Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 May 1.
Published in final edited form as: Biophys Chem. 2017 Mar 9;224:1–19. doi: 10.1016/j.bpc.2017.02.006

Calcium Triggers Reversal of Calmodulin on Nested Anti-Parallel Sites in the IQ Motif of the Neuronal Voltage-Dependent Sodium Channel NaV1.2

Liam Hovey 1,+, C Andrew Fowler 2,+, Ryan Mahling 1, Zesen Lin 1, Mark Stephen Miller 1, Dagan C Marx 1, Jesse B Yoder 1, Elaine H Kim 1, Kristin M Tefft 1, Brett C Waite 1, Michael D Feldkamp 1, Liping Yu 2, Madeline A Shea 1,*
PMCID: PMC5503752  NIHMSID: NIHMS862796  PMID: 28343066

Abstract

Several members of the voltage-gated sodium channel family are regulated by calmodulin (CaM) and ionic calcium. The neuronal voltage-gated sodium channel NaV1.2 contains binding sites for both apo (calcium-depleted) and calcium-saturated CaM. We have determined equilibrium dissociation constants for rat NaV1.2 IQ motif [IQRAYRRYLLK] binding to apo CaM (~3 nM) and (Ca2+)4-CaM (~85 nM), showing that apo CaM binding is favored by 30-fold. For both apo and (Ca2+)4-CaM, NMR demonstrated that NaV1.2 IQ motif peptide (NaV1.2|Qp) exclusively made contacts with C-domain residues of CaM (CaMC). To understand how calcium triggers conformational change at the CaM-IQ interface, we determined a solution structure (2M5E.pdb) of (Ca2+)2-CaMC bound to NaV1.2IQp. The polarity of (Ca2+)2-CaMC relative to the IQ motif was opposite to that seen in apo CaMC-Nav1.2IQp (2KXW), revealing that CaMC recognizes nested, anti-parallel sites in Nav1.2IQp. Reversal of CaM may require transient release from the IQ motif during calcium binding, and facilitate a re-orientation of CaMN allowing interactions with non-IQ NaV1.2 residues or auxiliary regulatory proteins interacting in the vicinity of the IQ motif.

Keywords: Molecular Recognition, Allostery, NMR, Titration, Binding, Linkage, Biosensor, Thermodynamics, Free Energy, Sodium Channels, FRET

Graphical abstract

graphic file with name nihms862796u1.jpg

1. Introduction

The human voltage-dependent sodium channel family (NaV) is responsible for the generation and propagation of action potentials.[1, 2] NaV channelopathies include epilepsy [3], “Long QT” syndrome [4], familial autism [5] and pain insensitivity.[6] NaV channels are believed to undergo calcium-dependent regulation[7] mediated by calmodulin (CaM), an essential intracellular calcium sensor.[8] However, the mechanisms of calcium-induced regulation remain poorly understood, and may vary among NaV subtypes.[9]

1.1 Universal Design for Apo CaM binding to NaV IQ motifs

The architecture of the Nav1.2 pore-forming α-subunit (2005 a.a.) is shown in Fig. 1A. It has 4 multi-helical transmembrane domains (green cylinders) that form the ion-conducting pore and contain voltage-sensing domains; they are connected by intracellular linkers.[10] The Nav1.2 C-terminal domain (CTD) contains a 4-helix bundle (EF-like or EFL domain) that preferentially binds FGF12[11] but also binds other members of the FGF homology factor (FHF) family. A solution structure (2KAV[12]) of a fragment of Nav1.2 that includes EFL (Fig. 1B) shows that the next (fifth) helix is connected to EFL by a flexible linker, allowing it to adopt many orientations relative to the 4-helix bundle. A schematic diagram of a fragment of the Nav1.2 CTD sequence below 2KAV shows the relative length and position of the helices within 2KAV that precede the helix containing the IQ motif (IQxxx[R,K]Gxxx[R,K][13]) that binds CaM tightly [1416]. The IQ motif is necessary for several members of the human NaV family to maintain a closed, inactivated state.[17, 18]

Figure 1. Schematics and Structures of CaM Binding to NaV IQ Motifs.

Figure 1

In all structures of CaM, residues 1–75 are blue, 76–80 are black, and 81–148 are red.

A: Cartoon of a voltage-gated sodium channel (NaV) α-subunit (green) with 4 multi-helical transmembrane units DI, DII, DIII, and DIV (cylinders). The C-terminal domain (CTD) contains a 4-helix bundle domain dubbed EF-like (EFL) and an lQ motif bound to apo CaMC (red) with apo CaMN (blue) connected by a flexible linker (black). The NaV inactivation gate (purple) links DIII and DIV, and contains a site for binding (Ca2+)4-CaM.

B: Solution structures of the NaV1.2 EF-like domain (PDB: 2KAV) depicting residues E1777 (blue sphere) to R1881 (red sphere); cylindrical helices are colored in shades of green correspond to the sections labeled H in the linear diagram of the NaV1.2 CTD residues 1777 to 1937 below.

C: Crystallographic structure of apo (calcium-depleted) CaM bound to a CTD fragment of NaV1.5 (PDB: 4DCK); shades of green match panel B. The NaV1.2 CTD in 4DCK was aligned (using PyMOL (Schrödinger, LLC)) to the corresponding residues in 2KAV in panel B based on EFL-located helical residues 1791–1866 and 1855–1866.

D: Overlay of structures having apo or Mg2+-bound CaM associated with NaV IQ motifs (PDB: 2KXW, 2L53, 4DCK, 4OVN, and 3WFN). For structures determined by NMR (2KXW and 2L53), only the minimized average structure is shown. Structures aligned with PyMOL (Schrödinger, LLC) based on position of CaM residues 102–112 and 117–128 (helices F and G).

E: Structures of individual CaM-IQ complexes in overlay in panel D; Mg2+ ions are gray spheres. For 2L53, all deposited models consistent with the NMR data are illustrated. Structures were aligned as in panel D. Spheres highlight IQ motif residues at positions 1, 2, 5 and 8. In, Nav1.2IQp these are I1912 (cyan), Q1913 (green), Y1916 and Y1919 (gold). In Nav1.5IQp and Nav1.6IQp, colors for corresponding residues match WebLogos in panel F.

F: WebLogo 3.3[26] analyses comparing IQ motif sequences for NaV1.2 from 53 species, NaV1.5 from 40 species and NaV1.6. from 45 species (sequence alignments shown in Supplemental Tables S1A, S1B, S1C). The height of each letter denotes relative conservation of residues. A star indicates the fully conserved Q residue of all three channels which is also labeled in panel E. Within the set of residues binding apo CaMC, complete conservation is observed for positions corresponding to Q1913, R1914, A1915, R1917 in NaV1.2.

Five high-resolution structures of apo (calcium-depleted) CaM bound to NaV IQ motifs are available. In all of these structures, CaMC adopts the semi-open tertiary structure also observed in complexes of apo CaMC bound to myosin V [19] and the SK Channel[20]. A crystal structure of apo CaM bound to a CTD fragment of cardiac NaV1.5 (4DCK, Fig. 1C) shows the C-domain of CaM (CaMC) binds to the IQ motif, while the N-domain (CaMN) makes few contacts with any part of the CTD. An overlay of this complex with apo CaM bound to the IQ motifs from NaV1.2[16], NaV1.5[2123] and NaV1.6[24] (Fig. 1D) illustrates that the position of apo CaMN is highly variable relative to CaMC.

The five structures are shown individually in Fig. 1E (note that CaMN is not shown for Nav1.2 (2KXW) because the structure included only CaMC). The solution structure of apo CaM bound to NaV1.5 (2L53) shows the ensemble of models included in the PDB deposit, with CaMN. having no preferred position relative to CaMC. In two crystallographic structures (4DCK and 4OVN) of CaM bound to a fragment of NaV1.5, there are distinct differences in the observed position of CaMN, perhaps attributable to differences in crystal contacts, solution conditions, or an additional auxiliary protein (FGF13U/FHF2) bound to the NaV1.5 fragment in 4DCK (not shown). Residues in NaV1.5 and NaV1.6 that make multiple contacts with CaM (spheres, Fig. 1D) include residues in positions analogous to NaV1.2 Y1916 and Y1919 which are highly conserved within NaV1.2 sequences.

Magnesium was resolved in the calcium-binding sites of two of the structures of apo CaM bound to NaV1.5 (Fig. 1E). In 4OVN, Mg2+ was observed in all 4 of the calcium-binding sites in CaM, whereas, in 4DCK, Mg2+ was seen only in CaMC sites III and IV (assignment of residues in CaMN sites I and II was not complete in 4DCK). Magnesium and calcium have different hydration shells, and chelation of magnesium is accomplished with a different geometric signature in the subset of residues required for pentagonal, bipyramidal chelation characteristic of calcium binding to EF-hand loops ([25] illustrated in Supplemental Figure S1A. In the crystallographic structure (3WFN) of apo CaM bound to NaV1.6 IQp, no metal ions were included in the deposited structure.

For NaV1.2, NaV1.5 and NaV1.6, WebLogos [26] in Fig. 1F illustrate the highly similar sequences preceding and comprising their IQ motifs [IQRAϕRxxK1 (alignments are provided in Supplemental Table S1). All are basic, amphipathic alpha helices (BAA motifs) characteristic of CaM binding domains. For example, the rat NaV1.2 IQ peptide has a calculated pl of 10.75, and a predicted net charge of +8 at pH 7. Consistent with their similarity in sequence, the available high-resolution structures (Fig. 1E) show that regardless of NaV isoform or structural method (NMR, XRD) used for determination, the “IQ” residues are in identical positions. The I (cyan) of NaV1.2 and NaV1.5, and L (cyan) in the equivalent position of NaV1.6 are buried in the semi-open cleft of CaMC. The fully conserved Q (green) residue of each IQ motif is oriented toward the F-G linker of CaM, forming a network of hydrogen bonds.

1.2 What Does Calcium Binding Do to CaM-NaV Interactions?

While these analyses illustrate the conformation of CaM associated with the NaV IQ motif under low “resting calcium” conditions, they do not explain the role of calcium in modulating CaM-NaV interactions. Early studies identifying the preferred sites for CaM binding to the CTD of NaV1.2 showed that sites for apo CaM and (Ca2+)4-CaM were overlapping but slightly different [14], suggesting that (Ca2+)4-CaM might slide towards the C-terminus of NaV1.2 upon binding calcium.

Calcium binding is known to trigger large changes in interhelical angles within both domains of CaM. In hundreds of high-resolution structures of CaM – whether free or bound to a peptide derived from a target protein–calcium-free domains of CaM are always observed to adopt a closed or semi-open conformation. The open tertiary structure is only observed when calcium occupies the loops in the helix-loop-helix motifs. This is illustrated in Fig. 2A, showing apo CaMC alone (1CFC) and bound to a myosin V peptide (2IX7), and calcium-saturated CaMC alone (1CLL) and bound to a CaMKII peptide (1CDM). The alignment by residues in helix G highlights the opening of the CaMC cleft in response to calcium binding; helices E and F are depicted in black. This grid also demonstrates that the orientation of the CaMBD (CaM-binding domain) peptide from myosin V bound to apo CaM is opposite to that of the CaMKII peptide bound to calcium-saturated CaMC. This plasticity in the calcium-dependent interactions of CaM with its targets is a hallmark feature of its role as a hub protein in signaling networks.[2732]

Figure 2. Ca2+-Induced Changes in Calmodulin Structure and Linkage to Target Binding.

Figure 2

A. Schematic showing structural changes of CaMC upon binding Ca2+ (horizontal) and a peptide target (vertical). Structures were aligned by CaM residues 117–128 (helix G). Interhelical angles for helices E and F (http://calcium.sci.yokohama-cu.ac.jp/efhand.html) shown in black are as follows: apo CaMC (1CFC, 54.8°), apo CaMC bound to an IQ motif in myosin V (2IX7, 79.5°), (Ca2+)2-CaMC (1CLL, 89.7°), and (Ca2+)2-CaMC bound to CaMKII (1CDM, 88.5°). For 2IX7 and 1CDM, the termini of the green target peptide are blue (amino) and red (carboxyl).

B. (Ca2+)4-CaM in 1CLL. CaM residues 1–75 are blue, 76–80 are black, and 81–148 are red; Ca2+ ions are yellow.

C. (Ca2+)4-CaM bound to CaMKII peptide in 1CDM. Color scheme the same as panel B, with peptide in green.

D. Alignment of CaMC residues in B and C showing E104 and E140 in 1CLL (orange) and 1CDM (red) as sticks in Ca2+-binding sites. Label gives the average distance between carboxyl oxygens of E140 and the Ca2+ in site IV. CaMC in each structure is shown in gray; structures were aligned using CaM residues 102–112 and 117–128 (helices F and G).

E. Overlay of apo CaM bound to a myosin V IQ motif in 2IX7, and half-saturated CaM (CaMN Ca2+-saturated but CaMC apo) bound to an SK Channel peptide (1G4Y) aligned by CaMC helices F and G.

F. Apo (Mg2+-bound) CaM and (Ca2+)4-CaM bound to NaV1.5 IQ motif (4DCK and 4JQ0). In sites III and IV, Mg2+ is gray, and Ca2+ ions listed in 4JQ0 are hot pink. Structures were aligned as in panel A.

G. Alignment of apo CaMC bound to SK Channel (1G4Y) shown in panel E, and apo (Mg2+ bound CaMC) bound to NaV1.5 (4DCK) and (Ca2+)4-CaM bound to NaV1.5 IQ motif (4JQ0) shown in panel F. Side chains of E104 and E140 in 1G4Y (cyan) and 4DCK (blue) are shown as sticks; only backbone atoms for E104 and E140 were assigned in 4JQ0 (hot pink). Label gives the average distance between carboxyl oxygens of E140 and the Mg2+ ion in site IV in 4DCK. Mg2+ is gray, and Ca2+ assigned in 4JQ0 is hot pink.

H. Dimer of heterotrimers (NaV-FGF13U-CaM) in the 4JPZ crystal structure of (Ca2+)4-CaM bound to NaV1.2 CTD (residues 1777–1937, green) and FGF13U (purple). Circled regions correspond to regions circled in panels I and J.

I. Blue circles highlight multiple contacts (≤ 6Å) between CaM in one heterotrimer of 4JPZ with FGF13U in the other heterotrimer. Heterotrimers are labeled with subscripts “a” and “b” for tracking interactions. They are identical in sequence, and nearly identical structurally.

J. Orange circle highlights highlight contacts (≤ 6Å) between CaMC in one heterotrimer of 4JPZ with CaMC in the other.

1.3 Distinct Responses by CaMC and CaMN

The linker between CaMC and CaMN allows the two domains to adopt many different relative orientations. Full-length (Ca2+)4-CaM may be extended as seen in the crystallographic structure in Fig. 2B where the linker is helical, fully separating CaMN and CaMC. In contrast, Fig. 2C shows the two domains of (Ca2+)4-CaM wrapped around the CaMBD peptide of CaMKII (compact ellipsoidal). However, an overlay (Fig. 2D) of CaMC from Figs. 2B and 2C shows that CaMC adopts an open tertiary conformation in both, and has identical geometry of calcium-chelating residues in the calcium-binding sites, with the bidentate glutamates E104 (site III) and E140 (site IV), each at position 12 of an EF-hand loop (sticks in red or orange), providing two oxygens for holding Ca2+ in the respective site. The geometry of the calcium-binding loops and interhelical angles is very different if calcium is not bound.

The domains of CaM function as distinct calcium-sensors and CaM can adopt a half-saturated state in which one domain is apo while both sites in the other domain are occupied by calcium. An example is the structure of half-saturated CaM (having (Ca2+)2-CaMN and apo CaMC) bound to a fragment of the SK channel (1G4Y). This is compared in Fig. 2E to a structure of apo CaM bound to a peptide from myosin V (2IX7). The apo CaMC domains are identical in tertiary structure, whereas CaMN has an open tertiary conformation with calcium bound in 1G4Y. It adopts different positions relative to CaMC depending on the target peptide bound to CaM. Structures of apo CaMC bound to IQ motifs of NaVs (Fig. 1E) match those observed in 2IX7 and 1G4Y.

Two crystal structures of half-saturated CaM bound to the IQ motif of NaV CTDs have been published in which calcium occupies sites I and II in CaMN. For CaM bound to NaV1.5, Fig. 2F shows an overlay of the IQ motif of NaV1.5 bound to apo CaM (4DCK) and calcium-bound CaM (4JQ0). In both structures, CaMC is in the semi-open tertiary conformation characteristic of apo CaM, and the interface between CaMC and the NaV1.5 IQ motif is identical. Like the overlay shown in Fig. 2E, CaMN in 4JQ0 has an open tertiary conformation with calcium bound and adopts positions relative to CaMC that differ from apo CaMN in 4DCK. The gray spheres depict the position of magnesium while the red depict calcium. Although both 4JPZ (3.02 A resolution) and 4JQ0 (3.84 A resolution) were deposited in the PDB with calcium having 100% occupancy in sites III and IV of CaMC, the authors commented that “These conformational differences of the individual CaM lobes in the NaV1.2/Ca2+ structure suggested that the C-lobe was unlikely to be fully occupied,” (p. 4) and that CaMC “is essentially unoccupied in the NaV1.2/Ca2+ structure” based on anomalous scattering data (p. 6) [33].

1.4 Magnesium and Calcium – Cousins, Not Twins

The conclusion that sites III and IV do not contain calcium is supported by the overlay in Fig. 2G comparing CaMC in 4DCK and 4JQ0 (both bound to the NaV1.5 IQ motif) and CaMC in 1G4Y (bound to an SK Channel peptide). All three structures show CaMC having a semi-open tertiary conformation. The bidentate glutamates (E104 and E140) in sites III and IV (blue 4DCK, cyan 1G4Y) are oriented away from the loop, consistent with the absence of calcium. These orientations are very different from what is observed in calcium-saturated sites shown in Fig. 2D. In 4JQ0, only the alpha carbon of residues E104 and E140 were assigned (the side chain atoms are missing). However the semi-open conformation, and absence of density for chelating carboxyl groups is consistent with the sites being apo.

The tertiary conformation of CaMC bound to the IQ motif in a fragment of NaV1.2 (4JPZ, Fig. 2H) is identical to that observed in 4JQ0 representing NaV1.5. The conclusion that calcium is absent from CaMC in 4JPZ is supported by the Fo-Fc electron density maps around the metals in sites III and IV of chains C and I compared to the same sites in a CaM-alone structure (1CLL; Supplemental Figure S1C). Despite the high concentration of calcium in the crystallization buffer (300 mM sodium acetate, 50 mM TRIS, pH 7.5, 2 mM CaCl2,), all of these data suggest that CaMC maintained the apo state despite calcium binding to CaMN. Thus, 4JPZ and 4JQ0 represent NaV CTD fragments bound to a half-saturated complex of CaM like the complex observed in 1G4Y with an SK Channel fragment bound to half-saturated CaM (Fig. 2E).

However, calcium-saturated CaM is known to bind tightly to the IQ motifs in NaV1.2 and NaV1.5.[15, 16, 23, 3335] A stoichiometric calcium titration of CaM bound to the NaV1.2 IQ motif showed 4 equivalents of calcium were required for saturation ([16], Fig. 7 therein). This raises the question of why sites III and IV in CaM would retain the apo conformation in structures 4JPZ and 4JQ0.

Figure 7. Models of CaM Retention, Release, or Reversal on NaV1.2.

Figure 7

A. Simulations of saturation of the NaV1.2 IQ motif by apo (dashed) and calcium-saturated (solid) CaM based on ΔG values in Table 1. Molecular models show calcium-induced reversal of CaMC relative to the IQ motif peptide.

B. Simulation of calcium binding to CaM bound to the NaV1.2 IQ motif (see Methods). Resting calcium levels indicated by blue bar; green bar centered on 10 03BCM indicates range of neuronal intracellular Ca2+ levels sufficient to achieve 50–75% calcium-saturated CaMC bound to NaV1.2-IQ. Schematics indicate calcium-induced reversal of CaMC on the IQ motif.

C. Schematic for proposed 3-state mechanism of CaM-mediated regulation based on calcium titration of full-length CaM bound to NaV1.2IQp monitored by NMR. Apo CaM binds to the IQ motif via CaMC with no preferred contacts between CaMN and Nav1.2IQp. Calcium binding to sites I and II affects only CaMN; calcium binding to sites III and IV triggers reversal of CaMC on the IQ motif.

D. CaM-target structures representative of the 3 states shown schematically in part C. Target is green with spheres for the I (cyan) and Q (green) residues of the IQ motif. CaM is gray except for CaMC helices F (red) and G (orange). Apo CaM bound to the NaV1.5 IQ motif (4OVN) or an IQ motif of myosin V (2IX7) shows I oriented into the semi-open cleft, and Q oriented towards the FG-linker in semi-open CaMC. The intermediate state of CaM bound to the NaV1.2 IQ motif (4JPZ) retains the same orientation of CaMC as the apo state. Ca2+-saturated CaM adopts an open conformation bound to the IQ motif of NaV1.2 (2M5E) and myosin V (4ZLK) with Q oriented towards the junction between CaMN and CaMC. All 5 structures were aligned by residues corresponding to 1909–1915 of NaV1.2 in 2M5E, where Iis 1912 and Q is at 1913.

E. Alternative models of calcium-mediated response include (Ca2+)4-CaM dissociation from NaV1.2, which might promote (Ca2+)4-CaM binding at the DIII–DIV linker (“inactivation gate”, purple) as seen in 4DJC, or binding at a pair of CaMBD sequences (orange, cyan) that are not in a continuous helix, such as seen in 5SY1, the STRA6 receptor for retinol uptake.

Those structures included a third protein – an intracellular FGF (FHF) isoform that recognizes the EFL domain within NaV. FGF13U was bound to the NaV1.2 CTD in 4JPZ, and FGF12U was bound to the NaV1.5 CTD in 4JQ0. For NaV1.2 (Fig. 2H), the asymmetric unit includes two heterotrimers (called A and B), and there are many close interactions between them. Close contacts between CaM in one heterotrimer with FGF in the other (i.e., CaMa-FGFb and CaMb-FGFa) are shown in Fig. 2I, while interactions between residues in the calcium-binding sites of CaM in each heterotrimer are shown in Fig. 2J. These suggest that a network of contacts observed in the crystal structure may provide tertiary constraints that favor retention of the CaMC-IQ binding interface observed in the absence of calcium.

1.5 Which CaM-NaV recognition determinants are encoded in the IQ Motif?

Because there is no physiological evidence of which we are aware that CaM interacts with FGF bound to a neighboring channel as seen in the dimer of heterotrimers in 4JPZ, it is expected that these interactions would not occur under native conditions in a neuron. Based on the well characterized effect of calcium triggering opening of hydrophobic clefts in the domains of CaM [36, 37], we hypothesize that calcium binding to CaMC bound to the NaV1.2 IQ motif should open the tertiary conformation of the domain to match the cleft observed for structures such as 1CLL or 1CDM, shown in Fig. 2A. Furthermore, based on the original report of CaM binding sites in NaV1.2 [14] and subsequent titrations and thermodynamic linkage studies [7, 15, 16] showing differences in free energies of binding apo and (Ca2+)4-CaM, we expect calcium binding to alter the CaMC-IQ interface seen in apo-CaM-IQ complexes in Fig. 1E.

To compare how apo and (Ca2+)4-CaM differ in their recognition of the NaV1.2 IQ motif, we determined calcium-dependent differences in CaM binding to a biosensor comprised of the NaV1.2 IQ motif sequence flanked by YFP and CFP.[38] The exquisite sensitivity of FRET between the autofluorescent proteins allowed direct estimates of equilibrium constants below 100 nM which allows us to distinguish between the very high affinity of both apo and calcium-saturated CaM for the IQ motif.

To understand how calcium alters the interface observed between the NaV1.2 IQ motif and apo CaMC, we focused exclusively on CaMC bound to the IQ motif (no EFL) in the absence of auxiliary proteins (no FGF). We eliminated the potential contribution of crystal contacts by determining a high-resolution solution structure (2M5E) of the complex of (Ca2+)2-CaMC bound to Nav1.2IQp. In this structure, (Ca2+)2-CaMC reverses its orientation on the IQ motif relative to the structure (2KXW) of apo CaMC bound to the same sequence. This pivot in place is a striking structural response to calcium binding.

Our findings expand the repertoire of CaM gymnastics on ion channels, and suggest a biophysical foundation for understanding an early step in the mechanism of calcium-mediated NaV modulation. The available crystallographic structures of half-saturated CaM bound to IQ motifs from NaV1.2 (4JPZ) and NaV1.5 (4JQ0), and our solution studies of calcium-saturated CaM bound to the IQ motif of NaV1.2 suggest that calcium regulates a 3-step transition by binding to CaMN at intermediate calcium levels, and at higher levels, binding to CaMC at the IQ motif, triggering its rotation around the IQ residues. The reversal of orientation of CaMC would move CaMN at the least, and might require release and re-association, allowing CaM molecules to interact elsewhere on NaV such as at the inactivation gate ([3944], other linkers within NaV1.2, or its intracellular tails in a manner similar to CaM interactions with CaV channels [4346])

2. Materials and Methods

2.1 Calmodulin

Full-length, wild-type CaM (paramecium sequence, 89% identical (131/148) to mammalian CaM) was overexpressed and purified as described previously[15, 47] with purity assessed by denaturing electrophoresis, UV/Vis spectroscopy and HPLC. Concentrations were determined using the BCA assay[48] (Pierce Biotechnology; Rockford IL) and by absorbance of denatured CaM in 0.1 M NaOH (ε293.3 of 2330 M−1cm−1).[49]

2.2 YFP-CFP Biosensors Containing NaV1.2 IQ

A biosensor pET21B vector expressing a gene for YFP upstream of a Kpnl site and CFP downstream of an AgeI site was a gift from A. Persechini and D.J. Black.[38, 50] A DNA oligomer optimized for bacterial expression and coding for the protein sequence of rat NaV1.2 residues 1901 to 1927 (KRKQEEVSAIVIQRAYRRYLLKQKVKK) was purchased from Integrated DNA Technology (Coralville, IA) and inserted between the KpnI and AgeI sites (underlined residues constitute the canonical IQ motif sequence, IQxxx[R,K]Gxxx[R,K]). Overexpression in BL21(DE3) cells was induced by addition of IPTG; cultures grew for 40 hours at 18 °C. Cells were lysed at 4 °C by sonication for 80 s in 50 mM HEPES, 100 mM KCl, 50 03BCM EGTA, 5 mM NTA, 1 mM MgCl2; pH 7.4 and clarified by ultracentrifugation (Beckman TI-60, 25K rpm, 45 min, 4 °C). Biosensor expression was verified by SDS-PAGE; concentration was determined by UV-Vis absorbance with ε514 of 83400 M−1cm−1.[51] CaM binding to biosensor in clarified cell lysate was indistinguishable from CaM binding to biosensor purified to homogeneity by nickel affinity chromatography.

2.3 Calmodulin Titrations of NaV1.2 IQ Biosensors

The biosensor concentration ranged from 0.5 nM (for binding apo CaM) to 5 nM (for binding calcium-saturated CaM). Buffer was 50 mM HEPES, 100 mM KCl, 50 pM EGTA, 5 mM NTA, 1 mM MgCl2; pH 7.4, 22 °C, with 1.5 μM BSA and 500 μM DTT. Calcium-saturating conditions included 1 mM CaCl2. Titrations were conducted in a 1 cm Suprasil quartz cuvette (Hellma USA, Inc., Plainview, NY); the solution was stirred continuously with a Teflon stir bar in a water-jacketed cuvette-holder held at 22 °C on a QuantaMaster-4 Steady State Spectrofluorometer System with a xenon lamp (Photon Technologies International; New Jersey, USA). The sample was excited at 430 nm and emission spectra were collected from 450–550 nm. Peaks were observed at 475 nm (CFP) and 525 (YFP) nm. An experimentally determined isoemissive point (typically at 511 nm) was also monitored during the titration. Bandpasses were 4 nm for excitation and 8 nm for emission; integration time was 4 to 8 s after each addition of CaM delivered by positive displacement pipets (Drummond Wiretrol) from stocks representing 10-fold serial dilutions spanning 6 orders of magnitude, from pM to pM. Normalized isotherms were determined from the fractional change in YFP intensity (525 nm) corrected for dilution (based on intensity at the isoemissive wavelength); the net decrease in YFP intensity (Fig. 3A) differed for binding apo and calcium-saturated CaM.

Figure 3. CaM Binding to NaV1.2IQ Biosensor and NaV1.2 IQp.

Figure 3

Nav1.2IQ biosensor samples excited at 430 nm; peaks in emission intensity correspond to CFP (475 nm) and YFP (525 nm).

A. Steady-state emission spectra of Nav1.2IQ biosensor alone (solid green) and saturated with apo CaM (black dashed) or calcium-saturated CaM (black solid) (>103-fold excess).

B. Emission spectra of Nav1.2IQ biosensor alone (solid green) and with >103-fold excess apo CaMN (blue solid) and calcium-saturated CaMN (blue dashed).

C. Equilibrium titrations of biosensor binding by apo full-length CaM (black dashed), CaMC (red dashed), and CaMN (blue dashed).

D. Equilibrium titrations of biosensor binding by calcium-saturated full-length CaM (black solid), CaMC (red solid), and CaMN (blue solid). Simulations in panels C and D are based on fits to the experimental data set shown in each panel. For CaMN, a horizontal line is provided to guide the eye.

E. 15N-HSQC spectral overlay of (Ca2+)2-CaMC (red) and (Ca2+)4-CaM (black) bound to Nav1.2IQp (black). Backbone resonance assignments are shown for residues of peptide-bound CaMC. Green ellipses encircle resonances for the same residue in each complex.

F. Expansion of crowded region of spectral overlay in panel E.

2.4 Free Energy of CaM binding to NaV1.2IQ Biosensors

Titrations were analyzed using nonlinear least squares analysis.[52] The fractional saturation ( Y¯) of the biosensor was described by Eq. (1) in which Ka, the association constant, is the reciprocal of Kd, the dissociation constant.

Y¯=Ka[CaMfree](1+Ka[CaMfree]) (1)

To calculate [CaM]free, the independent variables [CaM]total and [biosensor]total were used in an iterative convergence algorithm according to the quadratic equation described in Eq. (2) in which b is (1 + Ka·[biosensor] – Ka·[CaM]total).

[CaMfree]=b±b24Ka([CaMtotal])2Ka (2)

Equation 3 accounted for experimental variations in the observed end points of individual titration curves:

Signal=f(X)=Y[X]low+(Y¯Span) (3)

where Y¯ refers to average fractional saturation of the biosensor and Y[x]low corresponds to the intrinsic fluorescence intensity of biosensor in the absence of CaM. The variable Span represents the magnitude and direction of signal change upon titration (i.e., the difference between the endpoints at high and low [CaM]). Following the signal of fluorescence intensity of YFP at 525 nm, the Span was negative for increasing additions of CaM. The endpoints were fit directly in analysis of all titrations.

The quality of the nonlinear least squares fit was evaluated by judging the square root of the variance, values of the 67% asymmetric confidence intervals, span and randomness of the distribution of residuals, and absolute values of elements of the correlation matrix.[53] The square root of the residual variance was typically less than 0.02. The magnitudes of the confidence intervals were within a factor of two of the standard deviation observed between independent, replicate titrations. All average free energy values reported in Table 1 were based on 4 to 9 replicate determinations from at least two preparations of biosensor overexpressed from independent bacterial colonies. For CaMN, the deflection of YFP intensity was so small at high [CaMN] that it was not possible to estimate the affinity of CaMN for the biosensor.

Table 1.

Free Energy of CaM Binding to rNaV1.2 IQ Motif

Apo CaM Calcium-Saturated CaM

IQ ΔGApo1 Kd2 ΔΔG3 ΔGCa2+ Kd ΔΔG3 ΔΔG4
Ca2+ —Apo
CaM −11.48 ± 0.18 3.2 nM −9.55 ± 0.08 85 nM + 1.93
CaMC −10.72 ± 0.10 11.6 nM +0.76 −8.66 ± 0.09 386 nM +0.89 +2.06
CaMN >100 μM >100 μM
1

Values of ΔG reported in kcal/mol. Averages based on 4 to 7 determinations from at least 2 independent preparations of biosensors, and two independent preparations of CaM. Standard deviations from the mean are reported. Confidence intervals on independent trials ranged from ±.04 to ±.24 kcal/mol, with a median value of 0.10 kcal/mol.

2

Kd (equilibrium dissociation constant) calculated from average value of ΔG reported here.

3

ΔΔG = ΔGCaMc− ΔGCaM

4

ΔΔGCa2+−Apo= ΔGCa2+ − ΔGApo

Solution Conditions: 50 mM HEPES, 100 mM KCl, 50 μM EGTA, 5 mM NTA, 1 mM MgCl2, 1.5 μM BSA, 500 μM DTT, ± 1 mM CaCl2; pH 7.4, 22 °C

The change in normalized fluorescence intensity of CaM or CaMC binding to the biosensor was simulated in Fig. 3 using Eq. (3) that accounted for the plateau intensity of free and CaM-saturated biosensor, as well as the decrease in intensity upon CaM binding. Free energies (and their corresponding equilibrium constants) are given in Table 1. Single-site isotherms of fractional saturation of the IQ motif by apo or calcium-saturated CaM (Fig. 7A) were simulated according to Eq. (1) and the values in Table 1.

2.5 Simulation of Calcium-Binding Isotherm

We previously measured energies of calcium binding to CaM alone under the conditions used in this study; in the absence of the IQ motif, AG1 was −7.76 and ΔG2 was −15.92 kcal/mol.[54] Based on the principle of conservation of energy (shown in the linkage equations below), the difference in energy between apo CaM (ΔGa) and calcium-saturated CaM (ΔGd) binding to the NaV1.2 IQ biosensor (Table 1, ΔΔG +1.93) must be the same as the difference in calcium binding to CaM alone (ΔGc) and to CaM bound to the NaV1.2 IQ biosensor (206Gb).

graphic file with name nihms862796e1.jpg (4)

Thus, for illustration, we simulated a titration curve for calcium binding to sites III and IV of CaM bound to the NaV1.2 IQ motif (in Fig. 7B) with a total free energy of −13.99 kcal/mol and the same cooperativity as observed for CaM alone.

2.6 NMR Samples

Uniformly labeled 15N-CaMC or 13C,15N-CaMC was expressed in E. coli BL21(DE3) cells and contained a slight molar excess of commercially prepared peptide (GenScript, New Jersey). These represented NaV1.2 IQ motif residues 1901–1927, as previously described for NMR determination of the structure (2KXW) of the complex of Nav1.2IQp bound to apo CaMC.[16] In NMR samples, [CaM] was 1.5 mM in buffer composition identical to that used previously[16] (10 mM imidazole, 100 mM KCl, 0.01% NaN3, 50 03BCM EDTA, pH 6.8) in 90% H2O/10% D2O for amide-detected experiments or 100% D2O otherwise. Calcium in matching buffer was added to a final concentration of 3.3 mM, to exceed the concentration of sites (2 per complex).

2.7 NMR Data Acquisition

All experiments were run on either a 500 MHz Bruker Avance II NMR spectrometer equipped with a TXI probe or an 800 MHz Bruker Avance II NMR spectrometer equipped with a TCI cryoprobe. All experiments for assigning both backbone and side chain resonances of the protein were collected at 500 MHz; experiments for assigning the unlabeled peptide, as well as all NOESY experiments, were collected at 800 MHz. All NMR spectra were processed using NMRPipe [55] and analyzed using both Sparky[56] and CCPN Analysis.[57] Chemical shift differences were calculated on the basis of arithmetic differences in chemical shift (H or N) between the two spectra. ΔH and ΔN were scaled by the gyromagnetic ratios of 1H and 15N (normalized to 1H). The combined chemical shift difference, AS, was calculated as ΔH2+(ΔN0.1013)2.

2.8 NMR Resonance Assignments

Backbone spectra (HNCA, HN(CO)CA, HNCACB, HN(CO)CACB, HNCO, and HN(CA)CO) for CaMC were analyzed in Sparky[56], and the resulting peak lists were submitted to the PINE server[58] for automatic backbone assignment. The results were verified manually and corrected as necessary in Sparky. Backbone assignments were then imported into CCPN Analysis [57], which was used for all further assignments.

Side chain resonances were assigned using H(CCO)NH, C(CO)NH, and HCCH-TOCSY experiments, complemented by NOESY data for aromatic rings and methionine epsilons. Unlabeled peptide was assigned using 12C,14N filtered TOCSY (26 and 46 ms spinlock times) and 12C,14N double filtered NOESY (80 and 120 ms mixing times) experiments to suppress signals from labeled protein.[5961] 13C and 15N edited NOESY spectra (120 ms mixing times), as well as a 13C,15N edited, 12C,14N filtered NOESY (140 ms mixing time, [62]), were peak picked and initial assignments were made for unambiguous and somewhat ambiguous crosspeaks (those with 3 or fewer possibilities).

Backbone and non-stereospecific side chain assignments of (Ca2+)2-CaMC were straightforward and nearly complete (91% of all protons). As indicated in Supplemental Table S2, 35 of 38 missing assignments are possibly degenerate methylene protons, excluding M76 for which only the epsilon methyl is observed. Assignment of the residues for unlabeled Nav1.2IQp was more difficult due to chemical shift degeneracy and the lack of an isotopic label. No assignments were obtained for three residues at each end of the peptide (1901–1903 and 1925–1927). The remainder of the peptide, which contained all residues that contact CaMC, was assigned with sufficient completeness for structure determination (77% of expected protons); 19 of 33 missing peptide assignments may have degenerate chemical shifts and many are in solvent-exposed side chains (Supplemental Table S3).

2.9 Structure Calculations

Initial restraints for structure calculations were generated in CCPN Analysis.[57] Distance restraints were made for all assigned NOEs, and shift matching was used to generate additional ambiguous distance restraints. NOEs were grouped into bins based on measured peak intensities. Dihedral angle restraints were generated for both the protein and peptide using the built in DANGLE routine[63]; a residue-by-residue comparison showed that average values for the protein were highly consistent with those determined using TALOS+.[64] Additionally, a set of hydrogen bond restraints was generated to constrain alpha helices.

High-resolution models were generated using traditional solution NMR techniques of combining distance constraints derived from NOESY spectra with backbone φ and 03C8 dihedral angles derived from chemical shifts. Initial rounds of structure determination were carried out with Aria/CNS[65, 66] using standard protocols and importing restraints directly from CCPN Analysis.[57] Intermediate structures were generated in Xplor-NIH[67, 68] or CNS.[66] The results were used to manually curate restraint tables in an iterative manner. Additional hydrogen bond restraints were included to constrain residues in alpha helices. To verify the validity of these restraints, an ensemble calculated with the hydrogen bond restraints removed yielded a somewhat lower resolution but otherwise similar structure, suggesting that this data is not significantly biasing the final result (data not shown). After multiple rounds of structure calculations, restraint and violation analysis suggested no further changes to the distance restraints. The final ensemble of structures was generated as follows.

An ensemble of 300 structures was generated using CNS [66] from an extended starting structure of calcium-free CaMC using torsion angle dynamics, out of which the 20 with the lowest energies were retained. This stage used the default anneal.inp script and consisted of a brief initial minimization followed by 1000 steps of high temperature dynamics at 50000 K, 1000 steps of cooling from 2000 K to 0 K, and finally 2000 steps of minimization. For the 20 lowest energy structures from this calculation, calcium ions were then added and placed very roughly in the calcium-binding loops. An additional set of restraints for positioning the calcium ions was derived from a crystal structure of calcium-saturated CaM (1EXR.pdb)[69] and included in a further CNS [66] refinement protocol using Cartesian dynamics and consisting of 2000 steps at 2000K, 2000 cooling steps from 1000 K to 0 K, and finally 2000 steps of minimization. The final force constants used were 50 kcal mol−1 Å−1 for distance restraints and 200 kcal mol−1 rad−2 for torsion angle restraints. CNS [66] reported no backbone dihedral angle violations greater than 2° and no distance violations greater than 0.2 Å in this final ensemble. Complete structure statistics are reported in Table 2.

Table 2.

Structural Statistics for 2M5E* for the final ensemble (20 lowest energy structures)

Experimental restraints
 CaM NOEsa
  Intraresidue (i = j) 492
  Sequential (|i-j| =1) 393
  Medium range (1 < |i-j| < 5) 415
  Long range (|i-j| ≥ 5) 399
  Total CaM NOEs 1699
 Peptide NOEs 279
 Intermolecular NOEs 97
 Total unambiguous NOEs 2075
 NOEs with multiple assignments 651
 Total NOEs 2726
 Distance restraints per residue, CaMa 23.6
 Distance restraints per residue, peptideb 13.3
 phi/psi angles, CaM 136
 phi/psi angles, peptide 38
Restraint violations ensemble minimized average
 NOE distances violated > 0.2 Å 0 0
 dihedral angles violated > 2° 0 0
RMSDs from experimental restraints
 NOE (Å) 0.0039 ± 0.0001 0.004
 Dihedrals (°) 0.068 ± 0.021 0.082
XPLOR energies (kcal/mol)
 overall energy 50.63 ± 1.33 54.93
 bonds 0.74 ± 0.06 1.00
 angles 41.10 ± 0.28 41.92
 impropers 1.02 ± 0.08 1.09
 van der Waals 5.50 ± 0.96 8.67
 NOEs 2.21 ± 0.12 2.17
 dihedrals 0.05 ± 0.03 0.07
 RMSD from idealized geometry
  Bond lengths (Å) 0.0007 ± 0.0001 0.0008
  Bond angles (°) 0.300 ± 0.001 0.303
  Impropers (°) 0.090 ± 0.003 0.093
 Ramachandran plot statistics (%)c
  Most favored regions 92.0 90.6
  Additionally allowed regions 7.6 9.4
  Generously allowed regions 0.4 0.0
  Disallowed regions 0.1 0.0
  Coordinate RMSD (average difference to mean, Å)c
  Backbone atoms 0.643
  Heavy atoms 1.188

Restraint violations, RMSDs and energies are for all residues in the construct. Structural statistics are taken from the output of CNS[66], with the exception of the Ramachandran statistics which come from PROCHECK-NMR. [70]

a

for CaM residues 77–148 (all except M76)

b

for peptide residues 1904–1924 (all assigned)

c

for structured residues (CaM residues 79–147, peptide residues 1904–1924).

The final structures have good Ramachandran statistics as determined by PROCHECKNMR [70] (Table 2 and Supplemental Fig. S2). For the entire ensemble of 20 structures, 99.6% of non-glycine residues fall in the most favorable or allowed regions of torsion angle space (100% for the minimized average structure). All residues falling in the generously allowed (0.35%) and disallowed (0.05%) regions correspond to either Q1904 or K1924 which are only partially assigned and are at the disordered ends of the peptide. The PROCHECK-NMR [70] analysis omitted residues 76–78 and 148 in CaMC, which were poorly ordered, and residues 1901–1903 and 1925–1927 in Nav1.2IQp, for which no NMR data or restraints are available.

All molecular models were drawn with PyMOL Molecular Visualization System v. 1.7.2.3 (Schrödinger, LLC).

3. Results

3.1 FRET Intensity of CaM-Biosensor Assemblies

To evaluate separable roles of CaM domains in recognition of the NaV1.2 IQ motif, and to estimate equilibrium binding constants for both apo and calcium-saturated CaM and its domains, we monitored CaM-induced disruption of FRET in a biosensor protein[38, 50] that was engineered to contain the desired NaV1.2 sequence (residues 1901–1927) bracketed by YFP and CFP (see Methods). Steady-state emission spectra (λex = 430 nm) for the NaV1.2IQ biosensor alone (green, 4 nM) with λmax at 475 nm for CFP and 525 nm for YFP are shown in Fig. 3A.

FRET intensity is sensitive to small changes in distance between fluorophores induced by CaM binding to the IQ motif sequence in the Nav1.2IQ biosensor. We have shown previously that the IQ motif sequence is intrinsically disordered.[15] This allows CFP and YFP to be in close proximity for efficient non-radiative transfer of energy. The typical change in intensity of the biosensor after saturation with apo (black, dashed) or calcium-bound (black, solid) CaM is shown in Fig. 3A ([CaM]:[biosensor] = 103:1). The emission spectra were buffer-subtracted and corrected for dilution; the spectra were normalized to the peak intensity of free biosensor at 525 nm. The intrinsic fluorescence of CaM (which contains Phe and Tyr, but not Trp) does not contribute measurably to the intensities observed.

At λmax for YFP, the fractional decrease in intensity was consistently larger for saturation by apo CaM (36±1%) than (Ca2+)4-CaM (21±1%). We infer that apo CaM and (Ca2+)4-CaM differ in conformation when bound to the IQ motif, and thus differentially reduce energy transfer from CFP to YFP by changing their relative orientation or spatial separation. The isoemissive point for free and CaM-saturated biosensor was observed consistently at 511 nm throughout the titration, indicating that it behaved as a two-component solution.

3.2 Domain-Specific Binding of CaM

To investigate whether both domains of CaM were necessary for the change in fluorescence intensity seen upon binding full-length CaM, a C-domain fragment (CaMC, residues 76–148) or an N-domain fragment (CaMN, residues 1–75) was added to the Nav1.2IQ biosensor (4 nM) in large excess ([CaM]:[biosensor] = 103:1). The emission spectra of the Nav1.2IQ biosensor with apo and calcium-saturated CaMC bound (data not shown) were very similar to those in Fig. 3A for full-length CaM. However, Fig. 3B showed a negligible change (~1%) in intensity of the Nav1.2IQ biosensor after addition of apo CaMN (dashed blue), and a net change of ~3% for calcium-saturated CaMN (solid blue) at a concentration of 15 03BCM CaMN. The deflection in fluorescent intensity at 50 pM CaMN was less than 1% of that seen at 50 nM CaMC, showing high specificity for CaMC. These results are consistent with our prior studies using fluorescence anisotropy to study CaM and its domains binding to a fluorescein-labeled peptide corresponding to the same Nav1.2 IQ motif sequence [15, 16]

3.3 Energetics of CaM Binding

Thermodynamic linkage analysis previously showed that apo CaM binds more favorably than calcium-saturated CaM to rat Nav1.2IQp.[1416] However, the free energies of binding could not be determined quantitatively because the high concentration of Nav1.2IQp needed to obtain sufficient signal in fluorescence anisotropy or gel electrophoresis mobility shift experiments resulted in stoichiometric titrations. In contrast, the strong fluorescence intensity of the Nav1.2IQ biosensor permitted titrations to be conducted at concentrations as low as 0.5 nM (see Methods). The fractional change in intensity at λmax for YFP (normalized to biosensor alone at 100% and CaM-saturated biosensor as zero) was plotted as a function of total concentration of added CaM or CaMC, under either apo (Fig. 3C), or high calcium (1 mM, Fig. 3D) conditions.

These binding isotherms allowed direct estimates of Gibbs free energies of binding (AG) and corresponding dissociation constants (Kd) under equilibrium conditions (Table 1). Apo CaM binding to the Nav1.2IQ biosensor (−11.48 kcal/mol (Kd 3.2 nM) was more favorable by −1.93 kcal/mol than the binding of calcium-saturated CaM (−9.55 kcal/mol; Kd 85 nM). The difference was significantly larger than the standard deviations (0.18 and 0.08 kcal/mol) of multiple replicate titrations for apo and calcium-saturated CaM, respectively. The difference similarly exceeded the confidence intervals for individual determinations (median value of 0.10 kcal/mol), and suggested that the interface between apo CaM and the IQ motif changed when sites III and IV filled with calcium.

Because CaMN showed negligible binding to the Nav1.2IQ biosensor (Fig. 3B), it was not possible to observe a titration or estimate a free energy of interaction. However, to visually compare its behavior to that of CaM and CaMC under apo (Fig. 3C), or high calcium (1 mM, Fig. 3D) conditions, the small change in intensity observed upon addition of CaMN was normalized by assuming that the YFP signal of the biosensor would undergo the same deflection as observed for CaM (i.e., 36% for apo and 21% for calcium-saturated CaM) if it were saturated by CaMN. Even if the slight deflection observed at 10 03BCM CaMN were attributable to specific binding rather than non-specific quenching, the midpoint of a complete titration would be higher than 100 03BCM, corresponding to a very weak dissociation constant. It is also possible that this modest binding represents CaMN binding to the same NaV sequence that is occupied by CaMC when full-length CaM is bound, and thus represents an interaction that would not be populated in the cell where only full-length CaM is available. We observed a similar phenomenon for domain-specific binding of CaM to melittin.[71]

3.4 CaMC Binding to Nav1.2IQp Independent of CaMN

The titrations in Fig. 3 indicated that the driving force for CaM binding to the IQ motif comes primarily from residues in the C-domain of CaM. Although the binding of CaMC was slightly less favorable than CaM for each condition (see Table 1), this may reflect electrostatic differences in attraction for the very basic Nav1.2IQ sequence, rather than differences in close contacts such as van der Waals interactions or hydrogen bonding at the interface of CaM and Nav1.2IQp. The calculated net charge at pH 7 (http://www.scripps.edu/~cdputnam/protcalc.html) of CaM is −23.5 and CaMC is −12.6. Approximating CaM and CaMC as negative point charges binding to positive IQ motifs, the Coulombic contributions (proportional to q1*q2/r2 [72]) to their free energies of binding to the IQ motif would differ by the square of the ratio of their charges (23.5 vs. 12.6), or a factor of ~4. That represents a free energy of −RTln(4) or .813 kcal/mol at 22 °C which corresponds well with the ΔΔG values of 0.76 and 0.89 kcal/mol reported in Table 1.

To assess directly whether the CaM-peptide interface was the same for (Ca2+)4-CaM-Nav1.2IQp and (Ca2+)2-CaMC-Nav1.2IQp, HSQC spectra of 13C,15N-labeled complexes of (Ca2+)4-CaM (black) and (Ca2+)2-CaMC (red) bound to Nav1.2IQp were compared (Fig. 3E). The green ellipses encircle resonances for the same residue in each complex, and show that they are in identical or nearly identical positions, indicating that their chemical environment is very similar. Resonances in the highly overlapped central region boxed by a dashed line are shown in an enlarged format in Fig. 3F.

Consistent with the interface between Nav1.2IQp and CaM being dominated by close contacts with CaMC, resonances within the N-domain of (Ca2+)4-CaM bound to the Nav1.2IQ peptide matched those in a free fragment (Ca2+)2-CaMN (see Supplemental Figure S2). These studies match prior NMR studies showing that CaM binding Nav1.2IQp caused few perturbations of residues within the N-domain of either apo or calcium-saturated CaM.[16].

Therefore, to determine why (Ca2+)2-CaMC binds less favorably than apo CaMC to the NaV1.2 IQ motif, and understand how calcium binding changes the interface between CaMC and the NaV1.2 IQ motif without the influence of auxiliary proteins or flanking regions, we determined a high resolution structure of CaMC bound to a peptide representing the IQ motif (hereafter referred to as the CaMC-Nav1.2IQp complex). NMR allowed determination of the complex in solution without packing constraints that might be contributed by a crystallographic lattice.

3.5 Solution structure of Ca2+-CaMC bound to Nav1.2IQp

The structure was determined using a sample of 13C,15N-(Ca2+)2-CaMC bound to unlabeled Nav1.2IQp (see Methods). Nearly complete assignments of CaMC and central Nav1.2IQp peptide residues that interact with CaMC were obtained (see Supplemental Tables S2 and S3). Backbone traces of the 20 NMR-derived lowest energy structures of the complex are superimposed in Fig. 4A in two orientations: with Nav1.2IQp perpendicular and parallel to the plane of the figure. The Ramachandran plot is in Supplemental Figure S3; Table 2 provides structural statistics. In the minimized average NMR structure of the complex (Fig. 4B), (Ca2+)2−CaMC adopts a 4-helix bundle in the classic open conformation (see Fig. 2A), with side chains in the EF-hand sites consistent with the metal-coordination geometry adopted when calcium is bound (see Supplemental Fig. S1B). Residues 1904–1923 of Nav1.2IQp form an amphipathic α-helix closely associated with (Ca2+)2-CaMC. The terminal residues (1901–1903, 1924–1927) of Nav1.2IQp were disordered, as had been observed for Nav1.2IQp bound to apo CaMC.

Figure 4. NMR Models of Ca2+-Saturated CaMC Bound to NaV1.2IQp.

Figure 4

A. Wireframe backbone representations of the 20 lowest energy structures and the ensemble minimized average model of (Ca2+)2-CaMC bound to Nav1.2IQp (2M5E). For clarity, only well ordered helical residues 1904–1923 of Nav1.2IQp are shown.

B. Cartoon representation of minimized average model in 2M5E. Nav1.2IQp helix shown as a ribbon with gradient from blue (N-terminus) to magenta (C-terminus). Cylinders represent CaM helices E (gray), F (red), G (orange) and H (gray) with yellow spheres for calcium ions. Residues of the IQ motif making ≥ 3 short-range contacts (< 4.5 Å) with the hydrophobic cleft of CaM are shown in ball-and-stick representation along with Q1913 for reference.

C. Analysis by CSU (Contacts of Structural Units)[73] of the minimized average model of (Ca2+)2-CaMC-Nav1.2IQp. The primary sequence of residues 1904 to 1923 of the IQ motif is shown on the x-axis (I1912 in cyan, Q1913 in green). Boxes listing residues of CaM located within 4.5 Å of each IQ motif residue are shaded in orange; longer range contacts (4.5 to 6 Å) are in white boxes.

D. NMR strip plots corresponding to three CaM residues: A88α, V91γ2 and M124ε. For each residue, the top panel represents a 13C-edited, 12C,14N-filtered NOESY experiment, the middle panel represents an HCCH-TOCSY experiment, and the bottom panel is a 13C-edited NOESY with carbon chemical shift of the planes denoted in the label. Labels for atoms in Nav1.2IQp are green and in CaM are red. Residues surrounding A83, V91 and M124 of CaM are shown in images made with PyMOL (Schrödinger, LLC) next to the strip plots.

Analysis of the minimized average structure with CSU (Contacts of Structural Units [73]) in Fig. 4C showed residues in Nav1.2IQp (bottom row) that are in the binding interface with CaMC, and the vertical stacks show CaMC residues that have atoms within 6 A of Nav1.2IQp; contacts that were ≤ 4.5 Å are shaded orange.

The side chains V1911, I1912, A1915 and Y1916 (Fig. 4B) make many contacts with the hydrophobic cleft of CaMC, including the FLMM residues (F92, L112, M124, M144) observed to bind a large number of peptides in calcium-saturated CaMC complexes ([7476]). Residue Q1913 in the IQ motif of Nav1.2IQp points towards the C-terminus of CaMC.

For determining the orientation of CaMC relative to the peptide, the key intermolecular NOEs observed between CaMC and Nav1.2IQp were those associated with CaMC residues A88, V91 and M124 (Fig. 4D). Nav1.2IQp has only two aromatic residues (Y1916 and Y1919) with distinct chemical shifts, each of which makes unambiguous NOE contacts to both A88 and V91. The side chain of A88 comes within 2.4 Å of Y1916 (vs. 7.8 Å in 2KXW), and the V91 side chain is within 3 Å of Y1919 and just over 5 Å from Y1916 (both >10 Å in 2KXW). Furthermore, the side chain of M124 makes contacts with Hα and both Hβs of S1908, which is the only serine in Nav1.2IQp and more than 11 A away from M124 in the corresponding apo structure (2KXW).

To see the overall effect of peptide binding on (Ca2+)2-CaMC, a 1H-15N HSQC of (Ca2+)2-CaMC bound to Nav1.2IQp was compared to a corresponding spectrum of (Ca2+)2-CaMC alone (Fig. 5A). The chemical shift perturbations observed for each backbone amide upon binding to Nav1.2IQp localized mainly in helices E, F, and H (Fig. 5B) and the largest differences in chemical shift were in regions contacting Nav1.2IQ (Fig. 5C). The resonances for residues within sites III and IV in these two samples were nearly identical regardless of the peptide binding, suggesting that the geometry of calcium-chelation is also identical. As noted in Methods, NMR does not directly visualize the position of metal ions. Therefore, the positions of calcium ions shown in Fig. 5C were fit based on the side chain geometry of highly acidic coordinating residues. Such residues would undergo electrostatic repulsion if there were no divalent cation present.

Figure 5. Chemical Shift Differences and Contacts in (Ca2+)2-CaMC ± Nav1.2|Qp.

Figure 5

A. HSQC spectral overlay of (Ca2+)2-CaMC alone (red) and bound to Nav1.2IQp (black). Backbone resonance assignments are shown for residues of peptide-bound CaMC.

B. Effects of Nav1.2IQp on chemical shifts of (Ca2+)2-CaMC. Changes in chemical shifts are calculated as described in Methods. Location of CaM helices and calcium-binding loops are shown above the plot. Residues in calcium-binding sites III (93–104) and IV (128–140) are indicated by yellow shading.

C. Chemical shift perturbations mapped onto the structure of (Ca2+)2-CaMC in 2M5E. The greatest effect of Nav1.2IQp binding is indicated by the red/wide cartoon, and the weakest effect is shown by the blue/narrow cartoon.

3.6 Reversal of CaMC on the IQ motif

The most striking feature of the solution structure (2M5E) of calcium-saturated CaMC bound to Nav1.2IQp was the orientation of (Ca2+)2-CaMC relative to Nav1.2IQp. According to the NOEs shown in Fig. 4D, it was opposite to that observed in apo CaMo-Nav1.2IQp in 2KXW.[16] A nearly 180° rotation of a CaM domain on the same hydrophobic face of a target CaMBD helix was an unexpected response. The protein samples used for determining 2M5E were made by adding excess calcium to an equilibrated complex of apo CaMC bound to Nav1.2IQp that was from the same stock used for determining 2KXW (see Methods). Thus, the final structure shown in Fig. 4 required that (Ca2+)2-CaMC either pivoted in place, or released and re-associated with Nav1.2IQ.

In both 2M5E and 2KXW, the sequence of Nav1.2IQ was sufficiently long to encompass residues that interact closely with CaMC as well as containing terminal residues (1901–1904 and 1924–1927) that were non-interacting and highly disordered. Thus, the Nav1.2IQp sequence appeared to bracket residues needed for tight binding to CaMC. The comparison of HSQC spectra of (Ca2+)2-CaMC and apo CaMC, each in complex with unlabeled Nav1.2IQp (Fig. 6A), showed that few CaM resonances overlapped. A comparison of 12C,14N filtered NOESY spectra for both apo and calcium-saturated CaMC-Nav1.2IQp complexes (partial spectra shown in Supplemental Figure S4) shows striking differences consistent with calcium binding inducing a major conformational change as represented in Fig. 4 and 5.

Figure 6. Comparison of Nested Anti-Parallel Sites in Nav1.2IQp.

Figure 6

A. HSQC overlay of CaMC-Nav1.2IQp with (black) and without (green) saturating calcium. Backbone resonance assignments are shown for residues of (Ca2+)2-CaM-Nav1.2IQp.

B. Chemical shift differences between CaMC-Nav1.2IQp ± Ca2+ indicated changes in chemical environment of individual CaM residues. Location of CaM helices and calcium-binding loops are shown above the plot. Residues in calcium-binding sites III (93–104) and IV (128–140) are indicated by yellow shading.

C. Chemical shift perturbations mapped onto the structure of (Ca2+)2-CaMC in 2M5E. The greatest effect of Ca2+ binding is shown in the cartoon as red/wide, while the weakest effect is blue/narrow.

D. Comparison of (Ca2+)2-CaMC (2M5E) and apo CaMC (2KXW) bound to Nav1.2IQp aligned by superposition of the helical backbone of Nav1.2IQp. Residues I1912 (cyan), Q1913 (green), and A1915 (black) are in ball-and-stick. For clarity, the peptide backbone is omitted. The surface of solvent-exposed CaM residues is red and cleft residues are orange in semi-open (2KXW) and open (2M5E) CaM. CaMN would connect to CaMC via residue M76 (blue) highlighting the reversed orientations of CaMC on the IQ motif.

E. Comparison of clefts of Ca2+-CaMC (2M5E) and apo CaMC (2KXW) binding Nav1.2IQp. Structures were aligned based on helical residues of the peptide. IQ motif residues I1912 (cyan) Q1913 (green), A1915 (black), Y1916 (magenta), and Y1919 (dark green) are shown in ball-and-stick representation. For clarity, the backbone of the peptide is not shown.

Some changes in peak positions reflect the expected effects of calcium occupancy of sites and the opening of the tertiary structure triggered by calcium binding (i.e., semi-open to open conformation) that would occur for CaM alone. However, spectral comparisons of calcium-saturated CaMC ± Nav1.2IQp in Fig. 5A showed that binding of the peptide made additional changes in CaM peaks beyond those caused by calcium binding. The magnitude and direction of the perturbations seen in Fig. 6A reflect binding of calcium at sites III and IV, and binding of Nav1.2IQp in the open cleft of CaMC.

The chemical shift perturbation for each resonance that was assigned in both 2KXW and 2M5E (Fig. 6B) showed many large changes between the structures (average Δ03B4 of 0.5 ppm). With a false-color scale of red (representing greatest) to blue (least) perturbation, Fig. 6C highlights that the major differences outside calcium-binding sites III and IV were in the FG-linker which changes contacts dramatically when CaMC pivots relative to Nav1.2IQ. Consistent with analysis of short-range structural contacts (Fig. 4C) in (Ca2+)2-CaMC (2M5E), residues in the CaMC FG-linker contact A1915 in Nav1.2IQp whereas in apo CaMC (2KXW), the FG-linker made a network of hydrogen bonds with Q1913. Because of the reversal of (Ca2+)2-CaMC on Nav1.2IQp, Q1913 in 2M5E primarily contacts residues near the C-terminus of CaM and the residues that would be closest to CaMN.

Fig. 6D and 6E (90° rotations) show 2M5E and 2KXW aligned according to the backbone of Nav1.2IQp. Positions of selected side chains (I1912, Q1913, A1915, Y1916 and Y1919) that make multiple contacts with CaM are shown in ball-and-stick format. CaMC residue 76 (blue) indicates the junction with CaMN, highlighting the reversal of CaM relative to the channel peptide. The open conformation of (Ca2+)2-CaMC (2M5E) (Fig. 6D) has a more shallow hydrophobic cleft than the semi-open conformation of apo CaMC (2KXW). A view looking into the clefts (Fig. 6E) shows that in apo CaMC (2KXW), the side chains of I1912, Q1913, A1915, Y1916 are tightly clustered in the deep semi-open cleft whereas these residues are distributed over a wider surface of the hydrophobic cleft in 2M5E, with Y1916 oriented closer to Y1919 in the calcium-saturated structure.

Although the rotation of (Ca2+)2-CaMC on Nav1.2IQp was unexpected, the observed polarity of Nav1.2IQp relative to CaMC is consistent with hundreds of other complexes of calcium-saturated CaM bound to peptides representing CaMBDs from other ion channels, receptors and enzymes (see Fig. 2A). These motifs are commonly denoted by a numbering scheme (such as 1–5–8) indicating the positions of residues in the primary sequence that interact closely with CaM clefts. In 2M5E, Nav1.2IQp residues I1912–Y1916–Y1919 correspond to positions 1–5–8 in a typical motif, and I1912 is the sole residue contacting all 4 of the FLMM residues [75] in CaMC.

4. Discussion

CaM is an essential, intracellular, eukaryotic calcium sensor that regulates both aqueous and membrane-spanning target proteins. In turn, those targets are allosteric effectors of the calcium-binding free energies of CaM, making CaM an effective modulator over a range of intracellular calcium concentrations that spans at least three orders of magnitude, bracketing its micromolar affinity for calcium as a free protein.[77, 78] Voltage-dependent sodium channel NaV1.2 retains CaM as a constitutive subunit, but the mechanism of calcium-triggered change in the complex has been perplexing because previous structural studies did not see calcium bound to CaMC which anchors CaM to the channel.

4.1 All IQs are BAAs, But Not All IQs are Equal

The magnitude and direction of a target-linked shift in calcium-binding affinity of CaM and its conformational response when bound to a target is not predictable from the CaMBD sequence alone. All known CaMBDs are BAAs - Basic sequences (pl of 10–12) that form Amphipathic Alpha helices (usually continuous). Many bind exclusively to the open conformation of domains in (Ca2+)4-CaM, with no measurable affinity for apo CaM. Others, such as the NMDA receptor [79] and phosphodiesterase [80], retain apo CaM as a constitutive subunit but have a higher affinity for calcium-saturated CaM than apo CaM which binds primarily via contacts with CaMC. Calcium binding to CaMN changes the mode of interaction of CaM with the target protein.

The set of target proteins that binds the semi-open cleft of apo CaM domains more favorably than the open cleft of calcium-saturated CaM domains include neuromodulin [8183] and neurogranin [82, 84] as well as some myosins. Their CaMBDs are IQ motifs, a subclass of BAA motifs, and some release CaM when calcium binds CaM.[85] Sequence variations flanking the IQ motif influence whether only CaMC, or both domains of CaM, interact with the IQ motif.[27, 8688]

For the well conserved IQ motifs in human voltage-gated sodium channels, NMR and crystallographic studies of CaM binding to NaV1.2, 1.5, and 1.6 have shown that apo CaMC anchored CaM on the IQ motif, while apo CaMN adopted variable positions and made few contacts with the channel fragment. This was consistent with thermodynamic data that showed undetectable or very weak binding to these IQ motifs by CaMN. Through linkage or direct thermodynamic measurements, these studies also showed that the NaV IQ motifs have a higher affinity for apo CaM than (Ca2+)4-CaM.[15, 16, 22, 33, 41] These studies of NaV-CaM interactions contrast with the behavior of IQ motifs in voltage-dependent calcium channels such as CaV1.2, which interacts with both domains of CaM and has a higher affinity for (Ca2+)4-CaM than apo CaM [8991], despite the similarities between the calcium and sodium channels.[92]

4.2 Apo and (Ca2+)4-CaM binding to NaV1.2 IQ

The IQ-motif of NaV1.2 binds both apo CaM and (Ca2+)4-CaM with high affinity (Table 1). In our earlier studies of CaM binding to a fluoresceinated peptide of the NaV1.2 IQ motif, the stoichiometric conditions required for signal-to-noise ratios that yielded reproducible measurements allowed us only to put limits on the dissociation constants because the Kd was similar to, or much lower than, the concentration of peptide being titrated. This challenge of estimating dissociation constants plagues many biophysical methods such as quantitative mobility shift assays and isothermal titration calorimetry where it can be impossible to detect a concentration of macromolecule that is an order of magnitude lower than the value of the Kd, leading researchers to conduct studies under stoichiometric conditions.

To address the low spectral intensity inherent in our prior studies using fluorescence anisotropy, here we employed the biosensor method of Black and Persechini.[38] The high steady-state intensity of YFP-CFP FRET allowed titrations to be conducted biosensor concentrations as low as 0.3 nM biosensor, a factor of 10 lower than the most favorable observed Kd. Although it would have been preferable to use an even lower concentration of biosensor, in the nonlinear least squares analysis to determine free energies, the difference between free and total [CaM] was accounted for by using a function that calculated free [CaM] from the total [CaM] and the degree of saturation (see Methods).

Simulations of saturation of the NaV1.2 IQ motif based on the AG values reported in Table 1 for apo (dashed) and calcium-saturated (solid) CaM are shown in Fig. 7A. The 30-fold separation in midpoints indicates that the CaM-IQ interface must be different, but a small conformational change that resulted in the loss of a few hydrogen bonds might account for this weakening of affinity for (Ca2+)4-CaM. Instead, our solution structures 2KXW and 2M5E (insets in Fig. 7A) showed that calcium triggers an ~180° rotation of CaMC around I1912 in the IQ motif, causing many changes at the binding interface between CaMC and residues in the IQ motif. The final conformation is similar to that of CaMC in (Ca2+)4-CaM bound to peptides from a variety of targets such as CaV2.2 (3DVJ [27]), CaMKl (1MXE [93]), and NaV1.5 III–IV linker (4DJC [40]).

To infer the biological occupancy of an IQ motif by one of the endstates of CaM (with 0 or 4 calcium ions bound), it is essential to know both the equilibrium constants for binding the IQ motif and the local cellular concentration of free CaM. In cells, the level of CaM will depend on the expression levels of target proteins and their post-translational modifications, as well as the affinity of CaM for each target.[32, 83, 9497] The affinity is also a function of the degree of calcium occupancy of CaM. To infer binding occupancy of the NaV1.2 IQ motif, we estimated free [CaM] as 50–100 nM.[97, 98] Fig. 7A shows that this would correspond to 94–97% occupancy by apo CaM, and 37–54% occupancy by (Ca2+)4-CaM. This predicts that, regardless of the level of free calcium, CaM would be constitutively bound to the NaV1.2 IQ motif. This prediction agrees with proteomics studies of NaV1.2-associated proteins in neuronal cells, which found the level of enrichment of CaM in NaV1.2 pull-downs to be similar to that of other constitutively interacting proteins such as the NaV β2 subunit and FGF12 (FHF1).[11]

4.3 States of CaM-IQ: 0, 2, or 4 Ca2+

A cell experiences increases in calcium concentration due to release of calcium from internal stores or influx from the extracellular space. A calcium titration in a cuvette or NMR tube mimics this transition. Titrations of CaM-IQp monitored by NMR showed that calcium bound to CaM with a stoichiometry of 4 ions to 1 CaM.[16] However, the calcium-binding affinity of sites III and IV (CaMC) was diminished, while the affinity of sites I and II (CaMN) was similar to those sites in free CaM.[16] Thus, the CaM domains behave differently and may bind calcium sequentially but the CaM-IQp complex can attain full saturation with 4 calcium ions bound.

Crystallographic structures of an intermediate state having half-saturated CaM bound to the IQ motifs of NaV1.2 (4JPZ) or NaV1.5 (4JQ0) provide insight into the transition during a calcium influx. Both structures have calcium bound to sites I and II in CaMN, while sites III and IV are not saturated. The 4-helix bundles of CaMC in 4JPZ and 4JQ0 retain the semi-open tertiary conformation and orientation characteristic of apo CaM bound to the IQ motifs of NaV1.2 (2KXW) and NaV1.5 (2L53, 4DCK, and 4OVN) shown in Fig. 2. These are similar to the groundbreaking structure (1G4Y, Fig. 2E) of CaM bound to a large fragment of the SK channel [20] in which CaMC was apo, while CaMN was calcium-saturated.

The full ensemble of populated states of NaV1.2 changes in response to the abundance of both CaM and calcium. In this study, we did not directly measure calcium binding to the complex containing CaM bound to the NaV1.2 IQ motif. However, we can simulate this behavior based on the difference in free energy (ΔΔG) of binding apo and calcium-saturated CaM to the NaV1.2 IQ motif (1.93 kcal/mol in Table 1). Applying the principle of thermodynamic linkage, this value of ΔΔG must be the same as the difference between the AG of calcium binding to free CaM and to the apo CaM-NaV1.2 complex. The simulation in Fig. 7B represents predicted occupancy of the NaV1.2 IQ motif by (Ca2+)4-CaM.

Because the Kd for (Ca2+)4-CaM binding to the IQ motif is ~30-fold higher (weaker) than that of apo CaM, the preferred binding partner for NaV1.2 at resting calcium conditions is apo CaM, which is depicted schematically in the blue vertical bar. The green bar signifies high calcium; at 10 03BCM free calcium, occupancy of the IQ motif by (Ca2+)4-CaM would be ~68% or two-thirds. While 10 03BCM free calcium does not persist in cells for long periods of time, and does not occur uniformly throughout the cell, the level of calcium may rise to this level transiently and locally in nanodomains.[99]

4.4 Rotation in Place

A schematic diagram depicting the transition from 0 to 2 to 4 calcium ions bound to CaM when associated with the NaV1.2 IQ motif is shown in Fig. 7C. In this model, calcium binding to sites I and II affects only CaMN, while calcium binding to sites III and IV triggers reversal of CaMC orientation on the IQ motif. Our NMR data for apo and calcium-saturated CaM bound to Nav1.2IQp showed that backbone resonances for residues in CaMN in the CaM-IQP complex and free CaMN were essentially identical (Supp. Fig. S2) indicating that CaMN does not bind to NaV1.2IQP. This independence of CaMN was also seen in the NMR structure (2L53, Fig. 1E) of apo CaM bound to NaV1.5. Thus, in Fig. 7C, CaMN is represented as a free domain, without a preferred interface on NaV1.2.

However, apo CaMN might adopt positions similar to the two distinct, and well-separated locations seen in the crystallographic structures for apo CaMN bound to NaV1.5 (4OVN or 4DCK, Fig. 1E), while calcium-saturated CaMN may adopt a position similar to those seen in structures 4JPZ or 4JQ0. The reversal of CaMC occurs with or without CaMN. Future studies in solution of CaM binding to a longer fragment of NaV1.2 will be required to address to what extent the positions of CaMN observed in crystallographic studies represent energetically preferred orientations that are persistent in cells.

This simple 3-state model provides CaM with a vocabulary of at least three words for allosteric regulation of NaV1.2. The larger question is how these might align with three functional states of the channel broadly defined as closed, open and inactivated. However, each of those functional states is much more complex than a single protein conformation, and the physiological complexes include the NaV β subunits (with a single transmembrane helix), as well as auxiliary intracellular proteins such as intracellular FGFs (FHFs), ankyrins, CaMKII, synaptotagmin, and casein kinase II β which are also thought to modulate the NaV α subunit.[100, 101] Although our data cannot directly address the correlation between energetically accessible structures and physiological regulation of NaV1.2, they provide a striking new insight into an unexpected structural response to calcium binding, and suggest that it will be worth determining to what degree the calcium-saturated state observed in 2M5E is populated in cells.

4.5 Structural Correlates of 3-State Switch

CaM-target structures representing the 3 states depicted schematically in Fig. 7C are shown in Fig. 7D. They were aligned using PyMOL (Schrödinger, LLC) according to the positions of residues corresponding to residues 1909–1915 of NaV1.2 in 2M5E, where I1912 (cyan) and Q1913 (green) in NaV1.2 are shown as ball-and-stick, and residues at positions corresponding to those IQ residues are represented the same way in the other structures. To highlight the orientation of CaM, CaMC helices E and H are gray, while F (red) and G (orange) draw attention to the reversal of CaMC relative to I1912 and Q1913. Apo CaM bound to the NaV1.5 IQ motif (4OVN) and to an IQ motif of myosin V (2IX7) show that the eponymous Q side chain is oriented towards the FG-linker in semi-open apo CaMC as it is in 2KXW. The intermediate state of CaM (having two calcium ions bound to CaMN) associated with the NaV1.2 IQ motif (4JPZ) retains the same orientation of CaMC relative to the IQ motif as the apo form of CaM.

A recent report [102] of a calcium-triggered reversal of CaMC on IQ motif 1 of myosin V (4ZLK), shown in the right panel of Fig. 7D, is similar in orientation to 2M5E. In both structures, calcium binding to CaMC triggers opening and reversal of the 4-helix bundle, with the side chain of Q oriented towards the linker between CaMN and CaMC. The observed polarity of Nav1.2IQp relative to CaMC is consistent with most other complexes of calcium-saturated CaM bound to peptides [32] representing CaMBDs from other ion channels, receptors and enzymes such as CaMKII (1CDM) shown in Fig. 2A (termini of the CaMBD peptide (green) are indicated in blue/N and red/C).

4.6 Reversal by Release and Recapture

How CaM could regulate NaV channels while remaining very tightly bound to “one” site within the C-terminal tail has been a long-standing conundrum. The thermodynamic and structural studies presented here demonstrate that calcium binding reverses the polarity of CaM bound to the NaV1.2 IQ motif because CaMC recognizes nested, anti-parallel sites within a single helix in a calcium-dependent manner. This unexpected re-orientation of CaM illustrates the exquisite power of calcium-induced helical reorientations to result in large re-orientation of CaM to yield regulatory effects. Although studies here only addressed interactions of CaM with the IQ motif of NaV1.2, they suggest several hypotheses about the regulatory action of CaM on the channel as a whole. Because calcium binding to CaMC triggers it to rotate in the saddle of the hydrophobic face of the IQ motif, this might assist in modulating NaV1.2 by moving CaMN into an alternative position that could reach other sequences within NaV1.2 or its auxiliary regulatory proteins. Based on studies of CaM-NaV interactions, Young and Caldwell postulated that NaV1.4 might contain an NLBD – a non-IQ domain that binds CaMN.[101]

The observed calcium-induced reversal of CaM also provides a plausible mechanism for transient release of CaM from the NaV1.2 IQ motif, despite its very high affinity for both apo and calcium-saturated CaM. This would be analogous to IQCG (IQ motif-containing G protein) in which calcium causes release of CaMC but promotes binding of (Ca2+)2-CaMN.[103] Although none of the experiments reported here directly address the kinetic aspects of binding or release of CaM from the channel or the IQ motif, Fig. 7E illustrates some plausible models based on known CaM-target interactions. One study concluded that dissociated (Ca2+)4-CaM acts on NaV1.5 indirectly by binding to CaMKII which then phosphorylates the channel.[101] Other studies of NaV1.5 have proposed that high levels of calcium may cause CaM to release from the IQ motif and relocate at least one of its domains to the DIII-DIV linker, which has micromolar affinity for CaMC.[39, 41] For NaV1.5, (Ca2+)4-CaM binds at the DIII-DIV linker (“inactivation gate”) through CaMC as shown in the crystallographic structure 4DJC, which has the same polarity of CaMC-peptide interaction observed in many other calcium-saturated CaM complexes such as 1CDM (Fig. 2A). Given that the structure of the CTD of NaV1.5 (4OVN, 4DCK, 4JQ0) is highly similar to the CTD of NaV1.2 (4JPZ), release of (Ca2+)4-CaM from the IQ motif might result in one or both CaM domains binding the NaV1.2 DIII-DIV linker, post-IQ or other non-IQ CaM-binding sites within the CTD. Precedents exist for CaM binding the N-terminal tail of CaV channels [44, 45]; NaVs may have a similar site. (Ca2+)4-CaM may bind to the NaV β subunits or other auxiliary proteins such as those depicted in models of NaV1.5.[104]

Re-association or relocation of (Ca2+)4-CaM might occur as it does for the STRA6 receptor for retinol uptake in which the two domains of CaM bind and bridge distinct sequences in the receptor that are not found in a continuous helix.[105] This was also observed for CaM binding to non-inactivating voltage-dependent KCNQ potassium channels.[106] Another plausible arrangement might follow the example of CaM binding to the SK Channel (1G4Y[20]) wherein a second CaM molecule may be recruited to NaV1.2 when calcium levels rise. Such a mechanism would share features with the calcium-induced remodeling of myosin V, where one CaM molecule has been observed to bind two IQ motifs simultaneously.[85]

4.7 CaM-regulated Channels across Phyla

By studying Paramecium — a “unicellular neuron” — Kung and colleagues first discovered that CaM was a constitutive subunit of some calcium-regulated sodium and potassium channels, and that defects in each domain of CaM compromise a unique set of physiological responses to external stimuli.[107110] Under-reactive mutants of this excitable cell deviated from wild-type CaM at positions in CaMN that were primarily outside of the calcium-binding sites, while over-reactive mutants had changes in CaMC that were primarily localized within sites III and IV. Similar domain-specific results were later observed in mutagenic studies in yeast.[111] Exome analyses of humans have identified calmodulinopathies that are also specific to one domain of CaM.[112115] Some of the identified positions, such as D95 in site III, have been found to be deleterious in both Paramecium and human studies, suggesting that their mechanisms of action are conserved.

Given that the sequences of Paramecium and human CaM are 89% identical, and 96% similar, and that all of the mutants identified by Kung were at positions that are identical in the human and Paramecium wild-type CaM sequences, it was predictable that the initial discovery of CaM-regulated ion channels would generalize to several families of calcium-regulated ion channels in all eukaryotes.[92, 116119] For many of these channels, CaM is associated under resting (low calcium) conditions as it is with NaV1.2. CaM may activate, inhibit, or do both, depending on its fractional saturation by calcium. For example, studies of the ryanodine receptor RyR1 [120122] showed that apo CaM activated it, while calcium-saturated CaM inhibited it, showing that switching could be promoted by the addition or dissociation of calcium.

4.8 Power of Protein Ensembles

Prior to its recognition as a regulator of ion channels, studies of CaM revolved around its regulation of soluble enzymes including kinases, cyclases, phosphodiesterases, and phosphatases which are activated by (Ca2+)4-CaM.[123126] The two domains or lobes of CaM were viewed primarily as resulting from duplication of the gene for a primordial 4-helix bundle with metal-binding loops.[127130] Having two sets of paired EF-hands could be an evolutionary advantage for increasing avidity for targets in signaling pathways and enlarging the recognition site, thereby amplifying the response to an influx of calcium from the external medium, or internal stores. As membrane-spanning receptors and channels came to be recognized as non-enzymatic targets of CaM, some of them, such as NMDA [79, 131, 132], RyR1 [133135] and CaV1.2 [91], were found to bind apo CaM under resting conditions, but have a higher affinity for calcium-saturated CaM than apo CaM. In that respect, they were similar to kinases such as CaMKII.

Clearly, calcium-saturated CaM is a biologically significant state of CaM. In yeast, Davis and colleagues also demonstrated that apo CaM is essential to viability.[136138] There are 16 states total: 14 possible intermediates with 1, 2 or 3 calcium ions bound. Structures of (Ca2+)2− CaM bound to a fragment of the SK channel (1G4Y), and the IQ motif region of NaV1.2 (4JPZ) and NaV1.5 (4JQ0) are examples of an intermediate partially ligated CaM. Half-saturated CaM has also been observed in reverse – with apo CaMN and calcium-saturated CaMC bound to the anthrax edema factor.[139] It is even possible that CaM with one or three calcium ions bound may also have a significant role in some physiological processes. CaBP1 (Calcium Binding Protein 1) binds only 3 Ca2+, suggesting that there are some neuronal targets that prefer this stoichiometry.[140] Wild-type CaM with 4 functional sites is unlikely to adopt states with only one site filled in a domain because of the high degree of cooperativity between calcium-binding sites in the same domains. However, interactions between binding-site residues in CaM and side chains of residues in a target protein could alter the energetic balance to tip it in favor of an unoccupied site as proposed in the analysis of 4DCK.[22]

In neurons, sodium channels with multiple binding sites for auxiliary proteins integrate input from a panoply of modulators including CaM. In the structure reported here (2M5E), we showed that calcium saturation of CaM bound to the IQ motif of NaV1.2 triggers a half-turn rotation of CaMC around I1912 which has structural consequences for the geometry of CaMN as well as CaMC. The physiological consequences must be explored further. Having determined what CaM does on an isolated IQ motif, it remains to be seen whether CaM reverses polarity in solution on the full CTD of NaV1.2, and how the binding of intracellular FGFs (FHFs) may influence CaM-IQ interactions. Just as regulatory pathways in genetic circuits quantitatively integrate the effects of multiple proteins binding to DNA, and interacting both directly and allosterically with each other[141], neurons can take advantage of the full ensemble of energetically accessible states for auxiliary proteins binding to channels, and make use of ligand-triggered changes in their flexible structures.

Supplementary Material

2

Supplemental Figure S1

Comparison of Metal-Chelating Residue Positions and Difference Maps of CaMC

A and B. CaM C-domain residues 76–148 (CaMC) in 8 structures were aligned using CaM residues 95–100 in Site III (93–104). CaMC is gray, Ca2+ ions are yellow spheres unless otherwise noted, and Mg2+ ions are gray. Residues are labeled according to the sequence of mammalian CaM.* For each labeled residue, the average distance (A) between the metal ion (Ca2+ or Mg2+) and the coordinating oxygen is given. For clarity, only site III is labeled, but the geometry of ion-chelation in site IV is equivalent.

A. Structures of Semi-Open CaMC - apo (calcium-depleted) or ion-bound CaMC (Mg2+)4-CaM bound to the IQ motif of NaV1.2(4DCK) and NaV1.5 (4OVN), and the IQ motif of Myosin VI (3GN4 chain H), and (Ca2+)4-CaM bound to the IQ motif of NaV1.2 (4JPZ) and Nav1.5 (4JQ0). Mg2+ ions are gray spheres and Ca2+ in 4JPZ and 4JQ0 are magenta spheres. Side chains of metal-chelating residues in site III are highlighted with sticks: 4DCK - blue, 4OVN – cyan, 4JPZ - magenta, 4JQ0 - pink, and 3GN4 chain H - light blue.

B. Structures of Open CaMC (Ca2+)4-CaM alone (1CLL), and (Ca2+)4-CaM bound to a CaMKII peptide (1CDM) or a Myosin VI “Insert 2” (3GN4 Chain B) Side chains of Ca2+-chelating residues in site III are highlighted with sticks: 1CLL – orange, 1CDM – red, and 3GN4 chain B – salmon. Figures made with PyMOL (Schrödinger, LLC).

C. Calcium Fo-Fc Electron Density Maps The Fo-Fc density maps for two Ca2+ ions (yellow) located in sites III and IV in each calmodulin chain were contoured at +3 (green mesh, insufficient density) and −3 (red mesh, excess density) sigma to a radius of 5 Å from the center of each ion in calmodulin alone (1CLL), or in calmodulin bound to the IQ motif of NaV1.2 in 4JPZ [Chain C, Chain I]. Figures made with PyMOL (Schrödinger, LLC). PDB files and electron density maps were obtained from PDB_REDO (http://www.cmbi.ru.nl/pdb_redo/) Calmodulin residues 76–148 (gray) were aligned based on residues 117–128 (i.e., helix “G”).

*Mammalian CaM has 148 amino acids. Standard numbering begins with A1, and ends with K148. The Met listed as residue 1 in the UNIPROT database is not found in mammalian CaM, or in the bacterially expressed sequence of mammalian CaM.

Supplemental Figure S2

HSQC spectra of 13C,15N-(Ca2+)4-CaM(1–148)-Nav1.2IQp vs. 13C,15N-(Ca2+)2-CaM(1–75)

HSQC overlay of full-length CaM saturated with calcium and Nav1.2IQ [(Ca2+)4CaM-Nav1.2IQp; black] with free CaMN [(Ca2+)2-CaM(1–75); blue]. Peaks that are nearly identical are circled in green. The high level of agreement between these two spectra indicates that few residues in the N-domain of CaM are perturbed by CaM binding to the IQ motif and the C-domain makes most contacts in the CaM-Nav1.2IQp interface.

Supplemental Figure S3

Ramachandran Plot

PROCHECK-NMR analysis [70] of the minimized average structure in 2M5E.

Supplemental Figure S4

NOESY Overlay of Nav1.2IQp Bound to CaM ± Ca2+

Overlays comparing two regions of the doubly 12C,14N filtered NOESY spectra of NaV1.2IQp bound to apo (green) and Ca2+-saturated (black) CaM. Crosspeaks are shown (a) between upfield and downfield aliphatic protons and (b) between amide plus aromatic protons and aliphatic protons. The significant differences suggest that NaV1.2IQp is in a very different chemical environment when bound to apo CaM vs. calcium-saturated CaM.

Highlights.

  1. NaV1.2 IQ motif binds apo and (Ca2+)4-CaM tightly: how can Ca2+ binding trigger change?

  2. Direct measure of equilibrium Kd values: apo CaM (~3 nM) vs. (Ca2+)4-CaM (~85 nM).

  3. Solution structure of (Ca2+)2-CaMC bound to NaV1.2 IQ has “open” EF-hand conformation.

  4. 2KXW and 2M5E are first NMR structures of apo and Ca2+-CaMC bound to same channel.

  5. CaM binds nested, anti-parallel sites in NaV1.2 IQ; Ca2+ binding to CaMC reverses its orientation.

Acknowledgments

We thank A. Persechini and DJ Black for sharing a published vector used as the parent for the biosensor overexpression plasmid made in this study, and Sean A. Klein for critical input on experimental approaches.

These studies were supported by Iowa Center for Research by Undergraduates Fellowships to L.H., Z.L, and D.C.M,, a graduate fellowship for M.S.M. from the University of Iowa Center for Biocatalysis and Bioprocessing Training Grant in Biotechnology NIH T32 GM08365, a Biochemistry Summer Undergraduate Fellowship for B.C.W., the Roy J. Carver Charitable Trust Grant 01–224, and a grant to M.A.S. from the National Institutes of Health (R01 GM57001). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Abbreviations

BAA

Basic, Amphipathic Alpha Helix

CaM

Calmodulin

CaMBD

Calmodulin-Binding Domain

CaMC

C-domain of CaM

CaMN

N-domain of CaM

CFP

Cyan Fluorescent Protein

CTD

C-terminal Domain

EGTA

Ethylene Glycol Bis(aminoethylether)-N,N,N’,N’-tetra-acetic acid

FRET

Förster/Fluorescence Resonance Energy Transfer

HEPES

N-(2-hydroxy-ethyl)piperazine-N′-2-ethanesulfonic acid

HSQC

Heteronuclear Single Quantum Coherence

Nav1.2IQp

IQ motif peptide (residues 1901 to 1927) of rat NaV1.2

NaV

Voltage-gated Sodium Channel

NMR

Nuclear Magnetic Resonance

NOE

Nuclear Overhauser Effect

NTA

Nitrilo-triacetic acid

RMSD

Root-Mean-Squared Deviation

YFP

Yellow Fluorescent Protein

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Accession Numbers

The coordinates for the average minimized structure and 20 lowest energy models were deposited with the PDB code of 2M5E; there were preliminary conference reports of this structure.[142, 143] NMR assignment data (accession number 19050) were deposited with the BioMagResBank (BMRB http://www.bmrb.wisc.edu). Assignments are also contained in Supplemental Tables S2 and S3.

References

  • 1.Catterall WA. From Ionic Currents to Molecular Mechanisms: The Structure and Function of Voltage-Gated Sodium Channels. Neuron. 2000;26:13–25. doi: 10.1016/s0896-6273(00)81133-2. [DOI] [PubMed] [Google Scholar]
  • 2.Schaller KL, Caldwell JH. Expression and distribution of voltage-gated sodium channels in the cerebellum. Cerebellum. 2003;2:2–9. doi: 10.1080/14734220309424. [DOI] [PubMed] [Google Scholar]
  • 3.Meisler MH, Kearney J, Ottman R, Escayg A. Identification of epilepsy genes in human and mouse. Annual Review of Genetics. 2001;35:567–588. doi: 10.1146/annurev.genet.35.102401.091142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Rook MB, Bezzina Alshinawi C, Groenewegen WA, van Gelder IC, van Ginneken AC, Jongsma HJ, Mannens MM, Wilde AA. Human SCN5A gene mutations alter cardiac sodium channel kinetics and are associated with the Brugada syndrome. Cardiovascular research. 1999;44:507–517. doi: 10.1016/s0008-6363(99)00350-8. [DOI] [PubMed] [Google Scholar]
  • 5.Weiss LA, Escayg A, Kearney JA, Trudeau M, MacDonald BT, Mori M, Reichert J, Buxbaum JD, Meisler MH. Sodium channels SCN1A, SCN2A and SCN3A in familial autism. Molecular Psychiatry. 2003;8:186–194. doi: 10.1038/sj.mp.4001241. [DOI] [PubMed] [Google Scholar]
  • 6.Meisler MH, Kearney JA. Sodium channel mutations in epilepsy and other neurological disorders. Journal of Clinical Investigation. 2005;115:2010–2017. doi: 10.1172/JCI25466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Kim J, Ghosh S, Liu H, Tateyama M, Kass RS, Pitt GS. Calmodulin mediates Ca2+ sensitivity of sodium channels. Journal of Biological Chemistry. 2004;279:45004–45012. doi: 10.1074/jbc.M407286200. [DOI] [PubMed] [Google Scholar]
  • 8.Cheung WY. Calmodulin Plays a Pivotal Role in Cellular Regulation. Science. 1980;207:19–27. doi: 10.1126/science.6243188. [DOI] [PubMed] [Google Scholar]
  • 9.Ben-Johny M, Yang PS, Niu J, Yang W, Joshi-Mukherjee R, Yue DT. Conservation of Ca2+/Calmodulin Regulation across Na and Ca2+ Channels. Cell. 2014;157:1657–1670. doi: 10.1016/j.cell.2014.04.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Yu FH, Catterall WA. Overview of the voltage-gated sodium channel family. Genome Biology. 2003;4:207.201–207.207. doi: 10.1186/gb-2003-4-3-207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Wildburger NC, Ali SR, Hsu WC, Shavkunov AS, Nenov MN, Lichti CF, LeDuc RD, Mostovenko E, Panova-Elektronova NI, Emmett MR, Nilsson CL, Laezza F. Quantitative proteomics reveals protein-protein interactions with fibroblast growth factor 12 as a component of the voltage-gated sodium channel 1.2 (Nav1.2) macromolecular complex in Mammalian brain. Molecular and Cellular Proteomics. 2015;14:1288–1300. doi: 10.1074/mcp.M114.040055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Miloushev VZ, Levine JA, Arbing MA, Hunt JF, Pitt GS, Palmer AG. Solution structure of the NaV1.2 C-terminal EF-hand domain. Journal of Biological Chemistry. 2009;284:6446–6454. doi: 10.1074/jbc.M807401200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Rhoads AR, Friedberg F. Sequence motifs for calmodulin recognition. The Federation of American Societies for Experimental Biology Journal. 1997;11:331–340. doi: 10.1096/fasebj.11.5.9141499. [DOI] [PubMed] [Google Scholar]
  • 14.Mori M, Konno T, Ozawa T, Murata M, Imoto K, Nagayama K. Novel Interaction of the Voltage-Dependent Sodium Channel (VDSC) with Calmodulin: Does VDSC Acquire Calmodulin-Mediated Ca2+-Sensitivity? Biochemistry. 2000;39:1316–1323. doi: 10.1021/bi9912600. [DOI] [PubMed] [Google Scholar]
  • 15.Theoharis NT, Sorensen BR, Theisen-Toupal J, Shea MA. The Neuronal Voltage-Dependent Sodium Channel Type II IQ Motif Lowers the Calcium Affinity of the C-Domain of Calmodulin. Biochemistry. 2008;47:112–123. doi: 10.1021/bi7013129. [DOI] [PubMed] [Google Scholar]
  • 16.Feldkamp MD, Yu L, Shea MA. Structural and Energetic Determinants of Apo Calmodulin Binding to the IQ Motif of the Nav1.2 Voltage-Dependent Sodium Channel. Structure. 2011;19:733–747. doi: 10.1016/j.str.2011.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Cormier JW, Rivolta I, Tateyama M, Yang AS, Kass RS. Secondary Structure of the Human Cardiac Na+ Channel C Terminus. Journal of Biological Chemistry. 2002;277:9233–9241. doi: 10.1074/jbc.M110204200. [DOI] [PubMed] [Google Scholar]
  • 18.Mantegazza M, Yu FH, Catterall WA, Scheuer T. Role of the C-terminal domain in inactivation of brain and cardiac sodium channel. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:15348–15353. doi: 10.1073/pnas.211563298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Houdusse A, Gaucher JF, Krementsova E, Mui S, Trybus KM, Cohen C. Crystal structure of apo-calmodulin bound to the first two IQ motifs of myosin V reveals essential recognition features. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:19326–19331. doi: 10.1073/pnas.0609436103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Schumacher MA, Rivard AF, Bachinger HP, Adelman JP. Structure of the gating domain of a Ca2+-activated K+ channel complexed with Ca2+/calmodulin. Nature. 2001;410:1120–1124. doi: 10.1038/35074145. [DOI] [PubMed] [Google Scholar]
  • 21.Chagot B, Chazin WJ. Solution NMR Structure of Apo-Calmodulin in Complex with the IQ Motif of Human Cardiac Sodium Channel NaV1.5. Journal of molecular biology. 2011;406:106119. doi: 10.1016/j.jmb.2010.11.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Wang C, Chung BC, Yan H, Lee SY, Pitt GS. Crystal Structure of the Ternary Complex of a NaV C-Terminal Domain, a Fibroblast Growth Factor Homologous Factor, and Calmodulin. Structure. 2012 doi: 10.1016/j.str.2012.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Gabelli SB, Boto A, Kuhns VH, Bianchet MA, Farinelli F, Aripirala S, Yoder J, Jakoncic J, Tomaselli GF, Amzel LM. Regulation of the NaV1.5 cytoplasmic domain by calmodulin. Nature communications. 2014;5:5126. doi: 10.1038/ncomms6126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Reddy Chichili VP, Xiao Y, Seetharaman J, Cummins TR, Sivaraman J. Structural basis for the modulation of the neuronal voltage-gated sodium channel NaV1.6 by calmodulin. Scientific reports. 2013;3:2435. doi: 10.1038/srep02435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Strynadka NCJ, James MNG. Crystal Structures of the Helix-Loop-Helix Calcium-Binding Proteins. Annual Review of Biochemistry. 1989;58:951–998. doi: 10.1146/annurev.bi.58.070189.004511. [DOI] [PubMed] [Google Scholar]
  • 26.Crooks GE, Hon G, Chandonia JM, Brenner SE. WebLogo: a sequence logo generator. Genome Research. 2004;14:1188–1190. doi: 10.1101/gr.849004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Kim EY, Rumpf CH, Fujiwara Y, Cooley ES, Van Petegem F, Minor DL., Jr Structures of CaV2 Ca2+/CaM-IQ domain complexes reveal binding modes that underlie calcium-dependent inactivation and facilitation. Structure. 2008;16:1455–1467. doi: 10.1016/j.str.2008.07.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Cantrell AR, Catterall WA. Neuromodulation of Na+ Channels: An Unexpected Form of Cellular Plasticity, Nature Reviews. Neuroscience. 2001;2:397–407. doi: 10.1038/35077553. [DOI] [PubMed] [Google Scholar]
  • 29.Ikura M, Ames JB. Genetic polymorphism and protein conformational plasticity in the calmodulin superfamily: two ways to promote multifunctionality. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:1159–1164. doi: 10.1073/pnas.0508640103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Meador WE, Means AR, Quiocho FA. Modulation of calmodulin plasticity in molecular recognition on the basis of X-ray structures. Science. 1993;262:1718–1721. doi: 10.1126/science.8259515. [DOI] [PubMed] [Google Scholar]
  • 31.Homouz D, Sanabria H, Waxham MN, Cheung MS. Modulation of calmodulin plasticity by the effect of macromolecular crowding. Journal of molecular biology. 2009;391:933–943. doi: 10.1016/j.jmb.2009.06.073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Villarroel A, Taglialatela M, Bernardo-Seisdedos G, Alaimo A, Agirre J, Alberdi A, Gomis-Perez C, Soldovieri MV, Ambrosino P, Malo C, Areso P. The ever changing moods of calmodulin: how structural plasticity entails transductional adaptability. Journal of molecular biology. 2014;426:2717–2735. doi: 10.1016/j.jmb.2014.05.016. [DOI] [PubMed] [Google Scholar]
  • 33.Wang C, Chung BC, Yan H, Wang HG, Lee SY, Pitt GS. Structural analyses of Ca2+/CaM interaction with NaV channel C-termini reveal mechanisms of calcium-dependent regulation. Nature communications. 2014;5:4896. doi: 10.1038/ncomms5896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Chagot B, Potet F, Balser JR, Chazin WJ. Solution NMR structure of the C-terminal EF-hand domain of human cardiac sodium channel NaV1.5, Journal of. Biological Chemistry. 2009;284:6436–6445. doi: 10.1074/jbc.M807747200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Feldkamp MD, Yu L, Shea MA. Calmodulin Regulation of the Neuronal Voltage-Dependent Sodium Channel. Biophysical journal. 2010;98:310. [Google Scholar]
  • 36.Strynadka NCJ, James MNG. Model for the Interaction of Amphiphilic Helices With Troponin C and Calmodulin, Proteins: Structure. Function, and Genetics. 1990;7:234–248. doi: 10.1002/prot.340070305. [DOI] [PubMed] [Google Scholar]
  • 37.LaPorte DC, Wierman BM, Storm DR. Calcium-induced exposure of a hydrophobic surface on calmodulin. Biochemistry. 1980;19:3814–3819. doi: 10.1021/bi00557a025. [DOI] [PubMed] [Google Scholar]
  • 38.Black DJ, Leonard J, Persechini A. Biphasic Ca2+-dependent switching in a calmodulin-IQ domain complex. Biochemistry. 2006;45:6987–6995. doi: 10.1021/bi052533w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Sarhan MF, Van Petegem F, Ahern CA. A double tyrosine motif in the cardiac sodium channel domain III-IV linker couples calcium-dependent calmodulin binding to inactivation gating. Journal of Biological Chemistry. 2009;284:33265–33274. doi: 10.1074/jbc.M109.052910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Sarhan MF, Tung CC, Van Petegem F, Ahern CA. Crystallographic basis for calcium regulation of sodium channels. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:3558–3563. doi: 10.1073/pnas.1114748109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Shah VN, Wingo TL, Weiss KL, Williams CK, Balser JR, Chazin WJ. Calcium-dependent regulation of the voltage-gated sodium channel hH1: intrinsic and extrinsic sensors use a common molecular switch. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:3592–3597. doi: 10.1073/pnas.0507397103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Lee A, Zhou H, Scheuer T, Catterall WA. Molecular determinants of Ca2+/calmodulin-dependent regulation of Cav2.1 channels. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:16059–16064. doi: 10.1073/pnas.2237000100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Lee A, Scheuer T, Catterall WA. Ca2+/calmodulin-dependent facilitation and inactivation of P/Q-type Ca2+ channels. Journal of Neuroscience. 2000;20:6830–6838. doi: 10.1523/JNEUROSCI.20-18-06830.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Dick IE, Tadross MR, Liang H, Tay LH, Yang W, Yue DT. A modular switch for spatial Ca2+ selectivity in the calmodulin regulation of CaV channels. Nature. 2008;451:830–834. doi: 10.1038/nature06529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Zhou H, Yu K, McCoy KL, Lee A. Molecular mechanism for divergent regulation of CaV1.2 Ca2+ channels by calmodulin and Ca2+-binding protein-1. Journal of Biological Chemistry. 2005;280:29612–29619. doi: 10.1074/jbc.M504167200. [DOI] [PubMed] [Google Scholar]
  • 46.Leal K, Mochida S, Scheuer T, Catterall WA. Fine-tuning synaptic plasticity by modulation of CaV2.1 channels with Ca2+ sensor proteins. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:17069–17074. doi: 10.1073/pnas.1215172109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Putkey JA, Slaughter GR, Means AR. Bacterial expression and characterization of proteins derived from the chicken calmodulin cDNA and a calmodulin processed gene. Journal of Biological Chemistry. 1985;260:4704–4712. [PubMed] [Google Scholar]
  • 48.Smith PK, Krohn RI, Hermanson GT, Mallia AK, Gartner FH, Provenzano MD, Fujimoto EK, Goeke NM, Olson BJ, Klenk DC. Measurement of Protein Using Bicinchoninic Acid. Analytical Biochemistry. 1985;150:76–85. doi: 10.1016/0003-2697(85)90442-7. [DOI] [PubMed] [Google Scholar]
  • 49.Beaven GH, Holiday ER. Ultraviolet absorption spectra of proteins and amino acids. Advances in Protein Chemistry. 1952;7:319–386. doi: 10.1016/s0065-3233(08)60022-4. [DOI] [PubMed] [Google Scholar]
  • 50.Romoser VA, Hinkle PM, Persechini A. Detection in living cells of Ca2+-dependent changes in the fluorescence emission of an indicator composed of two green fluorescent protein variants linked by a calmodulin-binding sequence. A new class of fluorescent indicators. Journal of Biological Chemistry. 1997;272:13270–13274. doi: 10.1074/jbc.272.20.13270. [DOI] [PubMed] [Google Scholar]
  • 51.Tsien RY. The green fluorescent protein. Annual Review of Biochemistry. 1998;67:509544. doi: 10.1146/annurev.biochem.67.1.509. [DOI] [PubMed] [Google Scholar]
  • 52.Johnson ML, Frasier SG. Nonlinear Least-Squares Analysis. Methods in Enzymology. 1985;117:301–342. [Google Scholar]
  • 53.Sorensen BR, Shea MA. Interactions between domains of apo calmodulin alter calcium binding and stability. Biochemistry. 1998;37:4244–4253. doi: 10.1021/bi9718200. [DOI] [PubMed] [Google Scholar]
  • 54.VanScyoc WS, Newman RA, Sorensen BR, Shea MA. Calcium Binding by Calmodulin Mutants Having Domain-Specific Effects on Regulation of Ion Channels. Biochemistry. 2006;45:14311–14324. doi: 10.1021/bi061134d. [DOI] [PubMed] [Google Scholar]
  • 55.Delaglio F, Grzesiek S, Vuister GW, Zhu G, Pfeifer J, Bax A. NMRPipe: a multidimensional spectral processing system based on UNIX pipes. Journal of Biomolecular NMR. 1995;6:277–293. doi: 10.1007/BF00197809. [DOI] [PubMed] [Google Scholar]
  • 56.Goddard TD, Kneller DG. SPARKY, University of California, San Francisco [Google Scholar]
  • 57.Vranken WF, Boucher W, Stevens TJ, Fogh RH, Pajon A, Llinas M, Ulrich EL, Markley JL, Ionides J, Laue ED. The CCPN data model for NMR spectroscopy: development of a software pipeline. Proteins. 2005;59:687–696. doi: 10.1002/prot.20449. [DOI] [PubMed] [Google Scholar]
  • 58.Bahrami A, Assadi AH, Markley JL, Eghbalnia HR. Probabilistic interaction network of evidence algorithm and its application to complete labeling of peak lists from protein NMR spectroscopy. PLoS computational biology. 2009;5:e1000307. doi: 10.1371/journal.pcbi.1000307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Gemmecker G, Olejniczak ET, Fesik SW. An Improved Method for Selectively Observing Protons Attached to C-12 in the Presence of H-1-C-13 Spin Pairs. Journal of magnetic resonance. 1992;96:199–204. [Google Scholar]
  • 60.Ikura M, Bax A. Isotope-Filtered 2D NMR of a Protein-Peptide Complex: Study of a Skeletal Muscle Myosin Light Chain Kinase Fragment Bound to Calmodulin. Journal of the American Chemistry Society. 1992;114:2433–2440. [Google Scholar]
  • 61.Otting G, Wuthrich K. Extended Heteronuclear Editing of 2d H-1-Nmr Spectra of Isotope-Labeled Proteins, Using the X(Omega-1, Omega-2) Double Half Filter. Journal of magnetic resonance. 1989;85:586–594. [Google Scholar]
  • 62.Burgering MJ, Boelens R, Caffrey M, Breg JN, Kaptein R. Observation of inter-subunit nuclear Overhauser effects in a dimeric protein. Application to the Arc repressor, FEBS Letters. 1993;330:105–109. doi: 10.1016/0014-5793(93)80929-o. [DOI] [PubMed] [Google Scholar]
  • 63.Cheung MS, Maguire ML, Stevens TJ, Broadhurst RW. DANGLE: A Bayesian inferential method for predicting protein backbone dihedral angles and secondary structure. Journal of magnetic resonance. 2010;202:223–233. doi: 10.1016/j.jmr.2009.11.008. [DOI] [PubMed] [Google Scholar]
  • 64.Shen Y, Delaglio F, Cornilescu G, Bax A. TALOS+: a hybrid method for predicting protein backbone torsion angles from NMR chemical shifts. Journal of Biomolecular NMR. 2009;44:213–223. doi: 10.1007/s10858-009-9333-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Rieping W, Habeck M, Bardiaux B, Bernard A, Malliavin TE, Nilges M. ARIA2: automated NOE assignment and data integration in NMR structure calculation. Bioinformatics. 2007;23:381–382. doi: 10.1093/bioinformatics/btl589. [DOI] [PubMed] [Google Scholar]
  • 66.Brunger AT, Adams PD, Clore GM, DeLano WL, Gros P, Grosse-Kunstleve RW, Jiang JS, Kuszewski J, Nilges M, Pannu NS, Read RJ, Rice LM, Simonson T, Warren GL. Crystallography & NMR system: A new software suite for macromolecular structure determination. Acta Crystallographics Section D Biological Crystallography. 1998;54:905–921. doi: 10.1107/s0907444998003254. [DOI] [PubMed] [Google Scholar]
  • 67.Schwieters CD, Kuszewski JJ, Tjandra N, Clore GM. The Xplor-NIH NMR molecular structure determination package. Journal of magnetic resonance. 2003;160:65–73. doi: 10.1016/s1090-7807(02)00014-9. [DOI] [PubMed] [Google Scholar]
  • 68.Schwieters CD, Kuszewski JJ, Clore GM. Using Xplor-NIH for NMR molecular structure determination. Progress in nuclear magnetic resonance spectroscopy. 2006;48:47–62. [Google Scholar]
  • 69.Wilson MA, Brunger AT. The 1.0 A Crystal Structure of Ca2+-bound Calmodulin: an Analysis of Disorder and Implications for Functionally Relevant Plasticity. Journal of Molecular Biology. 2000;301:1237–1256. doi: 10.1006/jmbi.2000.4029. [DOI] [PubMed] [Google Scholar]
  • 70.Laskowski RA, Rullmannn JA, MacArthur MW, Kaptein R, Thornton JM. AQUA and PROCHECK-NMR: programs for checking the quality of protein structures solved by NMR. Journal of Biomolecular NMR. 1996;8:477–486. doi: 10.1007/BF00228148. [DOI] [PubMed] [Google Scholar]
  • 71.Newman RA, Van Scyoc WS, Sorensen BR, Jaren OR, Shea MA. Interdomain cooperativity of calmodulin to melittin preferentially increases calcium affinity of sites I and II. Proteins: Structure, Function, and Bioinformatics. 2008;71:1792–1812. doi: 10.1002/prot.21861. [DOI] [PubMed] [Google Scholar]
  • 72.Coulomb CA. Premier mémoire sur I’électricité et le magnétisme Histoire de I’Académie Royale des Sciences, ( 1785:569–577. [Google Scholar]
  • 73.Sobolev V, Sorokine A, Prilusky J, Abola EE, Edelman M. Automated analysis of interatomic contacts in proteins. Bioinformatics. 1999;15:327–332. doi: 10.1093/bioinformatics/15.4.327. [DOI] [PubMed] [Google Scholar]
  • 74.Yamniuk AP, Vogel HJ. Calmodulin’s flexibility allows for promiscuity in its interactions with target proteins and peptides. Molecular Biotechnology. 2004;27:33–57. doi: 10.1385/MB:27:1:33. [DOI] [PubMed] [Google Scholar]
  • 75.Ataman ZA, Gakhar L, Sorensen BR, Hell JW, Shea MA. The NMDA Receptor NR1 C1 Region Bound to Calmodulin: Structural Insights into Functional Differences between Homologous Domains. Structure. 2007;15:1603–1617. doi: 10.1016/j.str.2007.10.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Ishida H, Vogel HJ. Protein-peptide interaction studies demonstrate the versatility of calmodulin target protein binding. Protein and Peptide Letter. 2006;13:455–465. doi: 10.2174/092986606776819600. [DOI] [PubMed] [Google Scholar]
  • 77.Peersen OB, Madsen TS, Falke JJ. Intermolecular tuning of calmodulin by target peptides and proteins: differential effects on Ca2+ binding and implications for kinase activation. Protein Science. 1997;6:794–807. doi: 10.1002/pro.5560060406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.O’Donnell SE, Yu L, Fowler CA, Shea MA. Recognition of beta-calcineurin by the domains of calmodulin: thermodynamic and structural evidence for distinct roles. Proteins. 2011;79:765–786. doi: 10.1002/prot.22917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Akyol Z, Bartos JA, Merrill MA, Faga LA, Jaren OR, Shea MA, Hell JW. Apo–Calmodulin Binds with its COOH-terminal Domain to the N-methyl-D-aspartate Receptor NR1 C0 Region. Journal of Biological Chemistry. 2004;279:2166–2175. doi: 10.1074/jbc.M302542200. [DOI] [PubMed] [Google Scholar]
  • 80.Sonnenburg WK, Seger D, Kwak KS, Huang J, Charbonneau H, Beavo JA. Identification of inhibitory and calmodulin-binding domains of the PDE1A1 and PDE1A2 calmodulin-stimulated cyclic nucleotide phosphodiesterases. Journal of Biological Chemistry. 1995;270:30989–31000. doi: 10.1074/jbc.270.52.30989. [DOI] [PubMed] [Google Scholar]
  • 81.Black DJ, LaMartina D, Persechini A. The IQ domains in neuromodulin and PEP19 represent two major functional classes. Biochemistry. 2009;48:11766–11772. doi: 10.1021/bi9014874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Gerendasy DD, Herron SR, Jennings PA, Sutcliffe JG. Calmodulin stabilizes an amphiphilic a-helix within RC3/neurogranin and GAP-43/neuromodulin only when Ca2+ is absent. Journal of Biological Chemistry. 1995;270:6741–6750. doi: 10.1074/jbc.270.12.6741. [DOI] [PubMed] [Google Scholar]
  • 83.Liu Y, Storm DR. Regulation of free calmodulin levels by neuromodulin: Neuron growth and regeneration. TIPS. 1990;11:107–111. doi: 10.1016/0165-6147(90)90195-e. [DOI] [PubMed] [Google Scholar]
  • 84.Gerendasy DD, Herron SR, Watson JB, Sutcliffe JG. Mutational and biophysical studies suggest RC3/neurogranin regulates calmodulin availability. Journal of Biological Chemistry. 1994;269:22420–22426. [PubMed] [Google Scholar]
  • 85.Martin SR, Bayley PM. Regulatory implications of a novel mode of interaction of calmodulin with a double IQ-motif target sequence from murine dilute myosin V. Protein Science. 2002;11:2909–2923. doi: 10.1110/ps.0210402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Black DJ, Persechini A. Variations at the semiconserved glycine in the IQ domain consensus sequence have a major impact on Ca2+-dependent switching in calmodulin-IQ domain complexes. Biochemistry. 2010;49:78–83. doi: 10.1021/bi901695p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Black DJ, Persechini A. In calmodulin-IQ domain complexes, the Ca2+-free and Ca2+-bound forms of the calmodulin l direct the N-lobe to different binding sites. Biochemistry. 2011;50:10061–10068. doi: 10.1021/bi201300v. [DOI] [PubMed] [Google Scholar]
  • 88.Putkey JA, Kleerekoper Q, Gaertner TR, WAxham MN. A New Role for IQ Motif Proteins in Regulating Calmodulin Function. Journal of Biological Chemistry. 2003;278:49667–49670. doi: 10.1074/jbc.C300372200. [DOI] [PubMed] [Google Scholar]
  • 89.Evans TIA, Hell J, Shea MA. Thermodynamic Linkage Between Calmodulin Domains Binding Calcium and Contiguous Sites in the C-Terminal Tail of CaV1.2. Biophysical Chemistry. 2011;159:172–187. doi: 10.1016/j.bpc.2011.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Findeisen F, Tolia A, Arant R, Young Kim E, Isacoff E, Minor DL., Jr Calmodulin overexpression does not alter CaV1.2 function or oligomerization state. Channels. 2011;5 doi: 10.4161/chan.5.4.16821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Peterson BZ, DeMaria CD, Yue DT. Calmodulin Is the Ca2+ Sensor for Ca2+-Dependent Inactivation of L-Type Calcium Channels. Neuron. 1999;22:549–558. doi: 10.1016/s0896-6273(00)80709-6. [DOI] [PubMed] [Google Scholar]
  • 92.Ben-Johny M, Dick IE, Yue D, Yang W, Issa JB, Sang L, Lee SR, Limpitikul WB, Niu J, Kang PW, Banerjee R, Babich JS, Namkung H, Li J, Zhang M, Yang PS, Yue DN. Towards a Unified Theory of Calmodulin Regulation (Calmodulation) of Voltage-Gated Calcium and Sodium Channels. Current Molecular Pharmacology. 2014:1–18. doi: 10.2174/1874467208666150507110359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Clapperton JA, Martin SR, Smerdon SJ, Gamblin SJ, Bayley PM. Structure of the Complex of Calmodulin with the Target Sequences of Calmodulin-Dependent Protein Kinase I: Studies of the Kinase Activation Mechanism. Biochemistry. 2002;41:14669–14679. doi: 10.1021/bi026660t. [DOI] [PubMed] [Google Scholar]
  • 94.Tran QK, Black DJ, Persechini A. Dominant affectors in the calmodulin network shape the time courses of target responses in the cell. Cell Calcium. 2005;37:541–553. doi: 10.1016/j.ceca.2005.02.001. [DOI] [PubMed] [Google Scholar]
  • 95.Tran QK, Leonard J, Black DJ. A. Persechini, Phosphorylation within an autoinhibitory domain in endothelial nitric oxide synthase reduces the Ca2+ concentrations required for calmodulin to bind and activate the enzyme. Biochemistry. 2008;47:7557–7566. doi: 10.1021/bi8003186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Black DJ, Tran QK, Persechini A. Monitoring the total available calmodulin concentration in intact cells over the physiological range in free Ca2+ Cell Calcium. 2004;35:415–425. doi: 10.1016/j.ceca.2003.10.005. [DOI] [PubMed] [Google Scholar]
  • 97.Maier LS, Ziolo MT, Bossuyt J, Persechini A, Mestril R, Bers DM. Dynamic changes in free Ca-calmodulin levels in adult cardiac myocytes. Journal of Molecular and Cellular Cardiology. 2006;41:451–458. doi: 10.1016/j.yjmcc.2006.04.020. [DOI] [PubMed] [Google Scholar]
  • 98.Mori MX, Erickson MG, Yue DT. Functional Stoichiometry and Local Enrichment of Calmodulin Interacting with Ca2+ Channels. Science. 2004;304:432–435. doi: 10.1126/science.1093490. [DOI] [PubMed] [Google Scholar]
  • 99.Li R, Leblanc J, He K, Liu XJ. Spindle function in Xenopus oocytes involves possible nanodomain calcium signaling. Molecular Biology of the Cell. 2016;27:3273–3283. doi: 10.1091/mbc.E16-05-0338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Bosch MK, Nerbonne JM, Townsend RR, Miyazaki H, Nukina N, Ornitz DM, Marionneau C. Proteomic analysis of native cerebellar iFGF14 complexes. Channels. 2016;10:297–312. doi: 10.1080/19336950.2016.1153203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Young KA, Caldwell JH. Modulation of skeletal and cardiac voltage-gated sodium channels by calmodulin. Journal of Physiology. 2005;565:349–370. doi: 10.1113/jphysiol.2004.081422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Shen M, Zhang N, Zheng S, Zhang WB, Zhang HM, Lu Z, Su QP, Sun Y, Ye K, Li XD. Calmodulin in complex with the first IQ motif of myosin-5a functions as an intact calcium sensor. Proceedings of the National Academy of Sciences of the United States of America. 2016;113:E5812–E5820. doi: 10.1073/pnas.1607702113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Chen LT, Liang WX, Chen S, Li RK, Tan JL, Xu PF, Luo LF, Wang L, Yu SH, Meng G, Li KK, Liu TX, Chen Z, Chen SJ. Functional and molecular features of the calmodulin-interacting protein IQCG required for haematopoiesis in zebrafish. Nature Communications. 5:3811. doi: 10.1038/ncomms4811. [DOI] [PubMed] [Google Scholar]
  • 104.Adsit GS, Vaidyanathan R, Galler CM, Kyle JW, Makielski JC. Channelopathies from mutations in the cardiac sodium channel protein complex. Journal of Molecular and Cellular Cardiology. 2013;61:34–43. doi: 10.1016/j.yjmcc.2013.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Chen Y, Clarke OB, Kim J, Stowe S, Kim YK, Assur Z, Cavalier M, Godoy-Ruiz R, von Alpen DC, Manzini C, Blaner WS, Frank J, Quadro L, Weber DJ, Shapiro L, Hendrickson WA, Mancia F. Structure of the STRA6 receptor for retinol uptake. Science. 2016;353 doi: 10.1126/science.aad8266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Yus-Najera E, Santana-Castro I, Villarroel A. The identification and characterization of a noncontinuous calmodulin-binding site in noninactivating voltage-dependent KCNQ potassium channels. Journal of Biological Chemistry. 2002;277:28545–28553. doi: 10.1074/jbc.M204130200. [DOI] [PubMed] [Google Scholar]
  • 107.Ling KY, Preston RR, Burns R, Kink JA, Saimi Y, Kung C. Primary Mutations in Calmodulin Prevent Activation of the Ca2+-Dependent Na+ Channel in Paramecium, Proteins: Structure. Function, and Genetics. 1992;12:365–371. doi: 10.1002/prot.340120408. [DOI] [PubMed] [Google Scholar]
  • 108.Kung C, Preston RR, Maley ME, Ling KY, Kanabrocki JA, Seavey BR, Saimi Y. In vivo Paramecium mutants show that calmodulin orchestrates membrane responses to stimuli. Cell Calcium. 1992;13:413–425. doi: 10.1016/0143-4160(92)90054-v. [DOI] [PubMed] [Google Scholar]
  • 109.Schaefer WH, Hinrichsen RD, Burgess-Cassler A, Kung C, Blair IA, Watterson DM. A mutant Paramecium with a defective calcium-dependent potassium conductance has an altered calmodulin: A nonlethal selective alteration in calmodulin regulation. Proceedings of the National Academy of Sciences of the United States of America. 1987;84:3931–3935. doi: 10.1073/pnas.84.11.3931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Hennessey TM, Kung C. An anticalmodulin Drug, W-7, inhibits the voltage-dependent calcium current in Paramecium caudatum. Journal of Experimental Biology. 1984;110:169181. doi: 10.1242/jeb.110.1.169. [DOI] [PubMed] [Google Scholar]
  • 111.Ohya Y, Botstein D. Structure-based systematic isolation of conditional-lethal mutations in the single yeast calmodulin gene. Genetics. 1994;138:1041–1054. doi: 10.1093/genetics/138.4.1041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Makita N, Yagihara N, Crotti L, Johnson CN, Beckmann BM, Roh MS, Shigemizu D, Lichtner P, Ishikawa T, Aiba T, Homfray T, Behr ER, Klug D, Denjoy I, Mastantuono E, Theisen D, Tsunoda T, Satake W, Toda T, Nakagawa H, Tsuji Y, Tsuchiya T, Yamamoto H, Miyamoto Y, Endo N, Kimura A, Ozaki K, Motomura H, Suda K, Tanaka T, Schwartz PJ, Meitinger T, Kaab S, Guicheney P, Shimizu W, Bhuiyan ZA, Watanabe H, Chazin WJ, George AL. Novel Calmodulin (CALM2) Mutations Associated with Congenital Arrhythmia Susceptibility. Circulation: Cardiovascular Genetics. 2014 doi: 10.1161/CIRCGENETICS.113.000459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Limpitikul WB, Dick IE, Joshi-Mukherjee R, Overgaard MT, George AL, Jr, Yue DT. Calmodulin mutations associated with long QT syndrome prevent inactivation of cardiac L-type Ca currents and promote proarrhythmic behavior in ventricular myocytes. Journal Molecular and Cellular Cardiology. 2014 doi: 10.1016/j.yjmcc.2014.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Crotti L, Johnson CN, Graf E, De Ferrari GM, Cuneo BF, Ovadia M, Papagiannis J, Feldkamp MD, Rathi SG, Kunic JD, Pedrazzini M, Wieland T, Lichtner P, Beckmann BM, Clark T, Shaffer C, Benson DW, Kaab S, Meitinger T, Strom TM, Chazin WJ, Schwartz PJ, George AL., Jr Calmodulin mutations associated with recurrent cardiac arrest in infants. Circulation. 2013;127:1009–1017. doi: 10.1161/CIRCULATIONAHA.112.001216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Sondergaard MT, Sorensen AB, Skov LL, Kjaer-Sorensen K, Bauer MC, Nyegaard M, Linse S, Oxvig C, Overgaard MT. Calmodulin mutations causing catecholaminergic polymorphic ventricular tachycardia confer opposing functional and biophysical molecular changes. FEBS Journal. 2015;282:803–816. doi: 10.1111/febs.13184. [DOI] [PubMed] [Google Scholar]
  • 116.Saimi Y, Kung C. Ion Channel regulation by calmodulin binding. FEBS Letter. 1994;350:155–158. doi: 10.1016/0014-5793(94)00782-9. [DOI] [PubMed] [Google Scholar]
  • 117.Saimi Y, Kung C. Calmodulin as an Ion-Channel Subunit. Annual Review of Physiology. 2002;64:289–311. doi: 10.1146/annurev.physiol.64.100301.111649. [DOI] [PubMed] [Google Scholar]
  • 118.Ehlers MD, Augustine GJ. Cell signalling. Calmodulin at the channel gate, Nature, 399 ( 1999;105:107–108. doi: 10.1038/20079. [DOI] [PubMed] [Google Scholar]
  • 119.van Petegem F, Chatelain FC, Minor DL., Jr Insights into voltage-gated calcium channel regulation from the structure of the Ca(V)1.2 IQ domain-Ca2+/calmodulin complex. Nature Structural and Molecular Biology. 2005;12:1108–1115. doi: 10.1038/nsmb1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Hamilton SL. Ryanodine Receptors. Cell Calcium. 2005;38:253–260. doi: 10.1016/j.ceca.2005.06.037. [DOI] [PubMed] [Google Scholar]
  • 121.Xiong LW, Newman RA, Rodney GG, Thomas O, Zhang JZ, Persechini A, Shea MA, Hamilton SL. Lobe-dependent regulation of ryanodine receptor type 1 by calmodulin. Journal of Biological Chemistry. 2002;277:40862–40870. doi: 10.1074/jbc.M206763200. [DOI] [PubMed] [Google Scholar]
  • 122.Yamaguchi N, Xin C, Meissner G. Identification of apocalmodulin and Ca2+-calmodulin regulatory domain in skeletal muscle Ca2+ release channel, ryanodine receptor. Journal of Biological Chemistry. 2001;276:22579–22585. doi: 10.1074/jbc.M102729200. [DOI] [PubMed] [Google Scholar]
  • 123.Crivici A, Ikura M. Molecular and Structural Basis of Target Recognition by Calmodulin. Annual Review of Biophysics and Biomolecular Structure. 1995;24:85–116. doi: 10.1146/annurev.bb.24.060195.000505. [DOI] [PubMed] [Google Scholar]
  • 124.Klee CB. Interaction of calmodulin with Ca2+ and target proteins. In: Cohen P, Klee CB, editors. Calmodulin. Elsevier; New York: 1988. pp. 35–56. [Google Scholar]
  • 125.Klee CB, Vanaman TC. Calmodulin, Advances in Protein Chemistry. 1982;35:213321. doi: 10.1016/s0065-3233(08)60470-2. [DOI] [PubMed] [Google Scholar]
  • 126.Klee CB, Ren H, Wang X. Regulation of the calmodulin-stimulated protein phosphatase, calcineurin. Journal of Biological Chemistry. 1998;273:13367–13370. doi: 10.1074/jbc.273.22.13367. [DOI] [PubMed] [Google Scholar]
  • 127.Gifford JL, Walsh MP, Vogel HJ. Structures and metal-ion-binding properties of the Ca2+-binding helix-loop-helix EF-hand motifs. Biochemical Journal. 2007;405:199–221. doi: 10.1042/BJ20070255. [DOI] [PubMed] [Google Scholar]
  • 128.Kretsinger RH. Structure and Evolution of Calcium-Modulated Proteins. Critical Review in Biochemistry and Molecular Biology. 1980;8:119–174. doi: 10.3109/10409238009105467. [DOI] [PubMed] [Google Scholar]
  • 129.Persechini A, Moncrief ND, Kretsinger RH. The EF-hand family of calcium-modulated proteins. Trends in Neurosciences. 1989;12(11):462–467. doi: 10.1016/0166-2236(89)90097-0. [DOI] [PubMed] [Google Scholar]
  • 130.McPhalen CA, Strynadka NCJ, James MNG. Calcium-Binding Sites in Proteins: A Structural Perspective. Advances in Protein Chemistry. 1991;42:77–144. doi: 10.1016/s0065-3233(08)60535-5. [DOI] [PubMed] [Google Scholar]
  • 131.Ehlers MD, Zhang S, Bernhardt JP, Huganir RL. Inactivation of NMDA receptors by direct interaction of calmodulin with the NR1 subunit. Cell. 1996;84:745–755. doi: 10.1016/s0092-8674(00)81052-1. [DOI] [PubMed] [Google Scholar]
  • 132.Zhang S, Ehlers MD, Bernhardt JP, Su CT, Huganir RL. Calmodulin mediates calcium-dependent inactivation of N-methyl-D-aspartate receptors. Neuron. 1998;21:443–453. doi: 10.1016/s0896-6273(00)80553-x. [DOI] [PubMed] [Google Scholar]
  • 133.Moore CP, Rodney G, Zhang JZ, Santacruz-Toloza L, Strasburg G, Hamilton SL. Apocalmodulin and Ca2+ calmodulin bind to the same region on the skeletal muscle Ca2+ release channel. Biochemistry. 1999;38:8532–8537. doi: 10.1021/bi9907431. [DOI] [PubMed] [Google Scholar]
  • 134.Rodney GG, Williams BY, Strasburg GM, Beckingham K, Hamilton SL. Regulation of RYR1 Activity by Ca2+ and Calmodulin. Biochemistry. 2000;39:7807–7812. doi: 10.1021/bi0005660. [DOI] [PubMed] [Google Scholar]
  • 135.Samsó M, Wagenknecht T, Allen PD. Internal structure and visualization of transmembrane domains of the RyR1 calcium release channel by cryo-EM. Nature Structural & Molecular Biology. 2005;12:539–544. doi: 10.1038/nsmb938. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Geiser JR, van Tuinen D, Brockerhoff SE, Neff MM, Davis TN. Can Calmodulin Function without Binding Calcium? Cell. 1991;65:949–959. doi: 10.1016/0092-8674(91)90547-c. [DOI] [PubMed] [Google Scholar]
  • 137.Davis TN, Thorner J. Vertebrate and yeast calmodulin, despite significant sequence divergence, are functionally interchangeable. Proceedings of the National Academy of Sciences of the United States of America. 1989;86:7909–7913. doi: 10.1073/pnas.86.20.7909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Davis TN, Urdea MS, Masiarz FR, Thorner J. Isolation of the Yeast Calmodulin Gene: Calmodulin Is an Essential Protein. Cell. 1986;47:423–431. doi: 10.1016/0092-8674(86)90599-4. [DOI] [PubMed] [Google Scholar]
  • 139.Drum CL, Yan S, Bard J, Shen Y, LU D, Soelaiman S, Grabarek Z, Bohm A, Tang W. Structural basis for the activation of anthrax adenylyl cyclase exotoxin by calmodulin. Nature. 2002;415:396–402. doi: 10.1038/415396a. [DOI] [PubMed] [Google Scholar]
  • 140.Park S, Li C, Ames JB. Nuclear magnetic resonance structure of calcium-binding protein 1 in a Ca2+-bound closed state: implications for target recognition. Protein Science. 2011;20:1356–1366. doi: 10.1002/pro.662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Shea MA, Ackers GK. The OR Control System of Bacteriophage Lambda A physical-chemical model for gene regulation. Journal of Molecular Biology. 1985;181:211–230. doi: 10.1016/0022-2836(85)90086-5. [DOI] [PubMed] [Google Scholar]
  • 142.Shea MA, Miller MS, Yoder JB, Fowler CA, Feldkamp MD, Yu L. Calcium-Mediated Tailspin of Calmodulin on the IQ Motif of the Neuronal Voltage-Dependent Sodium Channel NaV1.2. Biophysical Journal. 2013;104:14a. [Google Scholar]
  • 143.Miller MS, Fowler CA, Feldkamp MD, Yu L, Shea MA. Calcium-Mediated Reversal of CaM on the NaV1.2 IQ Motif: Nested Anti-Parallel Sites. Biophysical Journal. 2014;106:48a. [Google Scholar]

Associated Data

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

Supplementary Materials

2

Supplemental Figure S1

Comparison of Metal-Chelating Residue Positions and Difference Maps of CaMC

A and B. CaM C-domain residues 76–148 (CaMC) in 8 structures were aligned using CaM residues 95–100 in Site III (93–104). CaMC is gray, Ca2+ ions are yellow spheres unless otherwise noted, and Mg2+ ions are gray. Residues are labeled according to the sequence of mammalian CaM.* For each labeled residue, the average distance (A) between the metal ion (Ca2+ or Mg2+) and the coordinating oxygen is given. For clarity, only site III is labeled, but the geometry of ion-chelation in site IV is equivalent.

A. Structures of Semi-Open CaMC - apo (calcium-depleted) or ion-bound CaMC (Mg2+)4-CaM bound to the IQ motif of NaV1.2(4DCK) and NaV1.5 (4OVN), and the IQ motif of Myosin VI (3GN4 chain H), and (Ca2+)4-CaM bound to the IQ motif of NaV1.2 (4JPZ) and Nav1.5 (4JQ0). Mg2+ ions are gray spheres and Ca2+ in 4JPZ and 4JQ0 are magenta spheres. Side chains of metal-chelating residues in site III are highlighted with sticks: 4DCK - blue, 4OVN – cyan, 4JPZ - magenta, 4JQ0 - pink, and 3GN4 chain H - light blue.

B. Structures of Open CaMC (Ca2+)4-CaM alone (1CLL), and (Ca2+)4-CaM bound to a CaMKII peptide (1CDM) or a Myosin VI “Insert 2” (3GN4 Chain B) Side chains of Ca2+-chelating residues in site III are highlighted with sticks: 1CLL – orange, 1CDM – red, and 3GN4 chain B – salmon. Figures made with PyMOL (Schrödinger, LLC).

C. Calcium Fo-Fc Electron Density Maps The Fo-Fc density maps for two Ca2+ ions (yellow) located in sites III and IV in each calmodulin chain were contoured at +3 (green mesh, insufficient density) and −3 (red mesh, excess density) sigma to a radius of 5 Å from the center of each ion in calmodulin alone (1CLL), or in calmodulin bound to the IQ motif of NaV1.2 in 4JPZ [Chain C, Chain I]. Figures made with PyMOL (Schrödinger, LLC). PDB files and electron density maps were obtained from PDB_REDO (http://www.cmbi.ru.nl/pdb_redo/) Calmodulin residues 76–148 (gray) were aligned based on residues 117–128 (i.e., helix “G”).

*Mammalian CaM has 148 amino acids. Standard numbering begins with A1, and ends with K148. The Met listed as residue 1 in the UNIPROT database is not found in mammalian CaM, or in the bacterially expressed sequence of mammalian CaM.

Supplemental Figure S2

HSQC spectra of 13C,15N-(Ca2+)4-CaM(1–148)-Nav1.2IQp vs. 13C,15N-(Ca2+)2-CaM(1–75)

HSQC overlay of full-length CaM saturated with calcium and Nav1.2IQ [(Ca2+)4CaM-Nav1.2IQp; black] with free CaMN [(Ca2+)2-CaM(1–75); blue]. Peaks that are nearly identical are circled in green. The high level of agreement between these two spectra indicates that few residues in the N-domain of CaM are perturbed by CaM binding to the IQ motif and the C-domain makes most contacts in the CaM-Nav1.2IQp interface.

Supplemental Figure S3

Ramachandran Plot

PROCHECK-NMR analysis [70] of the minimized average structure in 2M5E.

Supplemental Figure S4

NOESY Overlay of Nav1.2IQp Bound to CaM ± Ca2+

Overlays comparing two regions of the doubly 12C,14N filtered NOESY spectra of NaV1.2IQp bound to apo (green) and Ca2+-saturated (black) CaM. Crosspeaks are shown (a) between upfield and downfield aliphatic protons and (b) between amide plus aromatic protons and aliphatic protons. The significant differences suggest that NaV1.2IQp is in a very different chemical environment when bound to apo CaM vs. calcium-saturated CaM.

RESOURCES