干细胞之家 - 中国干细胞行业门户第一站

 

 

搜索
朗日生物

免疫细胞治疗专区

欢迎关注干细胞微信公众号

  
查看: 384251|回复: 262
go

CaMKII tethers to L-type Ca2 channels, establishing a local and dedic [复制链接]

Rank: 7Rank: 7Rank: 7

积分
威望
0  
包包
3465  
楼主
发表于 2009-3-6 08:52 |只看该作者 |倒序浏览 |打印
1 Department of Neurobiology, Stanford University, Stanford, CA 943052 ~! n; P  C$ r( @

' g+ o( ]+ l8 L2 Department of Molecular and Cellular Physiology, Stanford University, Stanford, CA 94305! Y/ n, a, Q5 O( w6 _* E
3 q3 N4 ], b- ?& M! R( b% C6 [4 g. _
3 Department of Pharmacology, College of Physicians and Surgeons of Columbia University, New York, NY 10032
" K: E1 i6 A, i% Z: }  a) [" \, F  f' A% x
4 Department of Medicine, College of Physicians and Surgeons of Columbia University, New York, NY 10032
4 Q8 H# ~, _# q- |; O+ ?  w
; p) L0 w" \9 n2 u- a5 Center for Molecular Cardiology, College of Physicians and Surgeons of Columbia University, New York, NY 10032
+ T5 t% i7 U/ e/ Y' U1 c6 k
0 G8 J* V; T  x& ?! Z  gCorrespondence to Richard W. Tsien: rwtsien@stanford.edu$ K* o7 h9 D: B2 ]# e5 ^' T* j# G
/ u9 D  e+ [6 Z  c4 u2 X
Ca2 -dependent facilitation (CDF) of voltage-gated calcium current is a powerful mechanism for up-regulation of Ca2  influx during repeated membrane depolarization. CDF of L-type Ca2  channels (Cav1.2) contributes to the positive force–frequency effect in the heart and is believed to involve the activation of Ca2 /calmodulin-dependent kinase II (CaMKII). How CaMKII is activated and what its substrates are have not yet been determined. We show that the pore-forming subunit 1C (Cav1.2) is a CaMKII substrate and that CaMKII interaction with the COOH terminus of 1C is essential for CDF of L-type channels. Ca2  influx triggers distinct features of CaMKII targeting and activity. After Ca2 -induced targeting to 1C, CaMKII becomes tightly tethered to the channel, even after calcium returns to normal levels. In contrast, activity of the tethered CaMKII remains fully Ca2 /CaM dependent, explaining its ability to operate as a calcium spike frequency detector. These findings clarify the molecular basis of CDF and demonstrate a novel enzymatic mechanism by which ion channel gating can be modulated by activity.
; Q1 N9 t6 k4 c3 O; C# y4 Z( P- ?6 a6 L! F3 D( e) j) ^* I
A. Hudmon's present address is Dept. of Neurology, Yale University, New Haven, CT 06516.
& [& H$ c; l; X/ b1 F) y% l$ R
& {1 [* q# p+ V6 U' }8 P0 mH. Schulman's present address is SurroMed, Inc., Menlo Park, CA 94025.
& x3 g' I" B0 a1 M9 b, l* j1 f2 x+ Z  `7 N- |# ]
Abbreviations used in this paper: AIP-2, autocamtide-2–related inhibitory peptide; AKAP, A-kinase anchor protein; CaMKII, Ca2 /calmodulin-dependent kinase II; CDF, Ca2 -dependent facilitation; CDI, Ca2 -dependent inactivation; HEK, human embryonic kidney; NMDAR, N-methyl-D-aspartate receptor; PKA, protein kinase A; Po, open probability; PP1, protein phosphatase 1.# K, T% `2 T/ E' s
. r- _3 m9 {/ x5 r
Introduction0 G. B; P$ }# ^# W8 H/ i2 T
6 a% v; c$ ~- x' s( L, ]' m: I8 k+ z
Ca2 -dependent facilitation (CDF) of calcium channels serves to potentiate the Ca2  influx through the L-type channels during repeated activity. CDF is a feed-forward form of adaptive plasticity that is a critical regulatory feature of many excitable cells. In the heart, frequency-dependent potentiation of Ca2  current through L-type channels (Cav1.2; Noble and Shimoni, 1981; Marban and Tsien, 1982; Lee, 1987; Schouten and Morad, 1989; Zygmunt and Maylie, 1990) contributes to the force–frequency relationship of cardiac contraction (Koch-Weser and Blinks, 1963). This increased contraction strength with faster heart rates contributes to the positive inotropic response during exercise (Ross et al., 1995) and is abnormal in heart failure (Feldman et al., 1988; Mulieri et al., 1992; Hasenfuss et al., 1994). In the brain, CDF of L-type channels may be important in relation to the privileged role of L-type channels in excitation–transcription coupling (Deisseroth et al., 2003). Despite these important physiological roles that are central to cardiac function and neuronal plasticity, there is little understanding of the molecular mechanism of CDF of L-type channels.
) `, w1 |+ R2 t0 B6 G, Y  k& Y' f& O1 G
Ca2 /calmodulin-dependent protein kinase II (CaMKII), a multifunctional Ser/Thr protein kinase, is a likely effector of CDF. Pharmacological inhibition of CaMKII abolishes CDF in the heart (Xiao et al., 1994; Yuan and Bers, 1994). Addition of activated CaMKII to the cytoplasmic face of cardiac myocyte membranes induces a high open-probability state of the channel that is consistent with the properties of Ca2  channels displaying CDF (Dzhura et al., 2000). Further, immunocytochemical data suggest that the Cav1.2 and CaMKII are localized close to each other on the cardiomyocyte sarcolemmal membrane (Xiao et al., 1994), suggesting that the kinase has easy access to the channel., A) R! S7 O. d. M5 ]' D9 |

5 K5 p6 V( e; H* @% p$ GCaMKII has structural and functional properties that make it an ideal candidate to sense the frequency of Ca2  transients during neuronal firing or changes in cardiac rhythm and translate that frequency signal into activity-dependent alterations such as CDF. CaMKII is a multimeric holoenzyme composed of 12 subunits, with the subunit isoforms being derived from a family of four closely related genes (, ?, , and ; Hudmon and Schulman, 2002b). In the brain, -CaMKII has been shown to play a key role in synaptic plasticity and learning/memory (Lisman et al., 2002). The  and  isoforms predominate in the heart and have been implicated in the regulation of gene expression as well as CDF (Zhang et al., 2002). In all of these isoforms, activation proceeds by Ca2 /CaM binding to an autoregulatory region, which causes the removal of a pseudosubstrate domain from the catalytic site. After the initial stimulus, autophosphorylation of Thr286 or its equivalent (Thr287 in non– isoforms) renders subsequent kinase activity independent (autonomous) of Ca2  and CaM (Miller et al., 1988) and increases the kinase's affinity for CaM by >10,000-fold ("CaM trapping"; Meyer et al., 1992). These properties endow CaMKII with the ability to become persistently activated in a transition that is sharply dependent on the frequency of Ca2  oscillations (De Koninck and Schulman, 1998; Eshete and Fields, 2001; Bayer et al., 2002; Bradshaw et al., 2003).
: Q+ B7 @( X- P# x$ u2 ]% ^
/ \3 g! i7 u# _  xWe now demonstrate for the first time that the subcellular localization of CaMKII is critical for its biological role as a frequency decoder of voltage-driven calcium spikes. We show that CaMKII phosphorylates 1C and that tethering of CaMKII to the 1C COOH terminus is an essential molecular feature of CDF. We present a molecular model for CDF in which a dedicated CaMKII holoenzyme acts as both a local sensor to monitor Ca2  channel activity and as a resident kinase effector to regulate Ca2  channel activity.
. c  C) u. p5 Y/ q! I, b3 {( `- }$ o: A
Results9 }0 j/ z0 x7 B9 c( k) K# ~

7 f: D4 \+ b' ^, w9 q7 ]The NH2 and COOH termini of 1C are substrates of CaMKII" S2 J, S; s- [( Z3 w* s7 j" K
6 L8 a$ y7 t. T; j; x" w; z) L. U4 z
Modulation of L-type channel gating by cytoplasmic delivery of constitutively active CaMKII is blocked by nonhydrolyzable analogues of ATP (Dzhura et al., 2000), suggesting that the kinase acts through phosphorylation of the channel or an associated regulatory protein. Because the kinase-induced increase in L-type Ca2  current by both protein kinase A (PKA) and Src results from phosphorylation of 1C (De Jongh et al., 1996; Bence-Hanulec et al., 2000), we first tested whether 1C was also a substrate for activated CaMKII (Fig. 1 A). The addition of activated CaMKII to 1C immunoprecipitated from lysates of L-type channel–expressing human embryonic kidney (HEK) cells resulted in the phosphorylation of protein migrating at 240 kD, consistent with the molecular mass of 1C. The kinase activity could be attributed to CaMKII and not to another kinase coimmunoprecipitated with 1C because inclusion of the CaMKII inhibitor autocamtide-2–related inhibitory peptide (AIP-2) prevented phosphorylation; continued presence of the 1C protein under this condition was confirmed by immunoblotting (Fig. 1 A, bottom). The immunoprecipitated and phosphorylated protein could be confidently identified as 1C in light of the findings that no 1C was immunoprecipitated and that 32P was not incorporated when immunoprecipitation was performed with control IgG or with lysates of HEK cells in which 1C had not been expressed. Interestingly, under conditions in which 1C was phosphorylated by CaMKII (Fig. 1 A, lane 3), we noticed a 32P-labeled protein (50 kD) corresponding to the autophosphorylated form of the  subunit of CaMKII that had been introduced for the kinase assay. Immunoblots with an anti-CaMKII antibody confirmed its identity (not depicted, but see Fig. 5 E). The retention of CaMKII, despite extensive washing of the immobilized 1C, suggested that 1C may serve as an anchoring protein as well as a substrate for the kinase. The near absence of retention when AIP-2 was added to the reaction gave an early indication about the mechanism of anchoring (see Fig. 6 B).9 G+ ~' W* Q. P1 A6 e  K# s, ?1 F
- A8 q7 @& ]7 k% N
Figure 1. Phosphorylation of the 1C subunit by CaMKII. (A) Autoradiogram showing phosphorylation of 1C by CaMKII. Lysates from HEK cells transfected with 1C and ?2b (lanes 2–4) or nontransfected cells (lane 1) were immunoprecipitated with an anti-1C antibody (lanes 1, 3, and 4) or control IgG (lane 2) and then incubated with purified -CaMKII in the presence of Ca2 /CaM and Mg2 /ATP32 as described in Materials and methods. 200 nM of the CaMKII inhibitor peptide AIP-2 (Calbiochem) was included (lane 4) to demonstrate kinase specificity. Phosphorylated 1C is indicated by an arrowhead; autophosphorylated CaMKII, retained after the kinase reaction despite extensive washing of the immunoprecipitate, is indicated with a double arrowhead. An anti-1C immunoblot of the samples used in the kinase reaction is shown below the autoradiogram. (B) Schematic of 1C. Thick black lines highlight regions used to generate GST fusion proteins. (C) GST fusion proteins enriched from bacterial lysates using glutathione–sepharose were incubated with purified -CaMKII in the presence of Ca2 /CaM and Mg2 /ATP32 as described in Materials and methods. After extensive washes, proteins were eluted using SDS-PAGE sample buffer. Autoradiogram of fusion proteins separated by SDS-PAGE after phosphorylation by CaMKII. C-term refers to the more distal COOH-terminal fusion protein, containing aa 1669–2171. Above the autoradiogram is the Coomassie blue–stained band for each fusion protein, indicating nearly equal loading of substrate for all fusion proteins.
+ I; Q& x  z2 G5 e  E; I, a! R9 ]# [( f! j& T/ o' C5 L& S
Having demonstrated that 1C was a CaMKII substrate, we next ran tests to determine which of the intracellular domains of 1C were phosphorylated by CaMKII. GST fusion proteins were generated for the entire sequence of each of the intracellular domains of the 1C subunit except the large cytoplasmic tail, which was represented by two complementary fragments (aa 1507–1622 and 1669–2171; Fig. 1 B). When the fusion proteins were tested in an in vitro kinase assay, significant incorporation of 32P was only observed for the NH2-terminal construct and the COOH-terminal fusion protein containing aa 1669–2171 (Fig. 1 C) and not the fusion protein containing aa 1507–1622 (not depicted). The finding that CaMKII can phosphorylate NH2- and COOH-terminal regions of 1C is provocative in light of previous data suggesting that these regions may be targets of kinase action for modulation of Cav1.2 function (Rotman et al., 1995; Bence-Hanulec et al., 2000; McHugh et al., 2000). Similar to results with the intact channel (Fig. 1 A), we again noticed in multiple lanes an 50-kD 32P-labeled protein corresponding to the autophosphorylated form of the  subunit of CaMKII that had been introduced for the kinase assay. The finding that -CaMKII could be retained by individual domains of 1C suggested that these domains might contribute to the kinase anchoring to the channel subunit as a whole.8 s" W- r/ l+ M+ H" q
. d4 p* R# J. l9 _3 Q) f
CaMKII interacts specifically with 1C
) y8 ^! ^8 {5 d& \# I$ {, c* B6 [+ S  ~, i5 t3 g
We tested the possibility that CaMKII tethers to 1C in the rat heart by attempting to coimmunoprecipitate CaMKII with 1C (Fig. 2 A). An anti-1C antibody (but not control IgG) coimmunoprecipitated an 58-kD protein from a rat heart that was easily detectable with a biotinylated calmodulin overlay, which was consistent with the properties of -CaMKII. Tethering of the kinase to the pore-forming subunit was further evaluated in experiments with HEK 293 cells coexpressing GFP-tagged CaMKII and Xpress-tagged 1C, along with the calcium channel accessory subunits 2 and ?2 (Fig. 2 B). We observed coimmunoprecipitation of the GFP-CaMKII by the antibody to epitope-tagged 1C (Fig. 2 B, lane 5), but not by a control IgG (lane 4).5 I( f2 C. M, s4 I! j
3 J9 K+ m2 [( p* }1 Z# \3 M  @
Figure 2. CaMKII coimmunoprecipitates and colocalizes with 1C. (A) Biotinylated calmodulin overlay of rat cardiac sarcolemmal membranes after immunoprecipitation with an anti-1C antibody. Purified -CaMKII was run as a control to demonstrate effectiveness of CaM overlay. An anti-1C antibody (but not control IgG) coimmunoprecipitated a protein identified as the  isoform of CaMKII by biotinylated CaM overlay and apparent molecular mass. (B) Anti-GFP immunoblot after immunoprecipitation of GFP-CaMKII by control IgG (lane 4) or anti-1C antibody (lane 5) from lysates of HEK 293 cells transiently transfected with GFP-CaMKII and 1C.
. H1 e& v' H: A0 n: N9 R2 j( Y- C$ X  r8 M. F. s" K
Activity-dependent association of CaMKII with multiple cytoplasmic regions of 1C3 h4 f% Q7 F( g: m4 ?+ ]

+ ^* c9 e6 L/ r) L" r* W0 S( z# w8 qTo further define the interaction between CaMKII and 1C, we constructed a pull-down binding assay using various 1C-GST fusion proteins (Fig. 3). The goal was to find out whether a direct interaction could be observed in vitro and whether different activation states of CaMKII modulated binding. When CaMKII was activated with Ca2 /CaM but not allowed to undergo autophosphorylation (ATP not included), the kinase bound to the NH2-terminal domain and the III-IV loop of 1C (Fig. 3, middle). Subsequent removal of Ca2 /CaM in the wash buffer reversed this binding (unpublished data). When CaMKII was activated in the presence of Ca2 /CaM plus ATP on ice, conditions previously shown to produce predominantly Thr286 autophosphorylation (Lai et al., 1987; Lou and Schulman, 1989; Ikeda et al., 1991), CaMKII again bound to the NH2 terminus and III-IV loop, but additionally bound to the I-II loop and the COOH terminus (Fig. 3, bottom). In contrast, CaMKII did not bind to any of the cytoplasmic region–containing GST fusion proteins in the absence of activating stimuli (Fig. 3, top). We concluded that the initiation of a direct interaction between CaMKII and 1C requires activation of the kinase by Ca2 /CaM. A subsequent activation state, that produced by autophosphorylation, was necessary for binding to additional cytoplasmic regions of 1C.
( y# O, g+ [0 s' A' W
$ h# `% A. X7 Q3 jFigure 3. Activity-dependent interaction of CaMKII with the cytoplasmic determinants of 1C. Immunoblots using an mAb (CB2) for CaMKII after a GST pull-down assay with 20 nM of native (top), Ca2 /CaM-activated (middle), or Ca2 /CaM/autophosphorylated -CaMKII (bottom). GST fusion proteins contained various cytoplasmic regions of 1C just as in Fig. 1 C. Panel above the immunoblots shows a representative Ponceau stain of each fusion protein. Although only one Ponceau staining profile is shown, all blots were run in parallel, and equal loading of all fusions proteins was independently verified.5 W. H. c. |' g' C; _
4 ]( z4 `, y: }; O$ w
To identify novel structural determinants of 1C that functionally affect CDF, we initially focused on the COOH terminus. This region displays an appropriate combination of attributes for CaMKII-mediated CDF: it is a target for phosphorylation by CaMKII (Fig. 1 C); it binds preferentially to autophosphorylated CaMKII (Fig. 3), a state of the kinase capable of supporting facilitation of single channels (Dzhura et al., 2000); and it has been implicated in Ca2 -dependent modulation of channel function (Hell et al., 1995; Zühlke et al., 1999; Gao et al., 2001). To delimit the locus of CaMKII binding within the COOH-terminal tail of 1C, we used a series of GST fusion proteins corresponding to different portions of this region (Fig. 4 A). We found a pattern of interactions with autophosphorylated CaMKII that suggested that the kinase bound between residues 1622 and 1669 of 1C. Because a weak interaction was also seen with a construct proximal to 1622, we generated a fusion protein spanning aa 1581–1690 for additional testing and found a clear interaction (Fig. 4 A). To further narrow down the CaMKII interaction site within this 110-aa region, we probed its interaction with autophosphorylated CaMKII and assessed interference by a series of 22 overlapping aa peptides (Fig. 4 B; Pitt et al., 2001). A peptide generated from residues 1639–1660 dramatically reduced the interaction of the kinase with the 1581–1690 fusion protein. In contrast, the CaMKII interaction was not inhibited by two peptides (1589–1610 and 1615–1636) that corresponded to sites important for tethering of apoCaM (Pitt et al., 2001; Kim et al., 2004). One stretch of six residues within the 1639–1660 peptide, TVGKF(Y/I)A, was identified as being nearly identical in 1C and 1A (Cav2.1), the pore-forming subunit of P/Q-type Ca2  channels, which display their own form of CDF (Lee et al., 1999; DeMaria et al., 2001). Accordingly, we constructed an 1C fusion protein containing the amino acids EEDAAA in place of TVGKFY within an otherwise wild-type sequence of residues 1581–1690 (Mut6). CaMKII binding to the Mut6 fusion protein was reduced by 87.3 ± 4.5% relative to binding to wild-type 1581–1690 fusion protein (Fig. 4 C). In contrast, the same amino acid substitution left CaM binding to this mutant fusion protein unaffected (Fig. 4 D).
) \* |% d+ D7 U) v8 C1 k
  |9 t, L+ f& C1 a/ D$ IFigure 4. Localization of the CaMKII binding site on the COOH terminus of 1C. (A) Diagram of 1C fusion proteins used in GST pull-down assays with autophosphorylated -CaMKII, exhibiting robust ( ), partial (±), and no (–) binding. (B) Immunoblot with CB2 after GST pull down of 20 nM of purified autophosphorylated -CaMKII, using 1C aa 1581–1690 fused to GST. Pull-down assay performed in the presence of 40 μM of the indicated peptide or the peptide diluent DMSO. (C) Quantification after immunoblot with CB2 of GST pull-down assays of purified autophosphorylated -CaMKII, using 1C aa 1581–1690 (wild type ), a 1644TVGKFY1649  EEDAAA mutant (Mut6), or GST alone shows that Mut6 blocks CaMKII binding. Panel above the immunoblots shows a representative Ponceau stain of each fusion protein. *, P 2 F9 x! X4 {: x5 i+ o5 M
  a& g$ w$ ]% _4 _+ w
Disruption of CaMKII binding to the COOH terminus of 1C prevents CDF
8 A: s  i, J3 n3 U: p7 U5 {0 s: y; E" s5 A
We then tested whether this site was critical for CDF by introducing the Mut6 mutation into 1C subunits of L-type channels expressed in Xenopus laevis oocytes. Because L-type channels also display a strong Ca2 -dependent inactivation (CDI) process that could diminish our ability to detect facilitation, we sought conditions under which CDF could be observed without the counteraction of CDI. Fortunately, robust CDF during trains of depolarizing pulses can be obtained by means of a point mutation within the IQ motif (I1654A; Zühlke et al., 1999, 2000) that eliminates CDI. In this setting, the Mut6 modification of the CaMKII interaction site completely abolished CDF (Fig. 5, A–C). There was no potentiation of ICa at any point during the train of 40 successive depolarizations within the entire range of frequencies tested (0.5–3.3 Hz). Abolition of the Ca2 -dependent facilitatory process was also observed in experiments using a two-pulse protocol and finely graded changes in interpulse interval. Ca2  currents evoked by the second pulse averaged 110% of those elicited by the first pulse at a time interval when the peak Ba2  current had only recovered to 95% (Fig. 5 D). A comparable difference between recovery of Ca2  and Ba2  currents was seen in wild-type 1C (Zühlke et al., 1999, 2000) but was likewise abolished by the Mut6 modification (unpublished data). Thus, in both of the approaches used to assess facilitation〞potentiation of ICa during trains of depolarizations and recovery from the aftereffects of a single pulse〞CDF was abolished by mutation of the COOH-terminal CaMKII interaction site on Cav1.2.
6 u* `$ y4 M5 l, ~) L3 C7 Q9 ~8 g7 a
Figure 5. CaMKII interaction with the COOH terminus of 1C is essential for CDF. (A) IBa and scaled ICa traces during a train of 40 test pulses of Vh from –90 mV to  20 mV at 3.3 Hz recorded from oocytes expressing 1C I1654A (I/A) or 1C I1654A/1644TVGKFY1649  EEDAAA (I/A-Mut6). Bars, 500 nA and 25 ms. (B) Peak IBa and ICa during trains of 40 repetitive test pulses at 3.3 Hz, normalized to the current amplitude at the beginning of each train (n = 4–5). Values indicate means ± SEM. (C) Changes in peak IBa and ICa conducted by 1C I1654A (I/A) or 1C I1654A/1644TVGKFY1649  EEDAAA (I/A-Mut6) at indicated stimulation frequencies (n = 4–5) Values indicate means ± SEM. (D) Summary of the recovery from inactivation after a two-step protocol for I/A and I/A-Mut6. The length of the prepulse was individually determined for each oocyte to produce 75–90% inactivation. (E) Autoradiograph showing phosphorylation of wild type (WT) or mutant 1C (Mut6) by CaMKII, performed as in Fig. 1 A. An anti-1C immunoblot of the samples used in the kinase reaction confirmed similar expression levels of the WT and mutant 1C subunits. An anti-CaMKII immunoblot with CB2 confirmed the identity of the retained 50-kD 32P-labeled protein as -CaMKII.: Z* x/ U) z1 X% n4 G% k. {

$ U5 B& }+ R7 k: v5 Y! JTo examine the mechanism by which the mutation prevented CDF, we tested whether the Mut6 channel was still a substrate for CaMKII using an in vitro assay like that in Fig. 1, in which the availability of kinase for phosphorylation was not limited and not dependent on tethering to the COOH terminus. Disruption of the CaMKII binding site on the COOH terminus by the Mut6 substitution did not reduce 32P incorporation into 1C (Fig. 5 E), suggesting that the 1C retained its ability to undergo phosphorylation. Together, these data support the hypothesis that CDF depends on tethering of CaMKII to this COOH-terminal site. Like the wild-type 1C, the Mut6 1C displayed an interaction with autophosphorylated (32P-labeled) CaMKII. The retention of kinase binding was not surprising in light of the multiple sites on 1C for CaMKII interaction that we had previously identified (Fig. 3).
  z7 l& v/ g6 U; D* e6 u8 K; {9 x% r- c8 P+ p0 e. E$ v/ z0 z+ G
The CaMKII binding site for the COOH terminus of 1C is conserved among multiple CaMKII isoforms and localizes to the catalytic domain0 T2 f. D& ^# J
9 s, }/ p; k. F: U
Although we had used -CaMKII, the predominantly brain-enriched isoform studied in the preceding in vitro experiments, there are several other CaMKII isoforms that differ in their cellular and subcellular distributions (Hudmon and Schulman, 2002a). The  isoform, the major CaMKII isoform in the heart (Edman and Schulman, 1994), was of particular interest (Fig. 2 A). Accordingly, we examined the generality of CaMKII interactions with the COOH-terminal tail of 1C across a range of isoforms. The , ?, B, A, and C isoforms were transiently expressed in HEK 293 cells for use as source material in pull-down assays and detected by the sensitive calmodulin overlay technique (Glenney and Weber, 1983; Fig. 6 A). In the absence of autophosphorylation, no binding was ever observed for any of the isoforms tested (unpublished data). However, once autothiophosphorylated, robust binding to the 1C COOH-terminal tail was observed for each of these CaMKII isoforms, with the sole exception of B-CaMKII. Thus, the capability of interaction with Cav1.2 is a widespread property of the CaMKII family, including the /? and  isoforms prevalent in brain and cardiac tissue.
% G1 n7 }/ G3 a' ^& M1 i# C) J$ [
2 c6 [4 W) @+ X" m4 PFigure 6. The binding site for the COOH terminus of 1C on CaMKII is localized near the catalytic domain. (A) Biotinylated CaM overlay of GST pull downs, using a fusion protein from the COOH terminus of 1C (aa 1509–1905) on lysates of HEK 293 cells transiently transfected with the CaMKII isoforms (, ?, A, C, and B; arrows) after thioautophosphorylation. In lanes 6 and 7, lysates of untransfected cells were run with ( ) and without (–) purified thiophosphorylated -CaMKII added to the lysate. (B) Immunoblot using an mAb (CB2) for CaMKII after GST pull downs, using a fusion protein from the COOH terminus of 1c (aa 1509–1905) and 20 nM of purified autophosphorylated -CaMKII. In addition, 20 μM of the indicated peptide was added to each binding reaction. (C) Sequence alignment of CaMKII binding sites from the COOH termini of NR2B and 1C with the autoregulatory domain from -CaMKII.3 r# [+ H1 s: `# E/ _

% T1 ~5 J+ Y7 L. N0 IWhere is the binding site for 1C on CaMKII? The conserved nature of the 1C binding site between brain and cardiac CaMKII isoforms favored a binding site that is conserved among the different kinase isoforms. We examined the conserved catalytic domain of -CaMKII, based on a recent report describing its interaction with the COOH terminus of the NR2B subunit of the neuronal N-methyl-D-aspartate receptor (NMDAR; Bayer et al., 2001). Indeed, binding of the COOH-terminal tail of 1C to autophosphorylated CaMKII was blocked by a peptide modeled after the CaMKII binding site of the NR2B subunit (NR2B peptide; Fig. 6 B). Further, binding of 1C to CaMKII was potently blocked by peptides designed around Thr286 and the autoregulatory domain of CaMKII, including the peptide substrate AC-2 and the peptide inhibitor AC-3i (Fig. 6 B), as well as AIP-2 (Fig. 1 A). As expected, the control peptide AC-3c had no effect on binding. Both sets of observations resemble previous findings using peptide inhibition to study binding of CaMKII to NR2B (Strack et al., 2000; Bayer et al., 2001). A logical conclusion is that similar or identical molecular determinants on CaMKII are responsible for binding either to 1C or to NR2B. The NR2B sequence that was found to support interaction with CaMKII closely resembles the autoregulatory domain of CaMKII surrounding Thr286 (Fig. 6 C; Bayer et al., 2001). In turn, both of these stretches of amino acids show significant resemblance to the region of 1C that we identified as critical for CaMKII interaction by peptide competition (Fig. 4 B), and that includes the TVGKFY sequence that was altered to the detriment of the 1C–CaMKII interaction. Although the corresponding regions of 1C and NR2B display points of sequence similarity (Fig. 6 C, dots and dashes), the overall degree of homology is limited.
& b8 K3 N9 k/ G" L2 }# a3 S, X( N2 t! r# k9 i3 s# Y5 G) z
CaMKII binding to the COOH terminus of 1C produces a dedicated Ca2  sensor+ z" ?5 B$ n/ t9 ]; }( B4 s

$ c- d7 L7 d4 C4 m! `, }, ~% I# QThe functional nature of the channel–kinase interaction could follow one of several possible scenarios. During recurrent rises and falls in Ca2 , the enzyme might cycle on and off the channel. Alternatively, CaMKII might remain anchored to 1C with its activity persistently switched on, like CaMKII associated with the NMDAR (Bayer et al., 2001). Finally, CaMKII might stay tethered to the 1C subunit, like PKA associated with Cav1.2 through A-kinase anchor protein (AKAP; Tavalin et al., 1999), but with kinase activity modulated by local changes in Ca2 /CaM, similar to the way that PKA is regulated by cAMP for ?-adrenergic modulation (Gao et al., 1997). To explore these possibilities, we tested whether CaMKII dissociated from the COOH-terminal tail on reversal of the Ca2  elevation or the kinase activation that initially drove the interaction.6 V& O) S, ^- }% F1 I
) t; j1 t) v3 d
When the Ca2  chelator EGTA was added immediately after the preautophosphorylation reaction, the binding of CaMKII to the 1C COOH-terminal tail was inhibited (Fig. 7 A). In contrast, once autophosphorylated CaMKII had bound to the 1C COOH-terminal tail, EGTA in the wash buffer (two or three rounds of washing, each lasting 5 min) failed to dissociate the kinase (Fig. 7 A). Dephosphorylation of autophosphorylated CaMKII with protein phosphatase 1 (PP1) before presenting the kinase to the 1C COOH-terminal fusion protein prevented binding (Fig. 7 B). However, dephosphorylation of CaMKII after binding did not. Even the combination of postbinding dephosphorylation and EGTA application failed to reverse binding (Fig. 7 B). In control experiments, immunoblotting with the phosphospecific antibody indicated that Thr286 had been completely dephosphorylated by PP1 treatment after the initial kinase binding (Fig. 7 B). Thus, although Ca2 /CaM and autophosphorylation were necessary for CaMKII to bind to the 1C COOH terminus, the same conditions were no longer required to sustain the interaction.
& ]1 f6 a( K/ c: x( Z' V$ W- h3 E3 ?; k& n* z' o- g: J  a+ `2 E
Figure 7. CaMKII interaction with the COOH terminus of 1C is not reversed by dephosphorylation or CaM dissociation, and tethered CaMKII requires autophosphorylation or Ca2 /CaM for activity. (A and B) Immunoblots with CB2 or a phosphospecific CaMKII mAb after GST pull-down assays, using 1C aa 1509–1905 and 20 nM of autophosphorylated -CaMKII. (A) 5 mM EGTA was present in the binding reaction and/or in the wash. (B) Purified recombinant PP1 was added before (PP1-Pre) or after (PP1-Post) the binding reaction in the presence or absence of 5 mM EGTA, as indicated. (C) Time course of reversal of CaMKII autonomous activity after PP1 treatment in solution (n = 4). (D) Activity measurements, using peptide AC-2 as a substrate, of CaMKII recovered in GST pull-down assays, using 1C aa 1509–1905. Ca2 /CaM-dependent and autonomous activity measurements of CaMKII recovered after treatment with recombinant PP1 for 30 min (PP1) or no treatment (–) in the binding assay (n = 4) Values indicate means ± SD.1 ~) K3 J* q( ~( |; y

" W, k. @" y6 _; `3 S$ r9 XTethered CaMKII retains its dependence on Ca2 /CaM for activity
* u2 A: o2 K/ q# p; C" X. T: G' t% o: y7 j8 A: j+ f9 k
Because the CaMKII binding for both 1C and NR2B appears to localize to the catalytic domain of the kinase, we asked whether 1C binding to CaMKII regulates its kinase activity, as in the case of NR2B. When bound to NR2B, CaMKII remains active in phosphorylating substrates even in the absence of Ca2 /CaM and autophosphorylation (Bayer et al., 2001). To determine how CaMKII is regulated when it is stably bound to the 1C COOH terminus, we examined the Ca2 /CaM-dependent and -independent (autonomous) activity after PP1 treatment. Dephosphorylation by PP1, assessed by tracking the loss of autonomous activity for soluble kinase, was complete within 30 min (Fig. 7 C). Under similar conditions, we observed that treatment of 1C-bound kinase with PP1 completely eliminated autonomous activity (remaining activity was 1.2 ± 0.6% of that without PP1 treatment; Fig. 7 D). Thus, autonomous activity of bound CaMKII was not maintained merely by interaction of the kinase with the 1C COOH terminus but depended strictly on CaMKII autophosphorylation. After PP1 treatment, tethered CaMKII could be reactivated by Ca2 /CaM. In these respects, CaMKII binding to 1C or to NR2B had very different effects on the activity of the kinase. As mentioned in Discussion, the association of CaMKII to the 1C COOH terminus is well suited to localize the kinase in close proximity to its regulatory target but not to keep the kinase constitutively active.
. t. k# M) v, Y9 g0 [; b% ?
7 T/ R8 R* V" SDiscussion0 a. C1 ^* |4 D! }7 ^& |( y/ D8 G

9 W' e8 r7 a6 s* l8 {CDF is a powerful positive feedback mechanism that allows excitable cells such as myocytes and neurons to modulate Ca2  entry through Ca2  channels according to the previous pattern of repetitive activity. The functional consequences are clearest in the heart, where CDF of L-type channels is required for sinoatrial pacemaker activity (Vinogradova et al., 2000) and contributes to the myocardial force–frequency relationship (Koch-Weser and Blinks, 1963). However, CDF or related phenomena have also been described for voltage-gated Ca2  channels in neurons (Cuttle et al., 1998), smooth muscle cells (McCarron et al., 1992), and adrenal glomerulosa cells (Wolfe et al., 2002). Although not described in neurons, CDF of L-type channels could play a major role in supporting their privileged status in mediating excitation–transcription coupling and long-term synaptic plasticity (Bradley and Finkbeiner, 2002; West et al., 2002; Deisseroth et al., 2003).
4 e5 ?: g" L' z; K4 ~3 I  d, a1 x. W% d
We have presented several new findings that advance our understanding of CDF of L-type channels. First, CaMKII associates with the pore-forming 1C subunit of L-type channels in the heart as indicated by coimmunoprecipitation. Second, specific regions of the 1C subunit have the capability to directly anchor activated CaMKII. Third, CaMKII can phosphorylate 1C in regions previously implicated in regulating channel function. Fourth, a mutation in the COOH terminus of 1C that disrupted CaMKII binding to that region completely abolished CDF. Fifth, once tethered to the COOH terminus, CaMKII can be completely dephosphorylated and deactivated, even though it persists in its association and retains its dependence on Ca2 /CaM. Thus, we conclude that the localization and targeting of CaMKII to the COOH terminus of the L-type channel is critical for CDF. Our experiments suggest that individual L-type channels can take advantage of CaMKII as a frequency detector for the activity-dependent regulation of their Ca2  influx. The tethered kinase provides a local and specific integrator of preceding channel activity that controls future channel function through feed-forward autoregulation.
, l; X7 J) L, ?2 f
0 h4 y6 m) J2 N9 _" KA working model for unifying disparate observations on CDF
% H1 R1 b$ t7 J  Q' a3 S. ^. A$ j; T' E: e3 Q+ E, |" ]4 e2 ^  T/ q
Our findings provide a biochemical and molecular explanation of earlier findings that suggested that CDF was mediated by CaMKII. Ca2  buffer experiments revealed that CDF depended on a calcium signal near the channel (Hryshko and Bers, 1990). Pharmacological inhibition of CaMKII abolished CDF (Anderson et al., 1994; Xiao et al., 1994; Yuan and Bers, 1994). Immunostaining showed that autophosphorylated CaMKII was concentrated near the surface membrane of cardiomyocytes (Xiao et al., 1994; Vinogradova et al., 2000). More recently, Dzhura et al. (2000) found that direct application of thiophosphorylated (constitutively activated) CaMKII to the cytoplasmic face of cardiac myocyte membranes induced a high open probability (Po) mode of L-type channel activity, thereby accounting for CDF; the modulatory effect could be prevented by nonhydrolyzable ATP analogues or CaM kinase blockers, further implicating the importance of phosphorylation by CaMKII.& s+ O% V& I% M8 M2 Y

& w, H+ U- S8 A7 E# w2 `Our results not only uncover key molecular underpinnings of those earlier studies but also resolve several unanswered questions. How can a ubiquitous CaMKII fulfill the requirement for a local Ca2  signal in CDF (Hryshko and Bers, 1990; Vinogradova et al., 2000)? Is autophosphorylated CaMKII concentrated near the cell surface (Xiao et al., 1994; Vinogradova et al., 2000) simply because Ca2  is highest near sites of influx (Hryshko and Bers, 1990)? Is a membrane localization of CaMKII achieved by tethering to L-type channels, and is such targeting necessary for CDF? Does CaMKII mediate CDF by directly phosphorylating the pore-forming 1C subunit or an auxiliary protein (Anderson et al., 1994)?
" X2 s6 V& I* A1 [1 `
: g0 k0 s" ~1 ]7 ~2 [8 a3 E0 XTentative answers to these questions can be put forward in the context of a working hypothesis that emerges from our findings on L-type channel–CaMKII interactions (Fig. 8). In a quiescent excitable cell, CaMKII is free in the cytoplasm (Fig. 8, bottom left) inasmuch as the inactive form of the kinase did not significantly interact with any of the cytoplasmic regions of 1C. After an initial Ca2  entry, recruitment to the channel takes place in an activity-dependent manner. CaM binding to soluble CaMKII targets the kinase to certain intracellular domains of 1C, and if the depolarization frequency suffices to produce CaMKII autophosphorylation on Thr286, the resulting displacement of the kinase's autoregulatory domain exposes a potent anchoring site for the 1C COOH terminus (Fig. 8, bottom middle). Observations that autophosphorylated CaMKII is concentrated at the myocyte sarcolemma (Xiao et al., 1994; Vinogradova et al., 2000) can be explained at least in part by a direct interaction of the kinase with 1C. Moreover, the requirement for a local Ca2  signal to trigger CDF (Hryshko and Bers, 1990; Vinogradova et al., 2000) would arise if the necessary phosphorylation could only be achieved by a tethered kinase that is modulated by CaM molecules in the immediate vicinity of the channel-anchored CaMKII.
& B, P' J/ _3 o9 S: P4 w) z
; `7 E! r6 @3 ^6 IFigure 8. Proposed mechanism of CaMKII binding to 1C to form a local and dedicated Ca2  spike integrator for CDF. A catalytic core and autoregulatory domain for a prototypical CaMKII inactive subunit is shown on the bottom left (inactive is indicated by green). Ca2 /CaM activation and Thr286 autophosphorylation displace the CaMKII autoregulatory domain within the catalytic lobe to activate the subunit (yellow) and to expose an 1C tethering site. The CaMKII holoenzyme remains bound to the 1C COOH terminus even after removal of the Ca2 /CaM stimulus, and CaMKII dephosphorylation produces an inactive subunit. CaMKII may remain tethered to other cytoplasmic domains of 1C as well. High depolarization frequencies would produce a threshold level of activated/autophosphorylated CaMKII subunits that increase the Po of the channel via phosphorylation of the NH2 and/or COOH termini (top left). At low depolarization frequencies and under the influence of phosphatase action, CaMKII activation would not be produced, favoring a low Po for 1C (top right).
' Z, Z+ ]6 a8 v4 z* W6 K$ \
: a3 h7 Z& S2 A6 POnce established, this interaction may persist even after Ca2  is lowered and the kinase is completely dephosphorylated (Fig. 7 D), so that CaMKII remains tightly tethered to the channel as long as the cell is intermittently active (Fig. 8, bottom right). This scenario capitalizes on the dodecameric structure of the CaMKII holoenzyme (Kolodziej et al., 2000) by using one or more kinase subunits for the purpose of subcellular localization (Fig. 8, top left). The existence of multiple CaMKII interaction sites on 1C (Fig. 3) may serve to couple the channel and the kinase more tightly and/or orient the large, dodecameric kinase for efficient phosphorylation. The securing of CaMKII in close proximity to key substrate sites on intracellular loops of the channel protein produces a high rate of channel phosphorylation and promotes a pattern of gating with high Po (mode 2; Dzhura et al., 2000). Lowering of the frequency of Ca2  influx reduces kinase activation and allows phosphatases to prevail in dephosphorylating both the channel and its associated CaMKII, driving the channel into a low Po gating mode (Fig. 8, top right). Because the resident CaMKII can be fully dephosphorylated while remaining associated with the channel, its modulatory activity can be graded over the widest possible working range. By virtue of its position, the anchored kinase has a tremendous kinetic advantage over cytosolic CaMKII molecules and essentially monopolizes the modulatory function. Accordingly, a mutation in 1C that rendered the cytoplasmic tail unable to bind CaMKII completely abolished CDF (Fig. 5). Thus, tethering of CaMKII to the COOH terminus of the channel is critical for making it competent for CDF. The combined channel–kinase complex represents a dedicated frequency detector that responds specifically to local Ca2  signaling.
/ q6 u+ _! G. [$ G  Y5 C  o' ~# l& o
Looking beyond Ca2  channels in surface membranes, Ca2  sequestration into intracellular Ca2  stores undergoes a frequency-dependent acceleration in myocardial cells, which is also critically dependent on CaMKII (DeSantiago et al., 2002). It remains unclear whether this action of CaMKII depends on activity-dependent targeting and whether frequency-dependent modulation is a common feature of Ca2  signaling proteins (Maier and Bers, 2002).
9 ]+ Q- b, ^/ b1 S9 d' k6 E6 B
2 \. b7 D2 u3 n, eComparisons with L-type channel modulation by other kinases
4 R& }8 s: G+ N& {# A1 Y2 d$ q  c3 n$ x2 w1 w  H+ F" i3 L7 s! p, I- U) u
The tethering of CaMKII to 1C adds some unique elements to the repertoire of mechanisms used by signaling molecules to link stimulus to cellular response. The L-type channel–CaMKII interaction takes advantage of the multimeric CaMKII holoenzyme, using one or a limited number of its 12 catalytic subunits for anchoring and therefore circumventing the use of auxiliary proteins such as AKAPs or receptors for activated protein kinase C, which tether PKA or PKC, respectively (Bunemann et al., 1999; Tavalin et al., 1999; Schechtman and Mochly-Rosen, 2001; Dorn and Mochly-Rosen, 2002). Another distinction lies in the persistent tethering of CaMKII and its catalytic domains to 1C. The spatial zone of catalytic activity is delimited by the distance from site of anchored subunit to most distant subunit of that holoenzyme. Dissociation of the PKA R2C2 complex from AKAPs leads to the immediate loss of catalytic localization once the C subunits are liberated and thereby activated over a much larger spatial volume. This mechanism is ideal for enabling catalytic subunits to diffuse from the site of activation to the nucleus (Harootunian et al., 1993) and is acceptable if ?-adrenergic potentiation of L-type Ca2  currents (Gao et al., 1997; Hulme et al., 2002) requires rapid responsiveness but only on infrequent occasions. The persistent tethering of the CaMKII holoenzyme might be better suited for continuous operation as an integrator of L-type Ca2  channel activity, endowed with briskly reversible Ca2  responsiveness and dedicated to a limited number of channels.1 a  k% j5 ^2 O8 p
* q$ C4 v; f0 v+ ~
Similarities and contrasts with CaMKII–NMDAR interactions0 l. T6 Z: z, z3 |
7 ~: b1 ?! |( b
Like L-type channels, NMDARs are predominant Ca2  entry pathways in neurons for triggering synaptic plasticity and signaling to the nucleus, and CaMKII is tethered to the NR1 and NR2B subunits of the NMDAR, so our experiments provide interesting points of comparison with previous work showing the direct binding of CaMKII to the NR2B and NR1 subunits of NMDARs (Strack and Colbran, 1998; Leonard et al., 1999, 2002; Strack et al., 2000; Bayer et al., 2001). There are telling similarities between NMDAR subunits and 1C as targets for CaMKII binding. First, completely inactive CaMKII will not initiate binding to any of these subunits. Second, in both NR2B and 1C, a COOH-terminal domain of the membrane protein competes with the autoregulatory domain of CaMKII for binding to the kinase, as shown by peptide competition (Fig. 6 B; Strack et al., 2000). This similarity was highlighted by the finding that a peptide based on the CaMKII binding site on NR2B prevented the kinase from interacting with the 1C COOH-terminal tail (Fig. 6 B). Third, in both NR1 and 1C, the site of CaMKII binding lies close to a site for CaM binding. In the C0 domain of NR1, the amino acids most critical for CaMKII binding lie three residues NH2-terminal to those most important for CaM binding (Leonard et al., 2002). Likewise, the 1C sequence implicated in the CaMKII interaction (Mut6) lies between stretches of amino acids, among them the IQ motif, that are critical for CaM tethering and effector action (Peterson et al., 1999; Zühlke et al., 1999, 2000; Pate et al., 2000; Romanin et al., 2000; Pitt et al., 2001; Erickson et al., 2003; Kim et al., 2004). Further studies will be needed to understand how the activity of the anchored CaMKII may be integrated with the Ca2 -sensing properties of the CaM–IQ domain complex for regulation of L-type channel gating and for downstream signaling to nuclear cAMP response element–binding protein (Dolmetsch et al., 2001).
/ P. E% f, k6 y* C, W0 k$ }6 q# k; i! ~2 T$ k# C
There are also critical functional differences between 1C and NR2B in their interaction with CaMKII. Although the COOH-terminal tails of 1C and NR2B use overlapping sites on CaMKII for binding, the two channels exhibit significant differences in kinase activation state requirements and in consequences of tethering. The NR2B COOH terminus displays a high-affinity interaction with CaMKII that merely requires Ca2 /CaM activation of CaMKII, not autophosphorylation (Bayer et al., 2001). In contrast, the COOH terminus of 1C only binds to autophosphorylated CaMKII (Fig. 3). Binding of CaMKII to NR2B alters kinase function, causing maintained kinase activity even in the absence of Ca2 /CaM or autophosphorylation. This is not the case for CaMKII binding to 1C; our experiments show that interaction with the 1C COOH terminus does not circumvent the autoinhibitory function of the bound kinase. The contrasting properties might arise from substantial differences in the respective COOH-terminal sequences of 1C and NR2B (Fig. 6 C) and might offer specific advantages appropriate to the different roles of the two channels. Establishment of sustained CaMKII activity after transient NMDAR signaling seems perfectly appropriate as a means of supporting enduring effects, e.g., long-term potentiation and long-term depression (Lisman et al., 2002). On the other hand, CDF of L-type channels would suffer a significant loss of dynamic range if the 1C COOH-terminal interaction with CaMKII were to cause constitutive kinase activity. The retention of dependence on Ca2 /CaM for enzymatic activity is well suited for the operation of CaMKII as a built-in integrator of the frequency of prior Ca2  signaling (Hudmon and Schulman, 2002a; Maier and Bers, 2002).) J) T$ H" O6 V+ V
4 ?% N- P. n6 o, f1 X
Materials and methods; [# l' z! ]& D) E: I7 L/ s- [
# N! u9 S4 M, c' v; X) W; q
Oocyte recordings
) c! t# O5 v8 Y' k
( B, u* x6 s+ J' X( g! NThe plasmid encoding the rabbit cardiac 1C subunit used for expression in X. laevis oocytes, pCARDHE, was a gift of W. Sather (University of Colorado, Denver, CO). In vitro transcription and microinjection into X. laevis oocytes (provided by J. Riley and S. Siegelbaum, Columbia University, New York, NY) of 1C, the auxiliary Ca2  channel subunits ?1 and 2, were performed as previously described (Zühlke et al., 2000). Before recording whole cell IBa or ICa, oocytes were injected with 25–50 nl of 100 mM BAPTA solution, pH 7.4, to minimize contaminating Ca2 -activated Cl– currents. IBa and ICa recordings were performed essentially as described previously (Zühlke et al., 2000) with a standard two-electrode voltage clamp configuration using an oocyte clamp amplifier (OC-725C; Warner Instrument Corp.) connected through a Digidata 3122A A/D interface (Axon Instruments, Inc.) to a personal computer. IBa and ICa were recorded in the same oocyte. Ionic currents were filtered at 1 kHz by an integral 4-pole Bessel filter, sampled at 10 kHz, and analyzed with Clampfit 8.1.% u  y$ e# i) Y" u
! y8 a3 B0 }" I5 N6 ^
GST fusion proteins2 P9 Y6 Q# v5 Y' u2 _9 o3 s, W

  n4 C; X$ p' Q1 e3 xPCR fragments corresponding to the a1C (available from Genbank/EMBL/DDBJ under accession no. X15539) NH2 terminus (aa 1–154), I-II intracellular loop (aa 435–554), II-III intracellular loop (aa 784–931), III-IV intracellular loop (aa 1197–1250), and two COOH-terminal fragments (aa 1581–1690 and 1669–2171) were cloned into pGEX-4T-1, and GST fusion proteins were generated. The plasmids encoding the COOH-terminal fragments CT5 (aa 1507–1622), CT12 (aa 1509–1905), and CT23 (aa 1622–1905) were provided by M. Hosey (Northwestern University, Evanston, IL).* d- X7 y5 Q$ |  }* T

" K9 X0 j7 J: o- ?7 Y5 k! Q& \+ ?Peptides
5 S. d2 O! S/ @! L6 |& ^' H. t* R% \7 ?! j& c6 w  g) V
Peptides spanning 1C residues 1581–1690 have previously been described (Pitt et al., 2001). The N-methyl-D-aspartate-L peptide (Bayer et al., 2001) and peptides AC-2 (Hanson et al., 1989), AC-3i (Braun and Schulman, 1995), and AC-3c (Braun and Schulman, 1995) have been described elsewhere.) \$ H. F4 ]( q9 W
5 t# D: E3 \6 m
Immunoprecipitation* X6 ^6 t( l" ^& c

9 \4 s/ P% B9 h( [; sRat cardiac sarcolemmal membranes were provided by S.O. Marx (Columbia University, New York, NY). Immunoprecipitation was performed with either an anti-1C (Alomone) or control IgG in 150 mM NaCl, 50 mM Tris, pH 8.0, 1% Triton, and Complete protease inhibitor cocktail (Roche). After SDS-PAGE, calmodulin overlay was performed with biotin-conjugated calmodulin (STI Signal Transduction) and detected with Vectastain ABC kit (Vector Laboratories). HEK 293 cells were transfected with 1C, 2, ?2, and GFP-CaMKII using Lipofectamine 2000 (Invitrogen) as instructed by the manufacturer. After 48 h, they were washed in ice-cold PBS and then lysed in 150 mM NaCl, 50 mM Tris, pH 8.0, 1% Triton, and Complete protease inhibitor cocktail, and immunoprecipitation was performed with the anti-1C antibody (Alamone). After SDS-PAGE, immunoblotting was performed with an anti-GFP antibody (Covance).  D' I8 u3 s% L3 S7 c9 Z. D4 M$ Z

6 P5 |1 F2 d% K0 k- ^0 tExpression and purification of CaMKII8 @5 X6 \1 S3 ~3 K

: }/ P  C- h1 Z% `% r-CaMKII was expressed and purified essentially as described previously (Bradshaw et al., 2002). Additional CaMKII isoforms were generated by transient expression in HEK 293 cells (Sr plasmid containing the , ?, A, C, or B isoforms). After 72 h, cells were lysed in 10 mM Tris/5% Betaine/150 mM sodium perchlorate, pH 7.5, by brief sonication. Cell lysates were centrifuged for 30 min at 14,000 g at 4°C, and the supernatants were aliquoted, snap frozen, and stored at –80°C.4 y7 I1 d7 {  m: u
% a; y; [! O' D% `
GST binding assay
, n4 R$ `+ S/ W3 ]! E% f/ q% ~& A* u# l
The binding reactions were accomplished in Tris-binding buffer (50 mM Tris, 150 mM NaCl, 0.1% T-20, pH 7.4, and 0.1% BSA) containing 20 nM purified CaMKII. The total protein from the HEK 293 cell lysates added to each binding reaction ranged from 9 to 22 μg, as determined by normalizing for the amount of CaMKII activity (Singla et al., 2001). Preautophosphorylayion of CaMKII (purified and lysate) was performed on ice for 5 min in Tris-binding buffer plus 1 mM CaCl2, 5 μM CaM, 1 mM ATP, and 5 mM MgCl2 to restrict the sites of autophosphorylation to primarily Thr286 (Lai et al., 1987; Lou and Schulman, 1989; Ikeda et al., 1991). Final concentration of these components in the binding reaction (1:40) was 0.025 mM CaCl2, 0.125 μM CaM, 0.025 mM ATP, and 0.125 mM MgCl2. The binding reaction was rocked for 1 h at 4°C, and the beads were extensively washed in Tris-binding buffer (2–3 times for 5 min each). CaMKII binding was quantified using densiotometric measurement of band intensity using 1D Image Analysis Software (Eastman Kodak Co.). Multiple exposure times, as well as a standard curve generated by dilution analysis, ensured linearity in the chemiluminescence intensity. One-way analysis of variance was performed, and Dunnett's test was used to identify specific pair-wise differences between the means. Comparison analyses were conducted using SPSS Version 10.1.3 (SPSS, Inc.).
9 }: h& ]! I2 E& m4 u- Y0 k; C  [
Calmodulin binding assay
$ w& r+ o2 t& Y) b# X- E% F* N
5 c4 d9 m# s# r5 a+ \& i3 f$ e/ TThe bound GST proteins–sepharose complex was prepared as described in the previous section. Purified CaM (Singla et al., 2001) was applied in the presence of 1 mM CaCl2 for 1 h before multiple washes of Tris-binding buffer plus 1 mM CaCl2. Immunoblotting was performed as described previously (Pitt et al., 2001).+ K/ Q4 T1 q) L! x' I- ^- |

' B) k0 b9 t7 Y1 D7 UCaMKII phosphorylation of 1C! Y- n! v% M- |3 ~6 w5 q: ?* t9 w4 u7 P
. @! C" v, e  Y3 g
Purified -CaMKII was incubated with bound GST fusion proteins or immunoprecipitated material bound to PKA in the presence of Ca2 /CaM (2 mM/10 μM) and Mg2 /ATP (5 mM/50 μM ATP) plus 10–50 μCi ATP32 for 15 min at RT. For the GST proteins, CaMKII was activated before exposure to the substrate reaction on ice (as described in GST binding assay) to produce an autophosphorylated enzyme. After the phosphorylation, the beads were washed extensively in PBS (plus 5 mM EDTA) and 2x SDS-PAGE sample buffer was added and SDS-PAGE was performed. The gels were Coomassie stained and exhaustively destained. The gels were dried down, and P32-labeled proteins were detected using autoradiography.: @* p; `- G% N" h( v$ ^9 @

8 T) g7 T- [3 c1 ?CaMKII dephosphorylation using PP18 Q6 @: a' @( j5 y1 |

: r  N: X( ~' u$ ^7 VCaMKII was dephosphorylated using a Hisx6-tagged PP1 catalytic subunit construct (provided by A. Nairn, Yale University, New Haven, CT) purified by Ni-NTA affinity chromatography (provided by M. Bradshaw, Stanford University, Stanford, CA).
4 R& M, k) L/ O, V
3 y# l5 K" A8 |8 s" ]& {Acknowledgments
8 K3 r. F7 V, ^, K
$ W& i* w% E4 m: O* f; RThe authors are grateful to Ben Barres and Michael Bradshaw for helpful discussion and comments on this manuscript.
& H  y" W4 G3 z; u# A7 p; I7 Z" Y7 J+ b7 U) e
This work was supported by grants from the National Institutes of Health to H. Schulman, R.W. Tsien, and G.S. Pitt; a grant from the Irma T. Hirschl Monique Weill-Caulier Trust to G.S. Pitt; and an award from the American Heart Association to A. Hudmon.; g; }3 D; v9 d
: m8 z: W1 s$ g: k& c
References
2 I2 D$ a8 r( y# C' I: q0 H4 a, a3 [: Y+ Q* ?1 j/ g/ G
Anderson, M., A. Braun, H. Schulman, and B. Premack. 1994. Multifunctional Ca2 /calmodulin-dependent protein kinase mediates Ca2 -induced enhancement of the L-type Ca2  current in rabbit ventricular myocytes. Circ. Res. 75:854–861.
% `5 b1 ?8 ]( r. y
7 ~- O, D2 Q  }4 X: M; ]Bayer, K.U., P. De Koninck, A.S. Leonard, J.W. Hell, and H. Schulman. 2001. Interaction with the NMDA receptor locks CaMKII in an active conformation. Nature. 411:801–805.
: i/ A" U: ~6 n/ J
$ }  P; I  ~* e: v5 ~3 `Bayer, K.U., P. De Koninck, and H. Schulman. 2002. Alternative splicing modulates the frequency-dependent response of CaMKII to Ca2  oscillations. EMBO J. 21:3590–3597.
# r# [( T* K/ h7 K4 b, e5 i. }8 s# @) `! k+ l" k- Z, _9 I
Bence-Hanulec, K.K., J. Marshall, and L.A. Blair. 2000. Potentiation of neuronal L calcium channels by IGF-1 requires phosphorylation of the 1 subunit on a specific tyrosine residue. Neuron. 27:121–131.' J/ ]1 V; k$ `  m+ H

; @# c! U: s  K0 h1 p% DBradley, J., and S. Finkbeiner. 2002. An evaluation of specificity in activity-dependent gene expression in neurons. Prog. Neurobiol. 67:469–477.
. k; V. i& T9 S9 x3 t, `3 ]* {4 D5 R* O0 d
Bradshaw, J.M., A. Hudmon, and H. Schulman. 2002. Chemical quenched flow kinetic studies indicate an intraholoenzyme autophosphorylation mechanism for Ca2 /calmodulin-dependent protein kinase II. J. Biol. Chem. 277:20991–20998.
( y9 O( v" {4 J
5 T6 J2 c* a+ l8 @- U7 q3 O) N1 wBradshaw, J.M., Y. Kubota, T. Meyer, and H. Schulman. 2003. An ultrasensitive Ca2 /calmodulin-dependent protein kinase II-protein phosphatase 1 switch facilitates specificity in postsynaptic calcium signaling. Proc. Natl. Acad. Sci. USA. 100:10512–10517.; |+ O5 E* [# j- r0 I

# @/ p. ~! c3 W" gBraun, A.P., and H. Schulman. 1995. A non-selective cation current activated via the multifunctional Ca2 -calmodulin-dependent protein kinase in human epithelial cells. J. Physiol. 488:37–55.) l* b- @6 s5 P& r( G

2 d* f+ S& o' {2 H; j" HBunemann, M., B.L. Gerhardstein, T. Gao, and M.M. Hosey. 1999. Functional regulation of L-type calcium channels via protein kinase A-mediated phosphorylation of the ?2 subunit. J. Biol. Chem. 274:33851–33854.  D) H0 K$ s0 }' B' L1 ^. m

' I) J8 l# W0 z6 dCuttle, M.F., T. Tsujimoto, I.D. Forsythe, and T. Takahashi. 1998. Facilitation of the presynaptic calcium current at an auditory synapse in rat brainstem. J. Physiol. 512:723–729.
, S+ z# u) P2 V0 ]9 f
' E9 v" U3 l$ R. s) S. S+ hDe Jongh, K.S., B.J. Murphy, A.A. Colvin, J.W. Hell, M. Takahashi, and W.A. Catterall. 1996. Specific phosphorylation of a site in the full-length form of the 1 subunit of the cardiac L-type calcium channel by adenosine 3',5'-cyclic monophosphate-dependent protein kinase. Biochemistry. 35:10392–10402." N9 b, a2 V( o9 F# m

5 L" C! @. h" `/ jDe Koninck, P., and H. Schulman. 1998. Sensitivity of CaM kinase II to the frequency of Ca2  oscillations. Science. 279:227–230./ @/ g8 U( R! I

3 n# i5 M# x: F6 PDeisseroth, K., P.G. Mermelstein, H. Xia, and R.W. Tsien. 2003. Signaling from synapse to nucleus: the logic behind the mechanisms. Curr. Opin. Neurobiol. 13:354–365.
! x! \/ d! \/ J1 N1 ?7 @$ p' n$ \9 k$ |! N
DeMaria, C.D., T.W. Soong, B.A. Alseikhan, R.S. Alvania, and D.T. Yue. 2001. Calmodulin bifurcates the local Ca2  signal that modulates P/Q-type Ca2  channels. Nature. 411:484–489.
  P1 d4 }, ]4 S+ o$ l. V4 u# U4 N7 @5 w9 P- X! [% [
DeSantiago, J., L.S. Maier, and D.M. Bers. 2002. Frequency-dependent acceleration of relaxation in the heart depends on CaMKII, but not phospholamban. J. Mol. Cell. Cardiol. 34:975–984.
! q& S2 B5 h9 l5 t+ e7 `: V' L" p- X: t4 e$ S- H  v. B& I. s
Dolmetsch, R.E., U. Pajvani, K. Fife, J.M. Spotts, and M.E. Greenberg. 2001. Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science. 294:333–339.
: U# K" X) {! v4 N7 _: l( o) R9 i4 L6 f. a  M
Dorn, G.W., II, and D. Mochly-Rosen. 2002. Intracellular transport mechanisms of signal transducers. Annu. Rev. Physiol. 64:407–429.+ H1 `* n  ^2 v# K) A6 q' }

# z0 C* E: j; B1 l+ G* V% HDzhura, I., Y. Wu, R.J. Colbran, J.R. Balser, and M.E. Anderson. 2000. Calmodulin kinase determines calcium-dependent facilitation of L-type calcium channels. Nat. Cell Biol. 2:173–177.
$ {( k; ?) O1 C( b2 u8 C7 P- ^
8 w0 v5 _6 o; I4 `; G3 V; K7 lEdman, C.F., and H. Schulman. 1994. Identification and characterization of B-CaM kinase and C-CaM kinase from rat heart, two new multifunctional Ca2 /calmodulin-dependent protein kinase isoforms. Biochim. Biophys. Acta. 1221:89–101.
$ ^: l! G# ~" I* Z3 M
& Q( u5 r) Y5 i* B9 E; pErickson, M.G., H. Liang, M.X. Mori, and D.T. Yue. 2003. FRET two-hybrid mapping reveals function and location of L-type Ca2  channel CaM preassociation. Neuron. 39:97–107.7 n; j) r3 r( s; d

! N* L( O* q5 J" m/ C% y+ n* R8 fEshete, F., and R.D. Fields. 2001. Spike frequency decoding and autonomous activation of Ca2 -calmodulin-dependent protein kinase II in dorsal root ganglion neurons. J. Neurosci. 21:6694–6705.
% o$ T6 V4 S# k4 D, @. Y4 z$ l, n% p0 t4 w- C2 z, ]9 l
Feldman, M.D., J.D. Alderman, J.M. Aroesty, H.D. Royal, J.J. Ferguson, R.M. Owen, W. Grossman, and R.G. McKay. 1988. Depression of systolic and diastolic myocardial reserve during atrial pacing tachycardia in patients with dilated cardiomyopathy. J. Clin. Invest. 82:1661–1669.
$ m( z: O7 Z1 x8 @- o! i, q
/ l! C. E$ R! u8 f" P( k5 ]Gao, T., A. Yatani, M.L. Dell'Acqua, H. Sako, S.A. Green, N. Dascal, J.D. Scott, and M.M. Hosey. 1997. cAMP-dependent regulation of cardiac L-type Ca2  channels requires membrane targeting of PKA and phosphorylation of channel subunits. Neuron. 19:185–196.
( k# j9 u: h5 Z) g+ E8 Z% R
; J8 y" K/ R; x- o  g. {& F9 GGao, T., A.E. Cuadra, H. Ma, M. Bunemann, B.L. Gerhardstein, T. Cheng, R.T. Eick, and M.M. Hosey. 2001. C-terminal fragments of the 1C (CaV1.2) subunit associate with and regulate L-type calcium channels containing C-terminal-truncated 1C subunits. J. Biol. Chem. 276:21089–21097.
3 S4 Y+ r0 {* s5 Z  p" D& V) D1 [7 n- {* A7 t3 r1 ^3 C
Glenney, J.R., Jr., and K. Weber. 1983. Detection of calmodulin-binding polypeptides separated in SDS-polyacrylamide gels by a sensitive calmodulin gel overlay assay. Methods Enzymol. 102:204–210., r7 K7 ]) t% i+ v, I6 Y( a# Q
8 h2 J2 U/ j; Q: I2 o% a
Hanson, P.I., M.S. Kapiloff, L.L. Lou, M.G. Rosenfeld, and H. Schulman. 1989. Expression of a multifunctional Ca2 /calmodulin-dependent protein kinase and mutational analysis of its autoregulation. Neuron. 3:59–70.3 r8 h6 Y5 W, v/ e! _/ U
; j3 n+ k5 R5 N  e3 v
Harootunian, A.T., S.R. Adams, W. Wen, J.L. Meinkoth, S.S. Taylor, and R.Y. Tsien. 1993. Movement of the free catalytic subunit of cAMP-dependent protein kinase into and out of the nucleus can be explained by diffusion. Mol. Biol. Cell. 4:993–1002.+ R' V5 N  d' W; X

7 p$ b7 x, {+ R% Q- YHasenfuss, G., C. Holubarsch, H.P. Hermann, K. Astheimer, B. Pieske, and H. Just. 1994. Influence of the force-frequency relationship on haemodynamics and left ventricular function in patients with non-failing hearts and in patients with dilated cardiomyopathy. Eur. Heart J. 15:164–170.' b* M# E8 F" n( Y

- u& U; w$ {: I# mHell, J.W., C.T. Yokoyama, L.J. Breeze, C. Chavkin, and W.A. Catterall. 1995. Phosphorylation of presynaptic and postsynaptic calcium channels by cAMP-dependent protein kinase in hippocampal neurons. EMBO J. 14:3036–3044.7 B; q9 {( n' n6 o4 T+ @( c
5 C* B- O- b6 L" U+ W7 o. o
Hryshko, L.V., and D.M. Bers. 1990. Ca current facilitation during postrest recovery depends on Ca entry. Am. J. Physiol. 259:H951–H961.
6 f! d3 A5 x$ n7 W* F: A
5 ^9 ^) h- Q# N/ j7 @Hudmon, A., and H. Schulman. 2002a. Neuronal Ca2 /calmodulin-dependent protein kinase II: the role of structure and autoregulation in cellular function. Annu. Rev. Biochem. 71:473–510.! c5 ]0 J7 u/ ]/ |

* K5 o. G3 |8 s; z& Q- YHudmon, A., and H. Schulman. 2002b. Structure-function of the multifunctional Ca2 /calmodulin-dependent protein kinase II. Biochem. J. 364:593–611.; T$ k9 D+ g7 F' P5 w

9 |6 u/ a5 d$ Q& V7 C2 yHulme, J.T., M. Ahn, S.D. Hauschka, T. Scheuer, and W.A. Catterall. 2002. A novel leucine zipper targets AKAP15 and cyclic AMP-dependent protein kinase to the C terminus of the skeletal muscle Ca2  channel and modulates its function. J. Biol. Chem. 277:4079–4087.
/ q( |) j, T* }! Z7 y; ?8 C7 Q
- o+ U) F; h1 zIkeda, A., S. Okuno, and H. Fujisawa. 1991. Studies on the generation of Ca2 /calmodulin-independent activity of calmodulin-dependent protein kinase II by autophosphorylation. Autothiophosphorylation of the enzyme. J. Biol. Chem. 266:11582–11588.
4 o5 q' o' G9 o, b
  a; S, t2 S% _: Q- TKim, J., S. Ghosh, D.A. Nunziato, and G.S. Pitt. 2004. Isolation of the components controlling inactivation of voltage-gated Ca2  channels. Neuron. 41:745–754.. t9 r- l, |' y3 A, ]# x7 I; B( r; Z" K

4 x! p- N# q( z. Q  ]% b& m; w& h% vKoch-Weser, J., and J. Blinks. 1963. The influence of the interval between beats on myocardial contractility. Pharmacol. Rev. 15:601–652.; b& ^* D' b$ e) I

, n, N' G* \  f0 _0 ~8 H7 {Kolodziej, S.J., A. Hudmon, M.N. Waxham, and J.K. Stoops. 2000. Three-dimensional reconstructions of calcium/calmodulin-dependent (CaM) kinase II and truncated CaM kinase II reveal a unique organization for its structural core and functional domains. J. Biol. Chem. 275:14354–14359.0 c% H0 f, o. ~1 W" b

: _; C4 l* \( s: [Lai, Y., A.C. Nairn, F. Gorelick, and P. Greengard. 1987. Ca2 /calmodulin-dependent protein kinase II: identification of autophosphorylation sites responsible for generation of Ca2 /calmodulin-independence. Proc. Natl. Acad. Sci. USA. 84:5710–5714.  U/ `0 ^' s$ n: o% P7 |; V9 Z, ]
! m! A7 d% a- X: \1 y4 O
Lee, A., S.T. Wong, D. Gallagher, B. Li, D.R. Storm, T. Scheuer, and W.A. Catterall. 1999. Ca2 /calmodulin binds to and modulates P/Q-type calcium channels. Nature. 399:155–159.
4 W$ x6 g2 k9 f/ x& w- ]+ v# t* q+ a% w( ^% w* u  J2 H* q
Lee, K.S. 1987. Potentiation of the calcium-channel currents of internally perfused mammalian heart cells by repetitive depolarization. Proc. Natl. Acad. Sci. USA. 84:3941–3945.
. x( i; \0 p1 W/ \
; X2 x- `$ t- I( S9 Q, F: s+ O& w2 |Leonard, A.S., I.A. Lim, D.E. Hemsworth, M.C. Horne, and J.W. Hell. 1999. Calcium/calmodulin-dependent protein kinase II is associated with the N-methyl-D-aspartate receptor. Proc. Natl. Acad. Sci. USA. 96:3239–3244.
% _9 n& V2 `1 @# ^% X
% I6 g5 W% g8 Y6 c* ~& oLeonard, A.S., K.-U. Bayer, M.A. Merrill, I.A. Lim, M.A. Shea, H. Schulman, and J.W. Hell. 2002. Regulation of calcium/calmodulin-dependent protein kinase II docking to N-methyl-D-aspartate receptors by calcium/calmodulin and -actinin. J. Biol. Chem. 277:48441–48448.8 c3 n$ v0 x3 T) |

) V: o, b5 N+ sLisman, J., H. Schulman, and H. Cline. 2002. The molecular basis of CaMKII function in synaptic and behavioural memory. Nat. Rev. Neurosci. 3:175–190.
! e; m, r8 D5 }' b3 K
  L# f9 ]- Q5 C2 o; s2 bLou, L.L., and H. Schulman. 1989. Distinct autophosphorylation sites sequentially produce autonomy and inhibition of the multifunctional Ca2 /calmodulin-dependent protein kinase. J. Neurosci. 9:2020–2032.
% {, k, C; J0 C1 y% {: K+ w; S2 X, b* S  J' q0 z# ?
Maier, L.S., and D.M. Bers. 2002. Calcium, calmodulin, and calcium-calmodulin kinase II: heartbeat to heartbeat and beyond. J. Mol. Cell. Cardiol. 34:919–939.& w) M6 T8 m8 l0 f/ R) j2 `* ]

/ E1 U  u, I9 o' n5 H& H; P9 EMarban, E., and R.W. Tsien. 1982. Enhancement of calcium current during digitalis inotropy in mammalian heart: positive feed-back regulation by intracellular calcium? J. Physiol. 329:589–614.9 @! V1 u# F0 ?/ r" q
+ ^6 R! D8 k- L+ ^+ P
McCarron, J.G., J.G. McGeown, S. Reardon, M. Ikebe, F.S. Fay, and J.V. Walsh Jr. 1992. Calcium-dependent enhancement of calcium current in smooth muscle by calmodulin-dependent protein kinase II. Nature. 357:74–77.5 b  f: R  X' H

' {" \/ ?( K. E4 d+ v, x- TMcHugh, D., E.M. Sharp, T. Scheuer, and W.A. Catterall. 2000. Inhibition of cardiac L-type calcium channels by protein kinase C phosphorylation of two sites in the N-terminal domain. Proc. Natl. Acad. Sci. USA. 97:12334–12338.! N# e1 k. L# H: Q. y
9 G" P0 u9 W$ F9 t4 F2 g  w
Meyer, T., P.I. Hanson, L. Stryer, and H. Schulman. 1992. Calmodulin trapping by calcium-calmodulin-dependent protein kinase. Science. 256:1199–1202.0 i; s6 c- U: R# h5 f1 f0 z7 x% c& a

" O6 n) e4 ^: U4 B8 CMiller, S.G., B.L. Patton, and M.B. Kennedy. 1988. Sequences of autophosphorylation sites in neuronal type II CaM kinase that control Ca2 -independent activity. Neuron. 1:593–604.
2 t* O8 B- r$ ]; l' T; O' n0 l/ y) m1 Z3 b6 c
Mulieri, L.A., G. Hasenfuss, B. Leavitt, P.D. Allen, and N.R. Alpert. 1992. Altered myocardial force-frequency relation in human heart failure. Circulation. 85:1743–1750.% s' x* Y+ ]5 X1 d

- C. \, [: N9 P2 FNoble, S., and Y. Shimoni. 1981. The calcium and frequency dependence of the slow inward current &staircase* in frog atrium. J. Physiol. 310:57–75.
3 p; R( t1 g( h0 q* {6 t) R
, f8 Y/ l( i9 }& d% }6 \Pate, P., J. Mochca-Morales, Y. Wu, J.Z. Zhang, G.G. Rodney, I.I. Serysheva, B.Y. Williams, M.E. Anderson, and S.L. Hamilton. 2000. Determinants for calmodulin binding on voltage-dependent Ca2  channels. J. Biol. Chem. 275:39786–39792.! e: Y, Z" j* ?$ D1 H7 J
! H  Q+ M, B" q# z
Peterson, B.Z., C.D. DeMaria, J.P. Adelman, and D.T. Yue. 1999. Calmodulin is the Ca2  sensor for Ca2 -dependent inactivation of L-type calcium channels. Neuron. 22:549–558. (published erratum appears in Neuron. 1999. 22:following 893)
+ e' ]7 V2 P2 C) P& N; M  Z5 L& w( |: `
Pitt, G.S., R.D. Zuhlke, A. Hudmon, H. Schulman, H. Reuter, and R.W. Tsien. 2001. Molecular basis of calmodulin tethering and Ca2 -dependent inactivation of L-type Ca2  channels. J. Biol. Chem. 276:30794–30802.
: ~6 \3 C$ Q! a$ u  i% f8 ]
/ @5 h" @* }( R6 L' E8 fRomanin, C., R. Gamsjaeger, H. Kahr, D. Schaufler, O. Carlson, D.R. Abernethy, and N.M. Soldatov. 2000. Ca2  sensors of L-type Ca2  channel. FEBS Lett. 487:301–306.* O9 F9 O# |' s& N- r+ Q& T

5 ^: t1 ?0 G$ C" M  NRoss, J., Jr., T. Miura, M. Kambayashi, G.P. Eising, and K.-H. Ryu. 1995. Adrenergic control of the force-frequency relation. Circulation. 92:2327–2332.
* d7 e/ A6 h, ~2 V1 N' W- _1 J7 s* }" G5 }( A' `( w7 n7 F
Rotman, E.I., B.J. Murphy, and W.A. Catterall. 1995. Sites of selective cAMP-dependent phosphorylation of the L-type calcium channel 1 subunit from intact rabbit skeletal muscle myotubes. J. Biol. Chem. 270:16371–16377.
1 q- _& O; Q! W3 K  W2 p; [7 M; o
& h  ]9 X& `1 r' eSchechtman, D., and D. Mochly-Rosen. 2001. Adaptor proteins in protein kinase C-mediated signal transduction. Oncogene. 20:6339–6347.8 D" j( \$ O% x; S4 n. H
# w) L8 f8 Y5 \0 r. `. \5 y
Schouten, V.J., and M. Morad. 1989. Regulation of Ca2  current in frog ventricular myocytes by the holding potential, c-AMP and frequency. Pflugers Arch. 415:1–11.2 A: \0 I* [' K( Q3 L, P
( b! X" B: [; W( @! ?
Singla, S.I., A. Hudmon, J.M. Goldberg, J.L. Smith, and H. Schulman. 2001. Molecular characterization of calmodulin trapping by calcium/calmodulin-dependent protein kinase II. J. Biol. Chem. 276:29353–29360.
; x0 l% S1 l2 a- k% T8 |
- y: x1 @4 r  H$ n, K: a6 VStrack, S., and R.J. Colbran. 1998. Autophosphorylation-dependent targeting of calcium/calmodulin-dependent protein kinase II by the NR2B subunit of the N-methyl-D-aspartate receptor. J. Biol. Chem. 273:20689–20692.4 L$ Z& u6 f3 K# d
3 H) m) w; p, b: w& v' y
Strack, S., R.B. McNeill, and R.J. Colbran. 2000. Mechanism and regulation of calcium/calmodulin-dependent protein kinase II targeting to the NR2B subunit of the N-methyl-D-aspartate receptor. J. Biol. Chem. 275:23798–23806.0 Q2 j. ^2 z, d3 s6 [" B

5 W" R4 c9 {; b* [* f9 J' H$ v- FTavalin, S.J., R.S. Westphal, M. Colledge, L.K. Langeberg, and J.D. Scott. 1999. The molecular architecture of neuronal kinase/phosphatase signalling complexes. Biochem. Soc. Trans. 27:539–542.: |( ?5 ^& I; K0 i2 p! _/ R& L7 `7 s
0 Y" W; ^% y8 K; D3 X) T! k
Vinogradova, T.M., Y.-Y. Zhou, K.Y. Bogdanov, D. Yang, M. Kuschel, H. Cheng, and R.-P. Xiao. 2000. Sinoatrial node pacemaker activity requires Ca2 /calmodulin-dependent protein kinase II activation. Circ. Res. 87:760–767.
+ G: a" q/ n1 i: ^: G/ _5 u% @3 y0 q8 P+ N  M+ J
West, A.E., E.C. Griffith, and M.E. Greenberg. 2002. Regulation of transcription factors by neuronal activity. Nat. Rev. Neurosci. 3:921–931.* {7 ~+ ?# I  D2 R1 T
* `! h4 D' A- E( Q
Wolfe, J.T., H. Wang, E. Perez-Reyes, and P.Q. Barrett. 2002. Stimulation of recombinant Cav3.2, T-type, Ca2  channel currents by CaMKIIC. J. Physiol. 538:343–355.
; |: U6 n" r' c' K) I
8 P% o5 z, l- C2 j" NXiao, R.P., H. Cheng, W.J. Lederer, T. Suzuki, and E.G. Lakatta. 1994. Dual regulation of Ca2 /calmodulin-dependent kinase II activity by membrane voltage and by calcium influx. Proc. Natl. Acad. Sci. USA. 91:9659–9663.
" y4 z- O: a/ U2 v: T: Y! b" l; {2 t! N" M# y
Yuan, W., and D.M. Bers. 1994. Ca-dependent facilitation of cardiac Ca current is due to Ca-calmodulin-dependent protein kinase. Am. J. Physiol. 267:H982–H993.
5 X* J  v* R# m3 b7 t$ m. t6 E8 W9 Q
Zhang, T., E.N. Johnson, Y. Gu, M.R. Morissette, V.P. Sah, M.S. Gigena, D.D. Belke, W.H. Dillmann, T.B. Rogers, H. Schulman, et al. 2002. The cardiac-specific nuclear B isoform of Ca2 /calmodulin-dependent protein kinase II induces hypertrophy and dilated cardiomyopathy associated with increased protein phosphatase 2A activity. J. Biol. Chem. 277:1261–1267.7 I, C1 }5 q- ^+ s/ w; N
& H" q5 \7 R- a; \' C" N
Zühlke, R.D., G.S. Pitt, K. Deisseroth, R.W. Tsien, and H. Reuter. 1999. Calmodulin supports both inactivation and facilitation of L-type calcium channels. Nature. 399:159–162.- y$ n/ T* y- L. I1 S
; o% y4 n! V. N  \( B
Zühlke, R.D., G.S. Pitt, R.W. Tsien, and H. Reuter. 2000. Ca2 -sensitive inactivation and facilitation of L-type Ca2  channels both depend on specific amino acid residues in a consensus calmodulin-binding motif in the 1C subunit. J. Biol. Chem. 275:21121–21129.
6 H- g) J6 ]5 x7 {& I( A% }3 q
' v* W& v: W9 B% m4 FZygmunt, A.C., and J. Maylie. 1990. Stimulation-dependent facilitation of the high threshold calcium current in guinea-pig ventricular myocytes. J. Physiol. 428:653–671.(Andy Hudmon1, Howard Schulman1, James Ki)

Rank: 2

积分
69 
威望
69  
包包
1788  
沙发
发表于 2015-6-8 20:09 |只看该作者
淋巴细胞

Rank: 2

积分
118 
威望
118  
包包
1769  
藤椅
发表于 2015-6-14 04:08 |只看该作者
世界上那些最容易的事情中,拖延时间最不费力。  

Rank: 2

积分
88 
威望
88  
包包
1897  
板凳
发表于 2015-6-23 13:18 |只看该作者
干细胞之家微信公众号
越办越好~~~~~~~~~`  

Rank: 2

积分
72 
威望
72  
包包
1942  
报纸
发表于 2015-6-30 10:27 |只看该作者
偶啥时才能熬出头啊.  

Rank: 2

积分
118 
威望
118  
包包
1769  
地板
发表于 2015-7-5 10:24 |只看该作者
快毕业了 希望有个好工作 干细胞还是不错的方向

Rank: 2

积分
136 
威望
136  
包包
1877  
7
发表于 2015-7-9 08:10 |只看该作者
我也来顶一下..  

Rank: 2

积分
72 
威望
72  
包包
1730  
8
发表于 2015-7-20 15:10 |只看该作者
这样的贴子,不顶说不过去啊  

Rank: 2

积分
162 
威望
162  
包包
1746  
9
发表于 2015-7-27 19:35 |只看该作者
慢慢来,呵呵  

Rank: 2

积分
122 
威望
122  
包包
1876  
10
发表于 2015-8-27 21:27 |只看该作者
厉害!强~~~~没的说了!  
‹ 上一主题|下一主题
你需要登录后才可以回帖 登录 | 注册
验证问答 换一个

Archiver|干细胞之家 ( 吉ICP备2021004615号-3 )

GMT+8, 2024-4-28 13:19

Powered by Discuz! X1.5

© 2001-2010 Comsenz Inc.