Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Jun 15.
Published in final edited form as: Arch Biochem Biophys. 2011 Apr 3;510(2):174–181. doi: 10.1016/j.abb.2011.03.009

Regulation of gastrointestinal motility by Ca2+/calmodulin-stimulated protein kinase II

Brian A Perrino 1,*
PMCID: PMC3134147  NIHMSID: NIHMS290565  PMID: 21443856

Abstract

Gastrointestinal (GI) motility ultimately depends upon the contractile activity of the smooth muscle cells of the tunica muscularis. Integrated functioning of multiple tissues and cell types, including enteric neurons and interstitial cells of Cajal (ICC) is necessary to generate coordinated patterns of motor activity that control the movement of material through the digestive tract. The neurogenic mechanisms that govern GI motility patterns are superimposed upon intrinsic myogenic mechanisms regulating smooth muscle cell excitability. Several mechanisms regulate smooth muscle cell responses to neurogenic inputs, including the multifunctional Ca2+/calmodulin-stimulated protein kinase II (CaMKII). CaMKII can be activated by Ca2+ transients from both extracellular and intracellular sources. Prolonging the activities of Ca2+-sensitive K+ channels in the plasma membrane of GI smooth muscle cells is an important regulatory mechanism carried out by CaMKII. Phospholamban (PLN) phosphorylation by CaMKII activates the sarcoplasmic reticulum (SR) Ca2+-ATPase (SERCA), increasing both the rate of Ca2+clearance from the myoplasm and the frequency of localized Ca2+ release events from intracellular stores. Overall, CaMKII appears to moderate GI smooth muscle cell excitability. Finally, transcription factor activities may be facilitated by the neutralization of HDAC4 by CaMKII phosphorylation, which may contribute to the phenotypic plasticity of GI smooth muscle cells.

Keywords: Digestive tract, Motility, Smooth Muscle, Ca2+/calmodulin-dependent protein kinase II, K+channels, Phospholamban

Introduction

The digestive tract is a muscular tube in which food is ingested, nutrients are absorbed, and waste is eliminated. Processing of nutrients requires movement of materials through the digestive tract, which is accomplished by the coordinated contractions and relaxations of the smooth muscles (tunica muscularis) comprising the various organs of the digestive tract. GI motility refers to the movement, or lack thereof, of material through the digestive tract. Motility patterns encompass the propulsive, mixing, and reservoir functions of the digestive tract necessary for the spatio-temporal processing of ingested food and elimination of waste products. Functional integration of multiple tissues and cell types is necessary to generate the various motility patterns characteristic of each organ in the digestive tract [1]. The smooth muscles of the digestive tract organs generate the force and perform the work; receiving regulatory input from various control systems, including motor neurons, ICC, hormones, paracrine substances and inflammatory mediators. GI smooth muscle cells also exhibit autonomous (myogenic) behavior arising from intrinsic regulatory pathways that can amplify or oppose signaling from the higher regulatory systems, depending upon the state of the resting membrane potential of the smooth muscle cells [2]. The modern definition of intrinsic (i.e. non-neural, non-hormonal) or myogenic regulation of gut motility now includes, in addition to smooth muscle mechanisms, the actions of ICC. The ICC affect the resting membrane potential of smooth muscle cells via gap junctions and generate the pacemaker activity that is the basis for electrical slow wave activity in phasic GI smooth muscles [3]. It is now recognized that the neurogenic mechanisms regulating GI motility patterns are superimposed upon the myogenic mechanisms. The enteric nervous system is of fundamental importance for generating major motor patterns, and regulating the amplitudes of contractions, but the smooth muscle response to external signals ultimately depends upon the excitability of the ICC-smooth muscle syncytium [4].

Regulation of smooth muscle excitability by membrane potential

Regulation of resting membrane potential is an important intrinsic mechanism that controls smooth muscle cell excitability by establishing the point from which smooth muscle cells respond to depolarizations from pacemaker cells and agonists [2]. Resting membrane potentials of GI smooth muscle cells vary widely in different regions of the gut (−85 mV to −40 mV), due to different ion channel expression patterns and open probabilities [5]. Membrane potential is coupled to opening of voltage-dependent Ca2+ channels, so depolarization of GI smooth muscle cells leads to Ca2+ influx and initiation of contraction. In regions with more negative membrane potentials, the open probability of Ca2+ channels is very low at the resting potential, and Ca2+ influx mainly occurs during excitable events such as slow waves and action potentials. These areas display phasic contractions generally timed by the frequency of slow waves. Phasic contractions contribute to organ level motility patterns such as peristalsis and segmentation. In regions with less negative potentials, there is low but regular openings of Ca2+ channels and continuous influx of Ca2+ (window current). These regions of muscle are characterized by tonic contractions and relaxation and contribute to motility patterns such as sphincter regulation and storage functions.

The wide range of resting potentials and electrical patterns of GI smooth muscles is mostly due to variable K+ channel expression of smooth muscle cells [6]. However, non-selective cation channels also contribute to the diversity in resting membrane potentials by shifting membrane potentials positive to the equilibrium potential of K+ ions. Several Ca2+-sensitive K+ channels expressed on smooth muscle cells are important determinants of GI smooth muscle excitability [2]. Investigations of the molecular basis of the Ca2+ sensitivity of these channels revealed a component regulated by phosphorylation [79]. CaMKII, a Ca2+/CaM-stimulated, multifunctional Ser/Thr protein kinase, modulates the activities of several Ca2+-sensitive proteins, enzymes, and ion channels [1012]. Several studies have now clearly demonstrated the contribution of CaMKII in regulating the activities of Ca2+ sensitive K+ channels expressed by GI smooth muscles [8,9,1315].

Structure-function of the multifunctional Ca2+/calmodulin-stimulated protein kinase II

It is now evident that the amplitude and frequency of cytosolic Ca2+ oscillations encodes information that is translated into physiological responses [1618]. CaMKII has the physical, catalytic, and regulatory properties required of a molecular decoder of Ca2+ oscillations [19]. Understanding how CaMKII responds to cytosolic Ca2+ transients is critical to understanding how it modulates GI smooth muscle contractile activity. The present review briefly summarizes our current knowledge of the physiological roles of CaMKII in regulating GI motility. Considering the extensive studies demonstrating the importance of CaMKII to cardiac muscle and vascular smooth muscle physiology, and the well established and critical roles of CaMKII for the function of central nervous system neurons, there is a lack of corresponding comprehensive and focused studies of CaMKII in GI smooth muscles and the enteric nervous system. However, recent studies have begun to delineate the various proteins and ion channels regulated by this Ca2+ sensitive multifunctional Ser/Thr protein kinase in GI smooth muscles.

CaMKII is ubiquitously expressed. It is a family of the closely related α, β, γ, δ isoforms, which are the products of four separate genes [20]. At the mRNA level, different tissues express different patterns of CaMKII subunit isoforms [2128]. The α isoform is restricted to neural tissues, while the β, γ and δ isoforms are widely distributed [29]. The CaMKII holoenzyme is a dodecamer of 12 kinase subunits [3032]. Each kinase subunit consists of an N-terminal catalytic/regulatory domain and a C-terminal association domain, connected by a linker region containing a number of discrete variable domains (Figure 1). The variable domains undergo alternative splicing, giving rise to the 38 known mammalian isozymes [29]. The variable domains confer unique regulatory properties to the holoenzyme. For example, variable domain 3, which is only found in γ and δ splice variants, contains a nuclear import signal, allowing CaMKII holoenzymes containing these splice variants to enter the nucleus and regulate transcription factor activity [12].

Figure 1. Structure of the CaMKII holoenzyme.

Figure 1

A. Domain organization of the kinase subunits. B. Surface view from above depicts the 12 kinase subunits arranged as dimers in a ring that is coplanar with the central hub as revealed by small-angle X-ray scattering and X-ray diffraction data [25].

The neuronal αCaMKII gene encodes three splice variants [33]. αCaMK-II is important in learning and memory, but is not necessary during embryonic development [33].αCaMKII knockout mice display defects in specific fear responses, spatial learning and spatial memory [3436]. The expression of αCaMKII in the enteric nervous system has not been examined.

βCaMKII is encoded as seven splice variants, none of which possess nuclear targeting sequences [29]. βCaMKII knockout mice are embryonic lethal [37]. βCaMKII expression is strongest in the adult brain, and is enriched in the cerebellum [38]. βCaMKII is also expressed in skeletal muscle, small intestine, and in endocrine tissues, such as the pancreas, adrenal, and pituitary [20,39]. We found weak expression of α and β CaMKII mRNA in gastric fundus and proximal colon smooth muscle tissues, but not in pure smooth muscle cells, suggesting that these two isoforms may be expressed in other cell types, such as enteric neurons [40].

The γCaMKII gene encodes 13 known splice variants [29]. γCaMKII is ubiquitously expressed, with highest expression in brain, oocytes and skeletal muscle and lowest in liver, testis, lung, thymus, and kidney [20,41]. It is essential for oocyte activation by regulating cell cycle resumption [41]. Multiple roles have been attributed to γCaMKII isozymes in the immune and endocrine systems [29].

The δCaMKII gene encodes at least 15 splice variants [29]. δCaMKIIs are important in embryogenesis and in cardiac and neuronal morphogenesis [33,42]. δCaMKII is highly expressed in brain, skeletal, vascular, and cardiac muscle, with significant quantities in intestine, liver, and lung [20,25,33,42,43]. In the heart δCaMKII phosphorylates several important Ca2+ handling proteins and channels with multiple functional consequences for cardiac myocytes [10].δCaMKII activity and expression are increased in cardiac hypertrophy, human heart failure, contributing to cardiac disease through excitation-transcription coupling [4446].

The different CaMKII isoforms can co-assemble into heteromeric holoenzymes in vitro, raising the possibility that both homomultimers and heteromultimers of CaMKII exist in vivo [4750]. However, except for brain, the subunit composition of CaMKII has not been determined in the tissues where its isoform expression has been investigated. Since the different subunit isoforms have different affinities for Ca2+/CaM, the holoenzyme composition also modulates the response of CaMKII to Ca2+ oscillations [19,23,51]. For example, homomeric CaMKII composed of either the α or β isoforms exhibit distinct variations in autophosphorylation and substrate phosphorylation rates [31]. These results suggest that different tissues express different subunit isoforms in order to assemble heteromeric CaMKII holoenzymes that optimally respond to specific frequencies and amplitudes of Ca2+ oscillations. However, except for brain, the subunit composition of CaMKII holoenzymes has not been determined in the tissues where its isoform expression has been investigated. We found that murine GI smooth muscle cells express γB, γC, γI, γJ, δA, δB, δC, δD, and δE [40]. The γ:δ CaMKII ratios are different in fundus and proximal colon smooth muscles, suggesting that these two functionally distinct GI smooth muscles express CaMKII holoenzymes with different subunit compositions and functional properties [40].

Early studies of the CaMKII three-dimensional structure suggested that the CaMKII holoenzyme consisted of two ‘wheels’ of six catalytic and regulatory domains tethered together by the twelve association domains (hub-and-spoke model) [31,52]. Other analyses suggested that there is actually an outer ring of 12 catalytic and regulatory domains, rather than two ‘wheels’ of six [53]. Higher resolution SAXS (small angle X-ray scattering) resolved this controversy, which showed that the CaMKII three-dimensional structure is highly unusual, with the 12 catalytic and regulatory domains organized into dimers, which are arranged in an outer ring that is coplanar with the ring formed by the association domains [54] (Figure 1). Another feature of the paired catalytic domain structure is that the ATP-binding site is occluded, and the T286 autophosphorylation site that confers calcium-independence to the enzyme, is blocked by the dimer interface [32,54]. Thus, catalytic domains dimers must separate to allow ATP access for kinase activity and to allow autophosphorylation for persistent activation by T286 autophosphorylation [32,54]. This structure also suggests that T286 autophosphorylation is not required for kinase activation, but subsequently prevents the reformation of inactive catalytic domain dimers [30,32]. This structural arrangement could allow for a circular array of switches around the holoenzyme with each of the six catalytic domain pairs acts as a separate “switch” [30,32]. The number of paired catalytic domains activated by Ca2+/CaM in each holoenzyme could potentially encode graded information [19,32]. In addition, the pattern of paired and unpaired catalytic dimers within a holoenzyme could have physiological significance by regulating subcellular targeting [30,32].

Following an increase in Ca2+, CaMKII is activated and autophosphorylated [55]. The holoenzyme kinase subunits are activated by Ca2+/CaM binding, resulting in phosphorylation of target substrates, and rapid intersubunit autophosphorylation of several sites [18,55]. Autophosphorylation of Thr286 (numbering according to α isoform, Thr287 for β, γ and δ) by adjacent subunits disrupts the interaction of the autoinhibitory domain with the catalytic domain, and converts the holoenzyme into a Ca2+/CaM independent kinase (autonomous activity) that allows CaMKII to continue to phosphorylate its substrates after cytosolic Ca2+ levels decrease [11,55]. In addition, the affinity of CaMKII for CaM increases about 1000-fold, a phenomenon termed ‘CaM trapping’ [56]. The ability of CaMKII to retain activity when Ca2+ levels decrease following prior activation by Ca2+ provides a form of ‘molecular memory’ that has been linked, for the α and β isoforms, to the processes of synaptic plasticity, learning, and memory [18,19]. The role of CaMKII in learning and memory is strongly supported by the αCaMKII knockout mouse model [57]. A modified form of memory has been linked to a prolongation of vascular tone in vascular smooth muscle [58]. Ca2+/CaM dissociation from the Thr286 phosphorylated protein allows subsequent autophosphorylation at Thr305, Thr306 and Ser314 to occur [59]. Autophosphorylation of Thr305 and Thr306 are inhibitory as they prevent further CaM binding [59]. However, Ser314 autophosphorylation has no effect on CaM binding [59]. Thus, the Thr286, and Thr305/306 autophosphorylation status further influences the response of the holoenzyme to Ca2+/CaM [18,19]. The physiological roles of autonomous CaMKII activity in GI smooth muscle cells is unclear, but may provide a mechanism for triggering changes in gene expression involved in the phenotypic remodeling of GI smooth muscle cells [60,61].

Ion channel regulation by CaMKII in GI smooth muscle cells

A component of the 4-aminopyridine (4-AP)-sensitive, tetraethylammonium (TEA)-insensitive, rapidly inactivating delayed rectifier current plays an important role in regulating the rhythmic electrical activity of the murine proximal colon [62]. This ‘A-like’ K+ current has properties similar to members of the Kv4 family of K+ channels [62]. Inhibition of this A-like current changes the pattern of electrical activity and induces repetitive Ca2+ action potentials [8]. CaMKII has been shown to regulate this delayed rectifier K+ current in murine colonic myocytes [8]. Whole cell studies revealed that the rate of inactivation of the delayed rectifier K+ current in murine colonic myocytes was increased by stronger Ca2+ buffering, and regulated by Ca2+ in the range of 10–100 nM. The CaMKII inhibitors KN-62 or KN-93 (but not the inactive KN-92 analog or PKC inhibitors) increased the rate of inactivation and slowed recovery from inactivation of the TEA-insensitive outward current, while dialysis of myocytes with constitutively active autothiophosphorylated CaMKII slowed the inactivation. The finding that CaMKII is a potent regulator of the TEA-insensitive current in colonic myocytes suggests a mechanism by which CaMKII is involved in regulating the pattern of electrical activity and thus the pattern of phasic contractions. CaMKII would be activated by the rise in cytosolic Ca2+ during bursts of action potentials, and phosphorylate the channel responsible for the A-type K+ current, resulting in a slowing of A-type current inactivation and reduced smooth muscle excitability [8].

A number of studies have suggested that members of the Kv4 family mediate the native A-type current in murine colonic myocytes [7,63]. Kv4.1 and Kv4.3 have consensus CaMKII phosphorylation sequences near the N-terminus of the channel protein which could mediate direct regulation of the Kv4 channels by CaMKII as has been suggested for Kv1.4 channels [64]. These channels undergo N-type inactivation similar to that shown by Kv4 channels and Kv1.4 phosphorylation slows inactivation and accelerates recovery from inactivation [64]. A similar mechanism contributing to the regulation of N-type inactivation of cloned Kv channels and the native GI smooth muscle delayed rectifier K+ currents was provided in a study by Sergeant et al., in which HEK-293 cells stably expressing the Kv4.3 channel were used to investigate the regulation of Kv4.3 currents by CaMKII [15]. Dialysis of cells with autothiophosphorylated CaMKII decreased the rateof inactivation and increased the rate of recovery from inactivation, while dialysis of cells with the CaMKII inhibitory peptide increased the rate of inactivation of Kv4.3 currents and decreased the recovery time from inactivation. KN-93 also increased the rate of inactivation of Kv4.3 currents and slowed the recovery from inactivation. CaMKII regulation of Kv4.3 currents appears to be mediated by direct phosphorylation of the pore-forming α-subunit. Site-directed mutation at Ser550 (i.e., S550A), a CaMKII consensus site near the COOH terminus, resulted in currents that were unaffected by dialysis with either autothiophosphorylated CaMKII or the CaMKII inhibitory peptide, suggesting that following activation by Ca2+/CaM, CaMKII slows the inactivation and accelerates the rate of recovery from inactivation of Kv4.3 currents by phosphorylation of Ser550. In addition, these currents inactivated more rapidly and recovered from inactivation at a slower rate than that of wild-type controls, suggesting that basal phosphorylation by autonomous CaMKII regulates the kinetics of wild-type Kv4.3 channels [15]. Studies have demonstrated that Kv4.3 is likely the major molecular determinant of the A-type current in GI smooth muscle cells and that this conductance participates in the maintenance of the resting potential [7,63]. Activation of Kv4.3 generates outward currents that oppose membrane depolarization and tend to stabilize membrane potential or reduce excitability [15]. Thus, CaMKII activation by elevations in cytosolic Ca2+ would slow Kv4.3 current inactivation and shift the window current range toward more positive potentials. CaMKII phosphorylation of the Kv4.3 pore-forming α-subunit would be expected to moderate membrane potential depolarizations in GI smooth muscle cells that use Kv4.3 currents to regulate membrane potential.

Another ion channel regulated by CaMKII is the small conductance Ca2+-activated K+ (SK) channel [9]. The inhibitory neurotransmitter ATP mediates a significant portion of the inhibitory regulation of GI smooth muscles [65]. Post-junctional responses to purinergic inhibition occur via stimulation of P2Y receptors and activation of SK channels [66,67]. SK channel activation leads to outward current and reduced membrane excitability [9]. Localized Ca2+ release from IP3 receptor-operated (IP3R) stores in GI smooth muscle cells activates SK channels in response to ATP release [9]. SK channels transduce intracellular Ca2+ transients into changes in membrane potential to regulate membrane excitability and the open probability of voltage-dependent Ca2+ channels [68]. Under whole-cell patch-clamp conditions of colonic myocytes, SK channel openings are resolved as charybdotoxin-insensitive spontaneous transient outward currents (STOCs) [9]. In excised patches, autothiophosphorylated CaMKII increased the open probability of SK channels while boiled CaMKII had no effect. KN-93 reduced the frequency of charybdotoxin-insensitive STOCs, but did not affect spontaneous Ca2+-release events. SK2, the predominant SK channel isoform expressed by GI smooth muscles, has four potential CaMKII phosphorylation sites suggesting that CaMKII affects SK channel by direct phosphorylation [9]. CaMKII regulation of SK channel activity in GI smooth muscle cells indicates its role in the neurogenic regulation of colonic motility, since these channels are typically activated by purines, which serve as enteric inhibitory motor neurotransmitters. Regulation of SK channel open probability by CaMKII provides a mechanism for regulation complementary to local changes in Ca2+ concentration, increasing the effect of a given Ca2+ transient on SK channel open probability, and leading to enhanced SK channel open probability and facilitation of channel openings and inhibitory responses with repetitive stimulation [9].

Regulation of intracellular Ca2+ transport by CaMKII

While extracellular Ca2+ ions enter smooth muscle cells mainly through L-type Ca2+ channels during muscle action potentials, intracellular storage and release of Ca2+ can occur during and between action potentials [69,70]. The sarcoplasmic reticulum (SR) is the main organelle in smooth muscle cells that stores and releases Ca2+ [71]. In GI smooth muscles, the uptake of Ca2+ by intracellular stores tends to reduce excitability and contractility by removing Ca2+ from the myoplasm [72]. Ca2+ release from the SR, however, can also reduce excitability by activating Ca2+-sensitive K+ channels in the plasma membrane, hyperpolarizing cells and reducing voltage-dependent Ca2+ entry [9]. Ca2+ release events occur from either IP3Rs (Ca2+ puffs), ryanodine receptors (RyR) (Ca2+ sparks), or from both types of receptors due to their Ca2+-sensitive properties and expression profiles [73,74]. Ca2+ puffs and sparks initiate STOCs in GI smooth muscles, affecting membrane potential and excitability [75,76]. Both SK and BK channels contribute to STOCs [77]. Ca2+ transients originating from the SR can also develop into intracellular Ca2+ waves that appear to propagate along a peripheral cellular compartment without activating contractile proteins [78]. Instead, the intracellular Ca2+ waves arising from Ca2+ puffs or sparks can hyperpolarize cells by increasing the open probability of Ca2+ sensitive K+ channels [14].

The level of luminal SR Ca2+ depends on the balance between Ca2+ efflux through Ca2+ release channels and Ca2+ influx through SERCA [71,79]. SERCA contributes to relaxation by lowering the [Ca2+]i and decreasing the amount of Ca2+ available for contraction [80,81]. In addition, elevations in luminal SR Ca2+ load by SERCA increases the frequency of Ca2+ sparks and or puffs, which also contributes to relaxation by increasing the frequency of STOCs [82]. SERCA activity is under dynamic control by the SR membrane protein PLN to continuously vary the rate of Ca2+ influx into the SR via its phosphorylation levels, from maximal SERCA inhibition of 50% by dephospho-PLN to maximal SERCA activity by phospho-PLN [83,84]. Phosphorylation of PLN by cAMP-dependent protein kinase (PKA) or cGMP-dependent protein kinase (PKG) at Ser16, or by CaMKII at Thr17 removes its inhibitory effects on SERCA, thereby increasing Ca2+ uptake into the SR and contributing to relaxation [8587]. Because the frequency and amplitudes of localized Ca2+ release events through ryanodine- or IP3-sensitive Ca2+ channels is in the range that will activate CaMKII [88], a number of investigations have examined the regulation of CaMKII activity and PLN phosphorylation by localized Ca2+ release events from internal Ca2+ stores of GI smooth muscles.

Gastric accommodation is a postprandial, vagally mediated reflex resulting in reduced tone of the proximal stomach that allows the proximal stomach to expand during food intake with no change in pressure [89]. Insufficient gastric accommodation is a frequent pathophysiological mechanism underlying functional dyspepsia [90]. Investigating the intrinsic excitation-contraction mechanisms of gastric fundus smooth muscles is necessary to enhance our understanding of stomach smooth muscle physiology and the pathophysiology of gastric motility disorders [91]. In gastric fundus smooth muscles, caffeine increases the frequency of ryanodine-sensitive Ca2+ release and causes relaxation that is sensitive to the Ca2+-sensitive K+ channel (KCa) blockers TEA and iberiotoxin, suggesting that membrane hyperpolarization and STOCs induced by large-conductance Ca2+-activated K+ channels contribute to fundus smooth muscle relaxation and gastric accommodation [9294]. Increasing the luminal SR Ca2+ content increases RyR open probability, resulting in increased Ca2+ spark frequency [77,82,87,95]. SERCA activity and luminal SR Ca2+ content are increased by cAMP-dependent protein kinase (PKA) or CaMKII phosphorylation of PLN, suggesting that CaMKII activation and PLN phosphorylation are involved in gastric fundus smooth muscle relaxation by ryanodine-sensitive Ca2+ release [8587].

In murine gastric fundus smooth muscles, caffeine was found to activate CaMKII in a tetracaine- and ryanodine-sensitive manner, suggesting that CaMKII activation involves increased Ca2+ spark activity [96]. Caffeine also increased PLN Thr17 phosphorylation, indicating that PLN is a substrate of CaMKII. The SERCA inhibitor cyclopiazonic acid (CPA) prevented caffeine-induced CaMKII activation and PLN phosphorylation, suggesting that SR luminal Ca2+ levels were lowered and insufficient to support caffeine-induced Ca2+ spark activity and CaMKII activation [9698]. In addition, the caffeine-induced relaxation and PLN Thr17 phosphorylation were inhibited by KN-93 [96]. These findings suggest that KN-93 inhibition of CaMKII and PLN Thr17 phosphorylation mimics the effects of SERCA inhibition. These results suggest that the inhibition of caffeine-induced relaxation of gastric fundus smooth muscle by KN-93 involves inhibition of PLN Thr17 phosphorylation by CaMKII, resulting in SERCA inhibition and decreased Ca2+ spark activity. The findings that relaxation of gastric fundus smooth muscles by caffeine is ryanodine sensitive, and involves PLN Thr17 phosphorylation by CaMKII point to a role for both increased Ca2+ clearance from the cytosol and increased Ca2+ spark activity in gastric fundus smooth muscle relaxation. These findings provide further insights into the Ca2+-sensitive processes that modulate excitation-contraction of tonic GI smooth muscles.

It is well established that the relaxant effects of nitric oxide (NO) and NO donors on smooth muscles are mediated by cGMP, but the mechanisms of NO-induced relaxation appear to differ between different smooth muscle types [87,99]. The NO donor sodium nitroprusside (SNP) increases Ca2+ spark frequency and activates STOCs via an NO-soluble guanylyl cyclase (sGC)-cGMP pathway in cerebral and coronary artery myocytes [95]. Similarly, SNP-induced relaxation of the rat gastric fundus is partially ryanodine sensitive and involves small-conductance KCa channels, while in guinea pig gastric antral myocytes, SNP increases KCa current and enhances STOCs, which are sensitive to IP3R inhibitors [100,101]. However, studies have also provided evidence that NO-induced smooth muscle relaxation involves uptake of myoplasmic Ca2+ into SR stores by increasing SERCA activity [102105]. PLN phosphorylation by PKG, PKA, or calmodulin (CaM) kinase II eliminates its inhibitory effects on SERCA, thereby activating Ca2+ uptake into the SR and lowering cytosolic Ca2+ levels, causing relaxation [81,106]. However, SR Ca2+ load is also an important modulator of the frequency of Ca2+ release events from the SR [77,107]. If SR Ca2+ load is increased by SERCA activation, Ca2+ spark frequency is increased, leading to increased STOC frequency and membrane potential hyperpolarization and relaxation [77]. Removal of Ca2+ from the myoplasm and increasing the SR Ca2+ load and the frequency of Ca2+ release events from the SR may both contribute to relaxation in a complementary manner [72].

The NO donor SNP has been used to address the role of PLN phosphorylation by CaMKII in NO-induced relaxation of the functionally distinct proximal (fundus) and distal (antrum) stomach [108,109]. SNP activated CaMKII, hyperpolarized the cells, and relaxed gastric fundus and antrum circular smooth muscles. Western blot analysis showed that SNP increased PLN Ser16 and Thr17 phosphorylation, suggesting that SNP activated both PKG and CaMKII. The membrane-permeablec GMP analog 8-bromo-cGMP relaxed gastric fundus and antrum smooth muscles and increased PLNSer16 and Thr17 phosphorylation. PLNThr17 phosphorylation, but not PLNSer16 phosphorylation, was inhibited by KN-93 and CPA, suggesting that PLNThr17 phosphorylation by CaMKII involves Ca2+ release from the SR. In contrast, both PLNser16 and Thr17 phosphorylation were inhibited by the soluble guanylate cyclase inhibitor 1H-[1,2,4]oxadiazolo-[4,3-α]quinoxalin-1-one (ODQ), suggesting that CaMKII activation by SR Ca2+ release is mediated by an NO-sGC-cGMP pathway [108,109]. These findings revealed a novel link between nitrergic and Ca2+/CaM-dependent signaling pathways in stomach smooth muscles. In response to cGMP production by sGC, PLNSer16 phosphorylation by PKG activates SERCA, increasing SR Ca2+ uptake and elevating SR Ca2+ levels. Higher SR Ca2+ levels increase the frequency SRCa2+ release events, activating CaMKII and increasing PLNThr17 phosphorylation. In addition, an increased frequency of Ca2+ release events through RyRs and IP3Rs would likely contribute to smooth muscle relaxation by activating KCa channels, resulting in hyperpolarization of the membrane potential. The SNP-induced release of SR Ca2+ may also potentiate the direct regulation of KCa channels by PKG and CaMKII [110,111]. For example, as noted earlier, CaMKII increases A-type current activity in GI smooth muscles by slowing its inactivation, suggesting that SR Ca2+ release could oppose depolarization by activating CaMKII to slow A-type current inactivation [96,108,109]. The physiological significance of CaMKII activation by NO remains unclear; however, its Ca2+-independent activity may provide a mechanism to prolong the effects of NO/cGMPon SERCA and KCa channel activities after the cGMP and Ca2+ signals have ceased.

CaMKII in ulcerative colitis

Ulcerative colitis and Crohn’s disease are inflammatory bowel diseases characterized by chronic intestinal inflammation associated with dysmotility, resulting in symptoms that include abdominal pain, weight loss and bloody diarrhea [112114]. Inflammation is associated with reduced colonic mixing patterns and impaired haustra formation, but increased propulsive motility from propagating contractions, causing rapid movement of the colonic contents to the rectum, accentuating the diarrhea [115118]. Multiple mechanisms involving changes in enteric neurotransmission, afferent sensory input, and intrinsic smooth muscle contractility underlie the dysmotility [116,118].

Improper calcium mobilization by the smooth muscles has been implicated in the development of colonic dysmotility in inflammatory bowel disease. Impaired Ca2+ influx in smooth muscles from the inflamed colon has been reported to occur by reduced L-type Ca2+ channel expression or by decreased L-type Ca2+ channel activity with no change in expression [119,120]. In addition, SERCA expression in rat colon smooth muscle is decreased in experimental colitis, suggesting that intracellular Ca2+ mobilization in smooth muscles is disrupted due to reduced SR Ca2+ uptake [121]. Understanding the molecular mechanisms underlying the disruptions in Ca2+ mobilization and the impaired contractile responses of colon smooth muscles to intestinal inflammation is important for understanding the myriad factors underlying the myogenic basis of the dysmotility that is associated with colitis.

The activities and expression of proteins involved with intracellular Ca2+ mobilization in colon smooth muscle cells are altered by dextran sulfate sodium (DSS)-induced colitis [61]. Functional disruptions are evident by impairments in spontaneous and acetylcholine-evoked contractions of distal colon smooth muscles from mice with DSS-colitis. DSS-colitis also increased the amplitudes, but not the frequency, of intracellular Ca2+ waves in colonic myocytes. The expression of both SERCA and PLN are reduced, suggesting that SERCA and PLN down-regulation lead to disruptions in intracellular Ca2+ mobilization and the altered intracellular Ca2+ wave patterns. In contrast, γCaMKII expression and activity are increased in distal colon smooth muscles from mice with DSS-colitis. Along with the increase in CaMKII activity, cytosolic levels of HDAC4 are increased, and IκBα levels are decreased in distal colon smooth muscles from DSS-treated mice [61]. IκBα degradation is necessary for NF-kB translocation into the nucleus and activation of NF-κB-dependent gene expression [122]. Additional transcriptional co-activators, including histone acetyltransferases (HATs) and histone deacetylases (HDACs) are required for NF-kB activity [122]. Histone acetylation enhances gene expression, while HDACs repress gene transcription by deacetylating histones to allow DNA to re-wind around the histone octamer [122]. HDAC4 repression of gene transcription is regulated, in part, by nucleo-cytoplasmic shuttling [123]. HDAC4 gene repression is neutralized by Ca2+ signals that activate CaMKII to phosphorylate HDAC4 and sequester it in the cytoplasm, allowing its target genes to be de-repressed and transcribed [123]. These findings link alterations in intracellular Ca2+ mobilization and contractile responses with changes in SERCA, PLN, and γCaMKII protein expression in distal colon smooth muscles from mice with DSS-colitis. The findings that IκBα levels are reduced, and cytosolic HDAC4 levels and binding to CaMKII were increased, suggest that CaMKII activation and HDAC4 phosphorylation may facilitate NF-κB-dependent gene transcription that contributes to the reduced contractility of colon smooth muscles as a result of colitis [61].

Summary

Higher regulatory systems (enteric neurons, ICC, and hormones) control and coordinate gut motility. However, GI smooth muscle cells are equipped with intrinsic regulatory pathways that can amplify or oppose signaling from these higher regulatory systems. The smooth muscle cells of the gut wall generate the force and carry out the work of the digestive tract organs; thus it is important to investigate the intrinsic excitation-contraction mechanisms of GI smooth muscles to advance our understanding of GI smooth muscle physiology and the pathophysiology of GI motility disorders.

Calcium sensitive K+ channels involved in setting membrane potentials and the excitability of GI smooth muscle cells are regulated by CaMKII. Kv4.3 is responsible for the A-type K+ current in GI smooth muscle cells with an important role in the maintenance of resting membrane potential [15]. In murine colonic myocytes, phosphorylation of the Kv4.3 pore-forming α-subunit by CaMKII slows the rate of inactivation and accelerates the rate of recovery from inactivation. Thus, during depolarization-induced Ca2+ influx, CaMKII activation would tend to limit the extent of depolarization and moderate smooth muscle excitability. The inhibitory neurotransmitter ATP contributes to the inhibitory regulation of GI smooth muscles. Stimulation of P2Y receptors activates SK channels via localized Ca2+ release from intracellular Ca2+ stores, leading to outward currents and decreased membrane excitability. Activation of CaMKII increases the open probability of SK channels, enhancing the impact of localized Ca2+ transients on SK currents and facilitating the inhibitory effects of purinergic inputs to reduce smooth muscle cell excitability [9].

In gastric fundus and antrum smooth muscles, the NO donor SNP increases PLN phosphorylation and activates CaMKII, revealing a novel contribution of Ca2+/CaM-dependent signaling pathways to nitrergic relaxation in stomach smooth muscles [109]. PLN phosphorylation by CaMKII at Thr17 stimulates SERCA, increasing Ca2+ uptake into the SR [8587]. Ca2+ clearance from the myoplasm into intracellular stores by SERCA tends to reduce excitability and contractility by reducing the availability of Ca2+ for contraction [72]. Elevations in luminal SR Ca2+ load also increase the frequency of Ca2+ sparks and or puffs, resulting in an increase in STOCs which hyperpolarizes the cells and reduces voltage-dependent Ca2+ entry [9,14].

Impaired Ca2+ mobilization by colonic smooth muscles has been implicated in the development of colonic dysmotility in inflammatory bowel disease. The findings that HDAC4 levels and binding to CaMKII are increased, and IκBα levels are decreased in an experimental colitis model suggest that CaMKII activation and HDAC4 phosphorylation may facilitate NF-κB-dependent gene transcription that contributes to the reduced contractility of colon smooth muscles as a result of colitis [61].

Overall, these studies suggest that the effects of CaMKII on the activities of K+ channels, intracellular Ca2+ handling proteins, and transcriptional activity, tend to reduce GI smooth muscle cell excitability. These findings are consistent with the additional regulation of MLCK by CaMKII in smooth muscle cells, in which MLCK phosphorylation by CaMKII decreases the Ca2+ sensitivity of myosin light phosphorylation, as has been reviewed elsewhere [124]. More work is needed to identify other targets and determine how spatio-temporal Ca2+ signaling regulates CaMKII expression, activity, and localization in GI smooth muscle cells. In addition, the expression and roles of CaMKII in ICC has not been studied. These cells are linked to the smooth muscle syncytium, so changes in membrane conductances, gap junction coupling, and pacemaker mechanisms could also be regulated by CaMKII. Finally, while CaMKII isoforms are expressed in enteric neurons little is known about how ionic conductances, Ca2+ handling mechanisms and reflex pathways are regulated by CaMKII in these cells. Sustained and focused studies of CaMKII in GI smooth muscles, ICC, and the enteric nervous system, are needed to obtain a more complete picture of its roles in the gut to more fully evaluate and appreciate its impact on GI motility, and to identify novel regulatory pathways for improved and innovative therapeutic strategies for GI motility disorders.

Acknowledgments

This work was supported by National Institutes of Health grants RR018751. The author is grateful to Kent Sanders for his helpful discussions and critiques during the preparation of this manuscript. The author declares no conflicts of interest.

Reference List

  • 1.Wood JD. Gastrointestinal Physiology. In: Rhoades RA, Tanner GA, editors. Medical Physiology. Lippincott Williams & Wilkins; Baltimore: 2003. [Google Scholar]
  • 2.Sanders KM. Regulation of smooth muscle excitation and contraction. Neurogastroenterology & Motility. 2008;20:39–53. doi: 10.1111/j.1365-2982.2008.01108.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Sanders KM, Koh SD, Ward SM. Interstitial Cells of Cajal as Pacemakers in the Gastrointestianl Tract. Annual Review of Physiology. 2006;68:307–343. doi: 10.1146/annurev.physiol.68.040504.094718. [DOI] [PubMed] [Google Scholar]
  • 4.Huizinga JD, Lammers WJEP. Gut peristalsis is governed by a multitude of cooperating mechanisms. Am J Physiol Gastrointest Liver Physiol. 2009;296:G1–G8. doi: 10.1152/ajpgi.90380.2008. [DOI] [PubMed] [Google Scholar]
  • 5.Lyford GL, Farrugia G. Ion channels in gastrointestinal smooth muscle and interstitial cells of Cajal. Curr Opin Pharmacol. 2003;3:583–587. doi: 10.1016/j.coph.2003.06.010. [DOI] [PubMed] [Google Scholar]
  • 6.Horowitz B, Ward SM, Sanders KM. Cellular and molecular basis for electrical rhythmicity in gastrointestinal muscles. Annu Rev Physiol. 1999;61:19–43. doi: 10.1146/annurev.physiol.61.1.19. [DOI] [PubMed] [Google Scholar]
  • 7.Amberg GC, Baker SA, Koh SD, Hatton WJ, Murray KJ, Horowitz B, Sanders KM. Characterization of the A-type potassium current in murine gastric antrum. J Physiol (Lond) 2002;544:417–428. doi: 10.1113/jphysiol.2002.025171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Koh SD, Perrino BA, Hatton WJ, Kenyon JL, Sanders KM. Novel regulation of the A-type current in murine proximal colon by calcium/calmodulin-dependent protein kinase II. J Physiol. 1999;517.1:75–84. doi: 10.1111/j.1469-7793.1999.0075z.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Kong ID, Koh SD, Bayguinov O, Sanders KM. Small conductance Ca2+-activated K+ channels are regulated by Ca2+-calmodulin-dependent protein kinase II in murine colonic myocytes. J Physiol. 2000;524.2:331–337. doi: 10.1111/j.1469-7793.2000.t01-1-00331.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bers DM, Grandi E. Calcium/calmodulin-dependent kinase II regulation of cardiac ion channels. J Cardiovasc Pharmacol. 2009;54:180–187. doi: 10.1097/FJC.0b013e3181a25078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kim HR, Appel S, Vetterkind S, Gangopadhyay SS, Morgan KG. Smooth muscle signalling pathways in health and disease. J Cell Mol Med. 2008;12:2165–2180. doi: 10.1111/j.1582-4934.2008.00552.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Skelding KA, Rostas JA. Regulation of CaMKII in vivo: the importance of targeting and the intracellular microenvironment. Neurochem Res. 2009;34:1792–1804. doi: 10.1007/s11064-009-9985-9. [DOI] [PubMed] [Google Scholar]
  • 13.Kim M, Hennig GW, Smith TK, Perrino BA. Phospholamban knockout increases CaM kinase II activity and intracellular Ca2+ wave activity and alters contractile responses of murine gastric antrum. Am J Physiol Cell Physiol. 2008;294:C432–C441. doi: 10.1152/ajpcell.00418.2007. [DOI] [PubMed] [Google Scholar]
  • 14.Kim M, Hennig GW, Park K, Han IS, Smith TK, Koh SD, Perrino BA. Modulation of murine gastric antrum smooth muscle STOC activity and excitability by phospholamban. J Physiol. 2008;586:4977–4991. doi: 10.1113/jphysiol.2008.156836. PMC Journal - In Process. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Sergeant GP, Ohya S, Reihill JA, Perrino BA, Amberg GC, Imaizumi Y, Horowitz B, Sanders KM, Koh SD. Regulation of Kv4.3 currents by Ca2+/calmodulin-dependent protein kinase II. Am J Physiol Cell Physiol. 2005;288:C304–C313. doi: 10.1152/ajpcell.00293.2004. [DOI] [PubMed] [Google Scholar]
  • 16.Berridge MJ. The AM and FM of calcium signalling. Nature. 1997;386:759–760. doi: 10.1038/386759a0. [DOI] [PubMed] [Google Scholar]
  • 17.Clapham DE. Calcium signaling. Cell. 1995;80:268. doi: 10.1016/0092-8674(95)90408-5. [DOI] [PubMed] [Google Scholar]
  • 18.Hudmon A, Schulman H. Structure-function of the multifunctional Ca2+/calmodulin-dependent protein kinase II. Biochem J. 2002;364:593–611. doi: 10.1042/BJ20020228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.DeKoninck P, Schulman H. Sensitivity of CaM kinase II to the frequency of Ca2+ oscillations. Science. 1998;279:227–230. doi: 10.1126/science.279.5348.227. [DOI] [PubMed] [Google Scholar]
  • 20.Tobimatsu T, Fujisawa H. Tissue-specific expression of four types of rat calmodulin-dependent protein kinase II mRNAs. J Biol Chem. 1989;264:17907–17912. [PubMed] [Google Scholar]
  • 21.Burgin KE, Waxham MN, Rickling S, Westgate SA, Mobley WC, Kelly PT. In situ hybridization histochemistry of Ca2+/calmodulin-dependent protein kinase in developing rat brain. J Neurosci. 1990;10:1788–1798. doi: 10.1523/JNEUROSCI.10-06-01788.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Kwiatkowski AP, McGill RL. Human biliary epithelial cell line Mz-ChA- 1 expresses new isoforms of calmodulin-dependent protein kinase II. Gastroenterology. 1995;109:1316–1323. doi: 10.1016/0016-5085(95)90594-4. [DOI] [PubMed] [Google Scholar]
  • 23.Kwiatkowski AP, McGill JM. Alternative splice variant of γ-calmodulin-dependent protein kinase II alters activation by calmodulin. Archives of Biochemistry and Biophysics. 2000;378:377–383. doi: 10.1006/abbi.2000.1846. [DOI] [PubMed] [Google Scholar]
  • 24.Schulman H. The multifunctional Ca 2+calmodulin-dependent protein kinases. Curr Opin Cell Biol. 1993;5:247–253. doi: 10.1016/0955-0674(93)90111-3. [DOI] [PubMed] [Google Scholar]
  • 25.Singer HA, Benscoter HA, Schworer CM. Novel Ca2+/Calmodulin-dependent protein kinase II y-subunit variants expressed in vascular smooth muscle, brain, and cardiomyocytes. J Biol Chem. 1997;272:9393–9400. doi: 10.1074/jbc.272.14.9393. [DOI] [PubMed] [Google Scholar]
  • 26.Takeuchi M, Fujisawa H. New alternatively spliced variants of calmodulin-dependent protein kinase from rabbit liver. Gene. 1998;221:107–115. doi: 10.1016/s0378-1119(98)00422-3. [DOI] [PubMed] [Google Scholar]
  • 27.Vallano ML, Beaman-Hall CM, Mathur A, Chen Q. Astrocytes express specific variants of CaM KII δ and γ, but not α and β, that determine their cellular localizations. Glia. 2000;30:154–164. doi: 10.1002/(sici)1098-1136(200004)30:2<154::aid-glia5>3.0.co;2-s. [DOI] [PubMed] [Google Scholar]
  • 28.Zhou ZL, Ikebe M. New isoforms of Ca2+/calmodulin-dependent protein kinase II in smooth muscle. Biochem J. 1994;299:489–495. doi: 10.1042/bj2990489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Tombes RM, Faison MO, Turbeville JM. Organization and evolution of multifunctional Ca(2+)/CaM-dependent protein kinase genes. Gene. 2003;322:17–31. doi: 10.1016/j.gene.2003.08.023. [DOI] [PubMed] [Google Scholar]
  • 30.Chao LH, Pellicena P, Deindl S, Barclay LA, Schulman H, Kuriyan J. Intersubunit capture of regulatory segments is a component of cooperative CaMKII activation. Nat Struct Mol Biol. 2010;17:264–272. doi: 10.1038/nsmb.1751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Gaertner TR, Kolodziej SJ, Wang D, Kobayashi R, Koomen JM, Stoops JK, Waxham MN. Comparative analyses of the three-dimensional structures and enzymatic properties of alpha, beta, gamma and delta isoforms of Ca2+-calmodulin-dependent protein kinase II. J Biol Chem. 2004;279:12484–12494. doi: 10.1074/jbc.M313597200. [DOI] [PubMed] [Google Scholar]
  • 32.Thaler C, Koushik SV, Puhl HL, Blank PS, Vogel SS. Structural rearrangement of CaMKII catalytic domains encodes activation. PNAS. 2009;106:6369–6374. doi: 10.1073/pnas.0901913106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Bayer KU, Lohler J, Schulman H, Harbers K. Developmental expression of the CaM kinase II isoforms: ubiquitous gamma- and delta-CaM kinase II are the early isoforms and most abundant in the developing nervous system. Brain Res Mol Brain Res. 1999;70:147–154. doi: 10.1016/s0169-328x(99)00131-x. [DOI] [PubMed] [Google Scholar]
  • 34.Chen C, Rainnie DG, Greene RW, Tonegawa S. Abnormal fear response and aggressive behavior in mutant mice deficient for alpha-calcium-calmodulin kinase II. Science. 1994;266:291–294. doi: 10.1126/science.7939668. [DOI] [PubMed] [Google Scholar]
  • 35.Silva AJ, Paylor R, Wehner JM, Tonegawa S. Impaired spatial learning in alpha-calcium-calmodulin kinase II mutant mice. Science. 1992;257:206–211. doi: 10.1126/science.1321493. [DOI] [PubMed] [Google Scholar]
  • 36.Silva AJ, Wang Y, Paylor R, Wehner JM, Stevens CF, Tonegawa S. Alpha calcium/calmodulin kinase II mutant mice: deficient long-term potentiation and impaired spatial learning. Cold Spring Harb Symp Quant Biol. 1992;57:527–539. doi: 10.1101/sqb.1992.057.01.058. [DOI] [PubMed] [Google Scholar]
  • 37.Karls U, Muller U, Gilbert DJ, Copeland NG, Jenkins NA, Harbers K. Structure, expression, and chromosome location of the gene for the beta subunit of brain-specific Ca2+/calmodulin-dependent protein kinase II identified by transgene integration in an embryonic lethal mouse mutant. Mol Cell Biol. 1992;12:3644–3652. doi: 10.1128/mcb.12.8.3644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.McGuiness TL, Lai Y, Greengard P. Ca2+/calmodulin-dependent protein kinase II. Isozymic forms from rat forebrain and cerebellum. J Biol Chem. 1985;260:1696–1704. [PubMed] [Google Scholar]
  • 39.Rochlitz H, Voigt A, Lankat-Buttgereit B, Goke B, Heimberg H, Nauck MA, Schiemann U, Schatz H, Pfeiffer AF. Cloning and quantitative determination of the human Ca2+/calmodulin-dependent protein kinase II (CaMK II) isoforms in human beta cells. Diabetologia. 2000;43:465–473. doi: 10.1007/s001250051330. [DOI] [PubMed] [Google Scholar]
  • 40.Lorenz JM, Riddervold MH, Beckett EAH, Baker SA, Perrino BA. Differential autophosphorylation of Ca2+/calmodulin-dependent protein kinase II from phasic and tonic smooth muscle tissues. Am J Physiol. 2002;283:C1399–C1413. doi: 10.1152/ajpcell.00020.2002. [DOI] [PubMed] [Google Scholar]
  • 41.Backs J, Stein P, Backs T, Duncan FE, Grueter CE, McAnally J, Qi X, Schultz RM, Olson EN. The gamma isoform of CaM kinase II controls mouse egg activation by regulating cell cycle resumption. Proc Natl Acad Sci U S A. 2010;107:81–86. doi: 10.1073/pnas.0912658106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Maier LS, Bers DM. Role of Ca2+/calmodulin-dependent protein kinase (CaMK) in excitation–contraction coupling in the heart. Cardiovascular Research. 2007;73:631–640. doi: 10.1016/j.cardiores.2006.11.005. [DOI] [PubMed] [Google Scholar]
  • 43.Schworer CM, Rothblum LI, Thekkumkara TJ, Singer HA. Identification of novel isoforms of the δ subunit of Ca2+/calmodulin-dependent protein kinase II. J Biol Chem. 1993;268:14443–14449. [PubMed] [Google Scholar]
  • 44.Backs J, Backs T, Neef S, Kreusser MM, Lehmann LH, Patrick DM, Grueter CE, Qi X, Richardson JA, Hill JA, Katus HA, Bassel-Duby R, Maier LS, Olson EN. 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;106:2342–2347. doi: 10.1073/pnas.0813013106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Sag CM, Wadsack DP, Khabbazzadeh S, Abesser M, Grefe C, Neumann K, Opiela MK, Backs J, Olson EN, Brown JH, Neef S, Maier SK, Maier LS. Calcium/calmodulin-dependent protein kinase II contributes to cardiac arrhythmogenesis in heart failure. Circ Heart Fail. 2009;2:664–675. doi: 10.1161/CIRCHEARTFAILURE.109.865279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Sossalla S, Fluschnik N, Schotola H, Ort KR, Neef S, Schulte T, Wittkopper K, Renner A, Schmitto JD, Gummert J, El-Armouche A, Hasenfuss G, Maier LS. Inhibition of elevated Ca2+/calmodulin-dependent protein kinase II improves contractility in human failing myocardium. Circ Res. 2010;107:1150–1161. doi: 10.1161/CIRCRESAHA.110.220418. [DOI] [PubMed] [Google Scholar]
  • 47.Bennet MK, Erondu NE, Kennedy MB. Purification and characterization of a calmodulin-dependent protein kinase that is highly concentrated in brain. J Biol Chem. 1983;258:12735–12744. [PubMed] [Google Scholar]
  • 48.Brocke L, Chiang LW, Wagner PD, Schulman H. Functional implications of the subunit composition of neuronal CaM kinase II. J Biol Chem. 1999;274:22713–22722. doi: 10.1074/jbc.274.32.22713. [DOI] [PubMed] [Google Scholar]
  • 49.Kolb SJ, Hudmon A, Ginsberg TR, Waxham MN. Identification of Domains Essential for the Assembly of Calcium/Calmodulin-dependent Protein Kinase II Holoenzymes. J Biol Chem. 1998;273:31555–31564. doi: 10.1074/jbc.273.47.31555. [DOI] [PubMed] [Google Scholar]
  • 50.Srinivasan M, Edman CF, Schulman H. Altemative splicing introduces a nuclear localization signal that targets multifunctional CaM kinase to the nucleus. J Cell Biol. 1994;126:839–852. doi: 10.1083/jcb.126.4.839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Soderling TR, Chang B, Brickey D. Cellular signaling through multifunctional Ca2+/calmodulin-dependent protein kinase II. J Biol Chem. 2001;276:3719–3722. doi: 10.1074/jbc.R000013200. [DOI] [PubMed] [Google Scholar]
  • 52.Kanaseki T, Ikeuchi Y, Sugiura H, Yamauchi T. Structural features of Ca2+/calmodulin-dependent protein kinase II revealed by electron microscopy. The Journal of Cell Biology. 1991;115:1049–1060. doi: 10.1083/jcb.115.4.1049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Morris EP, Török K. Oligomeric structure of alpha-calmodulin-dependent protein kinase II. J Mol Biol. 2001;308:1–8. doi: 10.1006/jmbi.2001.4584. [DOI] [PubMed] [Google Scholar]
  • 54.Rosenberg OS, Deindl S, Sung RJ, Nairn AC, Kuriyan J. Structure of the autoinhibited kinase domain of CaMKII and SAXS analysis of the holoenzyme. Cell. 2005;123:849–860. doi: 10.1016/j.cell.2005.10.029. [DOI] [PubMed] [Google Scholar]
  • 55.Lucic V, Greif GJ, Kennedy MB. Detailed state model of CaMKII activation and autophosphorylation. Eur Biophys J. 2008;38:83–98. doi: 10.1007/s00249-008-0362-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Meyer T, Hanson PI, Schulman H. Calmodulin trapping by calcium-calmodulin-dependent protein kinase. Science. 1992;256:1199–1202. doi: 10.1126/science.256.5060.1199. [DOI] [PubMed] [Google Scholar]
  • 57.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;411:309–313. doi: 10.1038/35077089. [DOI] [PubMed] [Google Scholar]
  • 58.Munevar S, Gangopadhyay S, Gallant C, Colombo B, Sellke F, Morgan K. CaMKIIT287 and T305 regulate history-dependent increases in agonist–induced vascular tone. Journal of Cellular and Molecular Medicine. 2008;12:219–226. doi: 10.1111/j.1582-4934.2007.00202.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Colbran RJ, Soderling TR. Calcium/calmodulin-independent autophosphorylation sites of calcium/calmodulin-dependent protein kinase II. Studies on the effect of phosphorylation of threonine 305/306 and serine 314 on calmodulin binding using synthetic peptides. J Biol Chem. 1990;265:11213–11219. [PubMed] [Google Scholar]
  • 60.Chen J, Chen H, Sanders KM, Perrino BA. Regulation of SRF/CArG-dependent gene transcription during chronic partial obstruction of murine small intestine. Neurogastroenterol Motil. 2008;20:829–842. doi: 10.1111/j.1365-2982.2008.01149.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Qureshi S, Song J, Lee HT, Koh SD, Hennig GW, Perrino BA. CaM kinase II in colonic smooth muscle contributes to dysmotility in murine DSS-colitis. Neurogastroenterol Motil. 2010;22:186–95. e64. doi: 10.1111/j.1365-2982.2009.01406.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Koh SD, Ward SM, Dick GM, Epperson A, Bonner HP, Sanders KM, Horowitz B, Kenyon JL. Contribution of delayed rectifier potassium currents to the electrical activity of murine colonic smooth muscle. J Physiol. 1999;515(Pt 2):475–487. doi: 10.1111/j.1469-7793.1999.475ac.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Amberg GC, Koh SD, Hatton WJ, Murray KJ, Monaghan K, Horowitz B, Sanders KM. Contribution of Kv4 channels toward the A-type potassium current in murine colonic myocytes. J Physiol. 2002;544:403–415. doi: 10.1113/jphysiol.2002.025163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Roeper J, Lorra C, Pongs O. Frequency-dependent inactivation of mammalian A-type K+ channel KV1.4 regulated by Ca2+/calmodulin-dependent protein kinase. J Neurosci. 1997;17:3379–3391. doi: 10.1523/JNEUROSCI.17-10-03379.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Hoyle CVH, Burnstock G. Neuromuscular transmission in the gastrointestinal tract, The Alimentary Canal, Motility and Circulation. American Physiological Society. 1989:435–464. [Google Scholar]
  • 66.Koh SD, Dick GM, Sanders KM. Small-conductance Ca2+-dependent K+ channels activated by ATP in murine colonic smooth muscle. Am J Physiol Cell Physiol. 1997;273:C2010–C2021. doi: 10.1152/ajpcell.1997.273.6.C2010. [DOI] [PubMed] [Google Scholar]
  • 67.Vogalis F, Goyal RK. Activation of small conductance Ca2+-dependent K+ channels by purinergic agonists in smooth muscle cells of the mouse ileum. J Physiol. 1997;502(Pt 3):497–508. doi: 10.1111/j.1469-7793.1997.497bj.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Xia XM, Fakler B, Rivard A, Wayman G, Johnson-Pais T, Keen JE, Ishii T, Hirschberg B, Bond CT, Lutsenko S, Maylie J, Adelman JP. Mechanism of calcium gating in small-conductance calcium-activated potassium channels. Nature. 1998;395:503–507. doi: 10.1038/26758. [DOI] [PubMed] [Google Scholar]
  • 69.Berridge MJ, Lipp P, Bootman MD. The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol. 2000;1:11–21. doi: 10.1038/35036035. [DOI] [PubMed] [Google Scholar]
  • 70.Sanders KM. Invited review: Mechanisms of calcium handling in smooth muscles. J Appl Physiol. 2001;91:1438–1449. doi: 10.1152/jappl.2001.91.3.1438. [DOI] [PubMed] [Google Scholar]
  • 71.Murray KJ. Cyclic AMP and mechanisms of vasodilation. Pharmacol Ther. 1990;47:329–345. doi: 10.1016/0163-7258(90)90060-f. [DOI] [PubMed] [Google Scholar]
  • 72.Oloizia B, Paul RJ. Ca2+ clearance and contractility in vascular smooth muscle: Evidence from gene-altered murine models. Journal of Molecular and Cellular Cardiology. 2008;45:347–362. doi: 10.1016/j.yjmcc.2008.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Bayguinov O, Hagen B, Bonev AD, Nelson MT, Sanders KM. Intracellular calcium events activated by ATP in murine colonic myocytes. Am J Physiol Cell Physiol. 2000;279:C126–C135. doi: 10.1152/ajpcell.2000.279.1.C126. [DOI] [PubMed] [Google Scholar]
  • 74.Gordienko DV, Bolton TB, Cannell MB. Variability in spontaneous subcellular calcium release in guinea-pig ileum smooth muscle cells. J Physiol. 1998;507(Pt 3):707–720. doi: 10.1111/j.1469-7793.1998.707bs.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Benham CD, Bolton TB. Spontaneous transient outward currents in single visceral and vascular smooth muscle cells of the rabbit. J Physiol. 1986;381:385–406. doi: 10.1113/jphysiol.1986.sp016333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Bolton TB, Imaizumi Y. Spontaneous transient outward currents in smooth muscle cells. Cell Calcium. 1996;20:141–152. doi: 10.1016/s0143-4160(96)90103-7. [DOI] [PubMed] [Google Scholar]
  • 77.ZhuGe R, Tuft RA, Fogarty KE, Bellve K, Fay FS, Walsh JV., Jr The influence of sarcoplasmic reticulum Ca2+ concentration on Ca2+ sparks and spontaneous transient outward currents in single smooth muscle cells. J Gen Physiol. 1999;113:215–228. doi: 10.1085/jgp.113.2.215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Hennig GW, Smith CB, O’Shea DM, Smith TK. Patterns of intracellular and intercellular Ca2+ waves in the longitudinal muscle layer of the murine large intestine in vitro. J Physiol. 2002;543:233–253. doi: 10.1113/jphysiol.2002.018986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Cahalan MD. Cell biology. How to STIMulate calcium channels. Science. 2010;330:43–44. doi: 10.1126/science.1196348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Inesi G, Sumbilla C, Kirtley ME. Relationships of molecular structure and function in Ca2+-transport ATPase. Physiol Rev. 1990;70:749–760. doi: 10.1152/physrev.1990.70.3.749. [DOI] [PubMed] [Google Scholar]
  • 81.Karaki H, Ozaki H, Hori M, Mitsui-Saito M, Amano K-I, Harada K-I, Miyamoto S, Nakazawa H, Won K-J, Sato K. Calcium movements, distribution, and functions in smooth muscle. Pharmacol Rev. 1997;49:157–230. [PubMed] [Google Scholar]
  • 82.Cheranov SY, Jaggar JH. Sarcoplasmic reticulum calcium load regulates rat arterial smooth muscle calcium sparks and transient K(Ca) currents. J Physiol. 2002;544:71–84. doi: 10.1113/jphysiol.2002.025197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Kim HW, Steenaart NA, Ferguson DG, Kranias EG. Functional reconstitution of the cardiac sarcoplasmic reticulum Ca2+-ATPase with phospholamban in phospholipid vesicles. J Biol Chem. 1990;265:1702–1709. [PubMed] [Google Scholar]
  • 84.Schmidt AG, Edes I, Kranias EG. Phospholamban: a promising therapeutic target in heart failure? Cardiovasc Drugs Ther. 2001;15:387–396. doi: 10.1023/a:1013381204658. [DOI] [PubMed] [Google Scholar]
  • 85.Ferrington DA, Yao Q, Squier TC, Bigelow DJ. Comparable levels of Ca-ATPase inhibition by phospholamban in slow-twitch skeletal and cardiac sarcoplasmic reticulum. Biochemistry. 2002;41:13289–13296. doi: 10.1021/bi026407t. [DOI] [PubMed] [Google Scholar]
  • 86.Nobe K, Sutliff RL, Kranias EG, Paul RJ. Phospholamban regulation of bladder contractility: evidence from gene-altered models. J Physiol. 2001;535.3:867–878. doi: 10.1111/j.1469-7793.2001.00867.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Wellman GC, Santana LF, Bonev AD, Nelson MT. Role of phospholamban in the modulation of arterial Ca2+ sparks and Ca2+-activated K+ channels by cAMP. Am J Physiol. 2001;281:C1029–C1037. doi: 10.1152/ajpcell.2001.281.3.C1029. [DOI] [PubMed] [Google Scholar]
  • 88.Perez GJ, Bonev AD, Nelson MT. Micromolar Ca2+ from sparks activates Ca2+-sensitive K+ channels in rat cerebral artery smooth muscle. Am J Physiol Cell Physiol. 2001;281:C1769–C1775. doi: 10.1152/ajpcell.2001.281.6.C1769. [DOI] [PubMed] [Google Scholar]
  • 89.Canon WB, Lieb CW. The receptive relaxation of the stomach. Am J Physiol. 1911;29:267–273. [Google Scholar]
  • 90.Tack J, Bisschops R, Sarnelli G. Pathophysiology and treatment of functional dyspepsia. Gastroenterology. 2004;127:1239–1255. doi: 10.1053/j.gastro.2004.05.030. [DOI] [PubMed] [Google Scholar]
  • 91.Parkman HP, Jones MP. Tests of Gastric Neuromuscular Function. Gastroenterology. 2009;136:1526–1543. doi: 10.1053/j.gastro.2009.02.039. [DOI] [PubMed] [Google Scholar]
  • 92.Petkov GV, Boev KK. The role of a ryanodine-sensitive Ca2+-store in the regulation of smooth muscle tone of the cat gastric fundus. Gen Physiol Biophys. 1998;17:225–237. [PubMed] [Google Scholar]
  • 93.Selemidis S, Cocks TM. Nitrergic relaxation of the mouse gastric fundus is mediated by cyclic GMP-dependent and ryanodine-sensitive mechanisms. Br J Pharmacol. 2000;129:1315–1322. doi: 10.1038/sj.bjp.0703174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Tokutomi Y, Tokutomi N, Nishi K. The properties of ryanodine-sensitive Ca2+ release in mouse gastric smooth muscle cells. Br J Pharmacol. 2001;133:125–137. doi: 10.1038/sj.bjp.0704048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Santana LF, Kranias EG, Lederer WJ. Calcium sparks and excitation-contraction coupling in phospholamban-deficient mouse ventricular myocytes. J Physiol. 1997;503(Pt 1):21–29. doi: 10.1111/j.1469-7793.1997.021bi.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Kim M, Cho SY, Han IS, Koh SD, Perrino BA. CaM kinase II and phospholamban contribute to caffeine-induced relaxation of murine gastric fundus smooth muscle. Am J Physiol Cell Physiol. 2005;288:C1202–C1210. doi: 10.1152/ajpcell.00299.2004. [DOI] [PubMed] [Google Scholar]
  • 97.Janiak R, Wilson SM, Montague S, Hume JR. Heterogeneity of calcium stores and elementary release events in canine pulmonary arterial smooth muscle cells. Am J Physiol Cell Physiol. 2001;280:C22–C33. doi: 10.1152/ajpcell.2001.280.1.C22. [DOI] [PubMed] [Google Scholar]
  • 98.van Breemen C, Chen Q, Laher I. Superficial buffer barrier function of smooth muscle sarcoplasmic reticulum. Trends Pharmacol Sci. 1995;16:98–105. doi: 10.1016/s0165-6147(00)88990-7. [DOI] [PubMed] [Google Scholar]
  • 99.Wellman GC, Nelson MT. Signaling between SR and plasmalemma in smooth muscle: sparks and the activation of Ca2+-sensitive ion channels. Cell Calcium. 2003;34:211–229. doi: 10.1016/s0143-4160(03)00124-6. [DOI] [PubMed] [Google Scholar]
  • 100.Ji G, Barsotti RJ, Feldman ME, Kotlikoff MI. Stretch-induced calcium release in smooth muscle. J Gen Physiol. 2002;119:533–544. doi: 10.1085/jgp.20028514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Yu YC, Guo HS, Li Y, Piao L, Li L, Li ZL, Xu WX. Role of calcium mobilization in sodium nitroprusside-induced increase of calcium-activated potassium currents in gastric antral circular myocytes of guinea pig. Acta Pharmacol Sin. 2003;24:819–825. [PubMed] [Google Scholar]
  • 102.Cohen RA, Weisbrod RM, Gericke M, Yaghoubi M, Bierl C, Bolotina VM. Mechanism of nitric oxide-induced vasodilatation: refilling of intracellular stores by sarcoplasmic reticulum Ca2+ ATPase and inhibition of store-operated Ca2+ influx. Circ Res. 1999;84:210–219. doi: 10.1161/01.res.84.2.210. [DOI] [PubMed] [Google Scholar]
  • 103.Petkov GV, Boev K. The role of sarcoplasmic reticulum and sarcoplasmic reticulum Ca2+-ATPase in the smooth muscle tone of the cat gastric fundus. Pflügers Archive. 1996;431:928–935. doi: 10.1007/s004240050087. [DOI] [PubMed] [Google Scholar]
  • 104.Raymond GL, Wendt IR. Force and intracellular Ca2+ during cyclic nucleotide-mediated relaxation of rat anococcygeus muscle and the effects of cyclopiazonic acid. Br J Pharmacol. 1996;119:1029–1037. doi: 10.1111/j.1476-5381.1996.tb15774.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Wayman CP, McFadzean I, Gibson A, Tucker JF. Inhibition by sodium nitroprusside of a calcium store depletion-activated non-selective cation current in smooth muscle cells of the mouse anococcygeus. Br J Pharmacol. 1996;118:2001–2008. doi: 10.1111/j.1476-5381.1996.tb15636.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Kuschel M, Karczewski P, Hempel P, Schlegel W-P, Krause EG, Bartel S. Ser16 prevails over Thr17 phospholamban phosphorylation in the β-adrenergic regulation of cardiac relaxation. Am J Physiol. 1999;276:H1625–H1633. doi: 10.1152/ajpheart.1999.276.5.H1625. [DOI] [PubMed] [Google Scholar]
  • 107.Nelson MT, Cheng H, Rubart M, Santana LF, Bonev AD, Knot HJ, Lederer WJ. Relaxation of arterial smooth muscle by calcium sparks. Science. 1995;270:633–637. doi: 10.1126/science.270.5236.633. [DOI] [PubMed] [Google Scholar]
  • 108.Kim M, Han IS, Koh SD, Perrino BA. Roles of CaM kinase II and phospholamban in SNP-induced relaxation of murine gastric fundus smooth muscle. Am J Physiol Cell Physiol. 2006;291:C337–C447. doi: 10.1152/ajpcell.00397.2005. [DOI] [PubMed] [Google Scholar]
  • 109.Kim M, Perrino BA. CaM kinase II activation and phospholamban phosphorylation by SNP in murine gastric antrum smooth muscles. Am J Physiol Gastrointest Liver Physiol. 2007;292:G1045–G1054. doi: 10.1152/ajpgi.00203.2006. [DOI] [PubMed] [Google Scholar]
  • 110.Robertson BE, Schubert R, Hescheler J, Nelson MT. cGMP-dependent protein kinase activates Ca-activated K channels in cerebral artery smooth muscle cells. Am J Physiol Cell Physiol. 1993;265:C299–C303. doi: 10.1152/ajpcell.1993.265.1.C299. [DOI] [PubMed] [Google Scholar]
  • 111.Schubert R, Nelson MT. Protein kinases: tuners of the BKCa channel in smooth muscle. Trends in Pharmacological Sciences. 2001;22:505–512. doi: 10.1016/s0165-6147(00)01775-2. [DOI] [PubMed] [Google Scholar]
  • 112.Ardizzone S, Bianchi PG. Biologic therapy for inflammatory bowel disease. Drugs. 2005;65:2253–2286. doi: 10.2165/00003495-200565160-00002. [DOI] [PubMed] [Google Scholar]
  • 113.Rufo PA, Bousvaros A. Current therapy of inflammatory bowel disease in children. Paediatr Drugs. 2006;8:279–302. doi: 10.2165/00148581-200608050-00002. [DOI] [PubMed] [Google Scholar]
  • 114.Sato K, Ohkura S, Kitahara Y, Ohama T, Hori M, Sato M, Kobayashi S, Sasaki Y, Hayashi T, Nasu T, Ozaki H. Involvement of CPI-17 downregulation in the dysmotility of the colon from dextran sodium sulphate-induced experimental colitis in a mouse model. Neurogastroenterol Motil. 2007;19:504–514. doi: 10.1111/j.1365-2982.2007.00911.x. [DOI] [PubMed] [Google Scholar]
  • 115.Ajaj WM, Lauenstein TC, Pelster G, Gerken G, Ruehm SG, Debatin JF, Goehde SC. Magnetic resonance colonography for the detection of inflammatory diseases of the large bowel: quantifying the inflammatory activity. Gut. 2005;54:257–263. doi: 10.1136/gut.2003.037085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Myers BS, Martin JS, Dempsey DT, Parkman HP, Thomas RM, Ryan JP. Acute experimental colitis decreases colonic circular smooth muscle contractility in rats. Am J Physiol. 1997;273:G928–G936. doi: 10.1152/ajpgi.1997.273.4.G928. [DOI] [PubMed] [Google Scholar]
  • 117.Pazdrak K, Shi XZ, Sarna SK. TNFα suppresses human colonic circular smooth muscle cell contractility by SP1- and NF-κB-mediated induction of ICAM-1. Gastroenterology. 2004;127:1096–1109. doi: 10.1053/j.gastro.2004.07.008. [DOI] [PubMed] [Google Scholar]
  • 118.Snape WJ, Jr, Williams R, Hyman PE. Defect in colonic smooth muscle contraction in patients with ulcerative colitis. Am J Physiol. 1991;261:G987–G991. doi: 10.1152/ajpgi.1991.261.6.G987. [DOI] [PubMed] [Google Scholar]
  • 119.Kinoshita K, Sato K, Hori M, Ozaki H, Karaki H. Decrease in activity of smooth muscle L-type Ca2+ channels and its reversal by NF-kappaB inhibitors in Crohn’s colitis model. Am J Physiol Gastrointest Liver Physiol. 2003;285:G483–G493. doi: 10.1152/ajpgi.00038.2003. [DOI] [PubMed] [Google Scholar]
  • 120.Liu X, Rusch NJ, Striessnig J, Sarna SK. Down-regulation of L-type calcium channels in inflamed circular smooth muscle cells of the canine colon. Gastroenterology. 2001;120:480–489. doi: 10.1053/gast.2001.21167. [DOI] [PubMed] [Google Scholar]
  • 121.Al-Jarallah A, Oriowo MA, Khan I. Mechanism of reduced colonic contractility in experimental colitis: role of sarcoplasmic reticulum pump isoform-2. Mol Cell Biochem. 2007;298:169–178. doi: 10.1007/s11010-006-9363-8. [DOI] [PubMed] [Google Scholar]
  • 122.Calao M, Burny A, Quivy V, Dekoninck A, Van Lint C. A pervasive role of histone acetyltransferases and deacetylases in an NF-kappaB-signaling code. Trends in Biochemical Sciences. 2008;33:339–349. doi: 10.1016/j.tibs.2008.04.015. [DOI] [PubMed] [Google Scholar]
  • 123.McKinsey TA. Derepression of pathological cardiac genes by members of the CaM kinase superfamily. Cardiovascular Research. 2007;73:667–677. doi: 10.1016/j.cardiores.2006.11.036. [DOI] [PubMed] [Google Scholar]
  • 124.Stull JT, Tansey MG, Tang DC, Word RA, Kamm KE. Phosphorylation of myosin light chain kinase: a cellular mechanism for Ca2+ desensitization. Mol Cell Biochem. 1993;127–128:229–237. doi: 10.1007/BF01076774. [DOI] [PubMed] [Google Scholar]

RESOURCES