inactivation. J Physiol 2005; 565:371-80; PMID:15774523; http://dx.doi.org/ 10.1113/jphysiol.2005.082909 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.George AL, Jr. Inherited disorders of voltage-gated sodium channels. J Clin Invest 2005; 115:1990-9; PMID:16075039; http://dx.doi.org/ 10.1172/JCI25505 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Antzelevitch C. Ion channels and ventricular arrhythmias: cellular and ionic mechanisms underlying the Brugada syndrome. Curr Opin Cardiol 1999; 14:274-9; PMID:10358800; http://dx.doi.org/ 10.1097/00001573-199905000-00013 [DOI] [PubMed] [Google Scholar]
- 137.Yan GX, Antzelevitch C. Cellular basis for the Brugada syndrome and other mechanisms of arrhythmogenesis associated with ST-segment elevation. Circulation 1999; 100:1660-6; PMID:10517739; http://dx.doi.org/ 10.1161/01.CIR.100.15.1660 [DOI] [PubMed] [Google Scholar]
- 138.Dumaine R, Wang Q, Keating MT, Hartmann HA, Schwartz PJ, Brown AM, Kirsch GE. Multiple mechanisms of Na+ channel–linked long-QT syndrome. Circul Res 1996; 78:916-24; PMID:8620612; http://dx.doi.org/ 10.1161/01.RES.78.5.916 [DOI] [PubMed] [Google Scholar]
- 139.Roden DM, George AL Jr., Bennett PB. Recent advances in understanding the molecular mechanisms of the long QT syndrome. J Cardiovasc Electrophysiol 1995; 6:1023-31; PMID:8589871; http://dx.doi.org/ 10.1111/j.1540-8167.1995.tb00379.x [DOI] [PubMed] [Google Scholar]
- 140.Wang DW, Yazawa K, George AL Jr., Bennett PB. Characterization of human cardiac Na+ channel mutations in the congenital long QT syndrome. Proc Natl Acad Sci U S A 1996; 93:13200-5; PMID:8917568; http://dx.doi.org/ 10.1073/pnas.93.23.13200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Bezzina C, Veldkamp MW, van Den Berg MP, Postma AV, Rook MB, Viersma JW, van Langen IM, Tan-Sindhunata G, Bink-Boelkens MT, van Der Hout AH, et al.. A single Na(+) channel mutation causing both long-QT and Brugada syndromes. Circul Res 1999; 85:1206-13; PMID:10590249; http://dx.doi.org/ 10.1161/01.RES.85.12.1206 [DOI] [PubMed] [Google Scholar]
- 142.Kapplinger JD, Tester DJ, Alders M, Benito B, Berthet M, Brugada J, Brugada P, Fressart V, Guerchicoff A, Harris-Kerr C, et al.. An international compendium of mutations in the SCN5A-encoded cardiac sodium channel in patients referred for Brugada syndrome genetic testing. Heart Rhythm 2010; 7:33-46; PMID:20129283; http://dx.doi.org/ 10.1016/j.hrthm.2009.09.069 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Makita N, Behr E, Shimizu W, Horie M, Sunami A, Crotti L, Schulze-Bahr E, Fukuhara S, Mochizuki N, Makiyama T, et al.. The E1784K mutation in SCN5A is associated with mixed clinical phenotype of type 3 long QT syndrome. J Clin Invest 2008; 118:2219-29; PMID:18451998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Makita N, Mochizuki N, Tsutsui H. Absence of a trafficking defect in R1232W/T1620M, a double SCN5A mutant responsible for Brugada syndrome. Circul J 2008; 72:1018-9; PMID:18503232; http://dx.doi.org/ 10.1253/circj.72.1018 [DOI] [PubMed] [Google Scholar]
- 145.Millat G, Chevalier P, Restier-Miron L, Da Costa A, Bouvagnet P, Kugener B, Fayol L, Gonzàlez Armengod C, Oddou B, Chanavat V, et al.. Spectrum of pathogenic mutations and associated polymorphisms in a cohort of 44 unrelated patients with long QT syndrome. Clin Genet 2006; 70:214-27; PMID:16922724; http://dx.doi.org/ 10.1111/j.1399-0004.2006.00671.x [DOI] [PubMed] [Google Scholar]
- 146.Potet F, Chagot B, Anghelescu M, Viswanathan PC, Stepanovic SZ, Kupershmidt S, Chazin WJ, Balser JR. Functional interactions between distinct sodium channel cytoplasmic domains through the action of calmodulin. J Biol Chem 2009; 284:8846-54; PMID:19171938; http://dx.doi.org/ 10.1074/jbc.M806871200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Cali T, Ottolini D, Brini M. Calcium signaling in Parkinson's disease. Cell Tissue Res 2014; 357:439-54; PMID:24781149; http://dx.doi.org/ 10.1007/s00441-014-1866-0 [DOI] [PubMed] [Google Scholar]
- 148.Alseikhan BA, DeMaria CD, Colecraft HM, Yue DT. Engineered calmodulins reveal the unexpected eminence of Ca2+ channel inactivation in controlling heart excitation. Proc Natl Acad Sci U S A 2002; 99:17185-90; PMID:12486220; http://dx.doi.org/ 10.1073/pnas.262372999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Mahajan A, Sato D, Shiferaw Y, Baher A, Xie LH, Peralta R, Garfinkel A, Qu Z, Weiss JN. Modifying L-type calcium current kinetics: consequences for cardiac excitation and arrhythmia dynamics. Biophys J 2008; 94:411-23; PMID:18160661; http://dx.doi.org/ 10.1529/biophysj.106.98590 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Crotti L, Johnson CN, Graf E, De Ferrari GM, Cuneo BF, Ovadia M, Papagiannis J, Feldkamp MD, Rathi SG, Kunic JD, et al.. Calmodulin mutations associated with recurrent cardiac arrest in infants. Circulation 2013; 127:1009-17; PMID:23388215; http://dx.doi.org/ 10.1161/CIRCULATIONAHA.112.001216 [DOI] [PMC free article] [PubMed] [Google Scholar]




