Abstract
Calcium-binding protein 1 (CaBP1), a calmodulin (CaM) homolog, endows certain voltage-gated calcium channels (CaVs) with unusual properties. CaBP1 inhibits CaV1.2 calcium-dependent inactivation (CDI) and introduces calcium-dependent facilitation (CDF). Here, we show that the ability of CaBP1 to inhibit CaV1.2 CDI and induce CDF arises from interaction between the CaBP1 N-lobe and interlobe linker residue Glu94. Unlike CaM, where functional EF hands are essential for channel modulation, CDI inhibition does not require functional CaBP1 EF-hands. Furthermore, CaBP1-mediated CDF has different molecular requirements than CaM-mediated CDF. Overall, the data show that CaBP1 comprises two structural modules having separate functions: similar to CaM, the CaBP1 C-lobe serves as a high-affinity anchor that binds the CaV1.2 IQ domain at a site that overlaps with the Ca2+/CaM C-lobe site, whereas the N-lobe/linker module houses the elements required for channel modulation. Discovery of this division provides the framework for understanding how CaBP1 regulates CaVs.
Introduction
Voltage-gated calcium channels (CaVs) serve as a major source of calcium influx in excitable cells (Catterall, 2000). Because calcium ions are chemical messengers (Clapham, 2007), influx through CaVs can directly link membrane potential charges to stimulation of intracellular signaling cascades (Catterall, 2000). Although high-voltage activated CaVs consist of four essential components (Van Petegem and Minor, 2006): a CaV1 or CaV2 pore-forming CaVα1 (Catterall, 2000), a cytoplasmic CaVβ (Dolphin, 2003), CaVα2δ (Davies et al., 2007), and calmodulin (CaM) (Pitt, 2007), the composition of these large protein complexes is not monolithic. In some contexts, such as cerebellar and hippocampal neurons (Lee et al., 2002; Zhou et al., 2004), photoreceptor synapses (Haeseleer et al., 2004), and auditory hair cells (Cui et al., 2007; Yang et al., 2006), members from a family of calcium binding proteins homologous to CaM, known as CaBPs (Haeseleer et al., 2000), can replace CaM. This component exchange has profound effects on how CaVs respond to calcium entry and results in channels that have strikingly different functional properties than those modulated by CaM (Cui et al., 2007; Few et al., 2005; Lautermilch et al., 2005; Lee et al., 2002; Yang et al., 2006; Zhou et al., 2004; Zhou et al., 2005).
When modulated by CaM, many CaV1s exhibit a strong calcium-dependent inactivation (CDI) that limits calcium influx during depolarization (Dunlap, 2007). In contrast, CaV1s under the influence of CaBP1, a CaBP abundant in the brain and retina (Haeseleer et al., 2000), have dramatically altered functional properties. CaBP1 blocks CaV1.2 (Zhou et al., 2004; Zhou et al., 2005) and CaV1.3 (Cui et al., 2007; Yang et al., 2006) CDI and introduces an increase in CaV1.2 (Zhou et al., 2004) peak current upon repetitive stimulation, calcium-dependent facilitation (CDF). These effects depend on displacement of CaM from the CaVα1 C-terminal IQ domain (Yang et al., 2006; Zhou et al., 2004), a channel element that is critical for CaM-mediated CDI (Erickson et al., 2003; Zühlke et al., 1999).
As with many calcium sensor proteins (Burgoyne et al., 2004; Haeseleer et al., 2002; Weiss and Burgoyne, 2002), both CaBP1 and CaM comprise two lobes that bear a pair of EF-hand calcium binding motifs (Haeseleer et al., 2000). Beyond this common architecture, there are a number of differences. CaBP1 is myristoylated at its N-terminus, has an N-lobe that is largely insensitive to calcium (Kd > 100 μM) (Wingard et al., 2005), and a longer interlobe linker. Apart from recent NMR structures of individual CaBP1 lobes (Li et al., 2009), little is known about the high-resolution structure of complete CaBPs or the details of how they interact with CaVs. Thus, it has remained uncertain which CaBP1 features contribute to the stark functional differences that CaBP1 brings to CaV modulation.
We set out to answer this question with respect to CaV1.2 and find that the key element is an interaction between the N-lobe and an interlobe linker residue that is conserved among CaBP family members. We further show that CaBP1 comprises two structural modules having separate functions. The CaBP1 C-lobe shares both structural and functional similarity with the CaM C-lobe and acts as a high-affinity anchor that binds the CaV1.2 IQ domain at a site that overlaps with the Ca2+/CaM C-lobe binding site. In contrast, the N-lobe/linker module contains the elements required for the unique functional effects CaBP1 has on CaV1.2. Delineation of these conserved CaBP structural elements provides the required framework for understanding how this family of calcium sensor proteins affects CaVs and other target proteins in a manner distinct from CaM.
Results
CaBP1 N-lobe and interlobe linker are crucial for CDI inhibition
We used two-electrode voltage clamp recording of Xenopus oocytes to test whether CaBP1 N-terminal myristoylation contributes to CaBP1 modulation of CaV1.2 as this lipid anchor is important for CaBP1 modulation of CaV2.1 (Few et al., 2005) and is an obvious difference from CaM (Figure 1A). Examination of calcium currents produced when CaV1.2 was co-expressed with a CaBP1 G2A mutant (Few et al., 2005) that blocks myristoylation (Rocque et al., 1993) demonstrated that CaBP1 G2A could inhibit CDI as effectively as wild type CaBP1 (Figures 1B and C, and Table 1). The enzyme that carries out myristoylation is unable to process G2A substrates and is conserved (Rocque et al., 1993). Nevertheless, to ensure that the N-terminal modification was eliminated, we also examined the consequences of deletion of the entire CaBP1 myristoylation site, CaBP1(Δ2-15). Similar to CaBP1 G2A, CaBP1(Δ2-15) fully inhibited CDI (Figures 1B and C, and Table 1). Thus, N-terminal myristoylation is not required for CaBP1 inhibition of CaV1.2 CDI.
Figure 1.
CaBP1-CaM chimeras demonstrate the importance of the N-lobe and interlobe linker. A, Structure based alignment of CaBP1 and CaM. N-lobe and C-lobe are shaded green and blue, respectively. CaBP1 N-terminal myristoylation segment and interlobe linker are indicated by the yellow and orange bars, respectively. EF hand positions are indicated above the sequence according to standard nomenclature (Gifford et al., 2007). Secondary structure elements are indicated. CaBP1 residues Y56 and E94 are highlighted in red. B, CaBP1-mediated CDI inhibition is unaffected by myristoylation site mutants. Representative normalized ICa traces at a test potential of +20 mV for CaV1.2 expressed with CaVβ2a and CaBP1 myristoylation mutants G2A and Δ2-15 (grey), CaBP1 (orange), or CaM (blue). C, ti300 values for data from ‘B’. D, Representative normalized ICa traces at a test potential of +20 mV for CaV1.2 expressed with CaVβ2a and CaBP1-CaM chimeras (grey), wild-type CaBP1 (orange), or CaM (blue). Icons show the origin of the domain elements: CaBP1 (orange), CaM (blue). Chimeras containing CaBP1 N-lobe all contain the myristoylation domain. E, ti300 values for data from ‘D’. F, Relative current amplitude changes of ICa in response to 40 +20 mV pulses delivered at 3 Hz for CaV1.2 expressed with CaVβ2a and the indicated chimeras. First (black), last (red) and every fourth trace (grey) are shown. G, Relative current changes between the last and first pulse for CaV1.2 expressed with CaVβ2a and CaBP1, CaM, or the indicated chimeras. Parentheses indicate the number of oocytes tested. Error bars indicate s.d. See also Figures S1 and S2.
Table 1.
CaV1.2 inactivation parameters
|
Data are expressed as mean values ± standard deviation. All experiments correspond to the mean of at least two separate oocyte batches.
Because CaBP1 (Li et al., 2009) and CaM (Masino et al., 2000; Tsalkova and Privalov, 1985) lobes fold independently, we designed a set of CaBP1-CaM chimeras based on exchange of three elements (N-lobe, interlobe linker, and C-lobe) and tested their effects on CaV1.2 expressed in Xenopus oocytes. We designate the chimeras with three letter names, e.g. MMB, in which the first, second, and third positions denote N-lobe, interlobe linker, and C-lobe, respectively, and each letter denotes the element origin: ‘M’ for CaM and ‘B’ for CaBP1.
All CaM N-lobe chimeras (MMB, MBB, MBM) lacked the hallmark CaV1.2 CDI inhibition caused by CaBP1, and instead supported CDI similar to CaM (Figures 1D, E, and Table 1). CaM interlobe linker chimeras (MMB, BMB and BMM) also allowed CDI to proceed. Notably, the chimera having the CaM interlobe linker exchanged into CaBP1, BMB, failed to inhibit CDI. Pulldown assays showed that all chimeras retained ability to interact with the CaV1.2 IQ domain (Figure S1) and eliminate the possibility that the absence of CDI inhibition arose from failure of the chimeras to fold properly and bind the CaV1.2 IQ domain. Together, these results suggest that the inability of MMB, MBB, MBM, MMB, BMB, and BMM to inhibit CDI arises from the absence of elements in the CaBP1 N-lobe (MMB, MBB, MBM, MMB) and CaBP1 interlobe linker (MMB, BMB, and BMM). In further support of this, we found that BBM, which has CaBP1 N-lobe and interlobe linker joined to CaM C-lobe, blocks CaV1.2 CDI even more potently than CaBP1 (Figures 1D and E). BBM additionally causes slower CaV1.2 activation (Figure 1E). Taken together, the results from the chimeras strongly suggest that the key elements underlying the CaBP1 and CaM functional differences with respect to CDI reside in the N-terminal lobe and interlobe linker.
Besides CDI inhibition, CaBP1 causes CaV1.2 CDF (Zhou et al., 2004) (Figure 1F). We tested whether the CaBP1-CaM chimeras retained this property. Chimeras bearing either the CaM N-lobe (MMB, MBB, MBM, MMB) or CaM interlobe linker (MMB, BMB, and BMM) were unable to support CaV1.2 CDF. Except for BMB, which introduced a larger progressive loss in current amplitude (Figures 1F and G), channels expressed with these chimeras were indistinguishable from channels expressed with CaM. In contrast, BBM caused CaV1.2 CDF that was ~2-fold stronger than that of CaBP1 (Figures 1F and G). Hence, BBM embodies both main functional properties of CaBP1, the ability to inhibit CaV1.2 CDI and the ability to confer CDF. Together, the data indicate that the CaBP1 N-lobe and interlobe linker bear the modulatory elements unique to CaBP1, whereas CaBP1 and CaM C-terminal lobes perform similar functions.
CaBP1 interlobe linker functional properties
The CaBP1 and CaM interlobe linker lengths differ by four residues (Figure 1A), a divergence conserved among CaBPs (Haeseleer et al., 2000) (Figure S2). Given the apparent importance of the interlobe linker, we investigated whether its length, composition, or both were crucial for CaBP1 function. CaBP1 constructs having an interlobe linker composed of the first four (93-96, ‘AETA’) or last four interlobe linker residues (97-100, ‘DMIG’) failed to inhibit CDI (Figures 2A and B). Replacement of the CaBP1 interlobe linker with a duplication of the CaM interlobe linker (DTDSDTDS), octa-alanine (8A), or octa-glycine (8G) also failed to inhibit CDI (Figures 2A and B).
Figure 2.
Interlobe linker length and sequence have a crucial role in CaBP1 function. A, Representative normalized ICa traces at a test potential of +20 mV for CaV1.2 expressed with CaVβ2a and indicated CaBP1 linker mutants (grey), CaBP1 (orange), or CaM (blue). B, ti300 values for data from ‘A’. BMB values are from Figure 1. C, Relative current amplitude changes of ICa and IBa in response to 40 +20 mV pulses delivered at 3 Hz for CaV1.2 expressed with CaVβ2a and CaBP1, CaM, or the DMIG linker mutant. First (black), last (red) and every fourth trace (grey) are shown. D, Averaged frequency dependent change of ICa current amplitude for CaV1.2 expressed with CaVβ2a and CaBP1, CaM, or the indicated linker mutants. E, Comparison of ICa recovery from inactivation for CaV1.2 expressed with CaVβ2a and CaM (blue) or the CaBP1 DMIG linker mutant (orange) in response to two 450 ms pulses to +20 mV separated by variable time intervals. Parentheses indicate the number of oocytes tested. Error bars indicate s.d. See also Figures S1 and S2.
Unexpectedly, the protocol used to induce CDF caused CaV1.2 to display a strong, calcium-dependent reduction in current amplitude in the presence of all of the CaBP1 interlobe linker mutants (Figure 2C). This phenomenon, which we term ‘CDI tachyphylaxis’, is stronger than the small current suppression seen with CaM (Figure 2D) and provides evidence that the interlobe linker manipulations did not incapacitate the CaBP1 mutants. This is corroborated by pulldown experiments that show the individual mutants retain the ability to bind the CaV1.2 IQ domain (Figure S1B). Further analysis of the DMIG mutant showed that the CDI tachyphylaxis arises from changes in recovery from inactivation (Figures 2E). Following a depolarization pulse, CaV1.2 co-expressed with CaM shows essentially full recovery after 750 ms. In contrast, ~7% of CaV1.2 co-expressed with DMIG fail to recover in the same period. Over longer interpulse periods, both CaM and DMIG containing channels recover fully (Figure 2E). Taken together, the data from the chimeras and interlobe linker mutants establish that both the length and composition CaBP1 interlobe linker are crucial for modulation of CaV1.2.
N- and C-lobes contribute to CaBP1-CaV1.2 IQ domain affinity
We turned to isothermal titration calorimetry (ITC) to investigate how CaBP1 interacts with the CaV1.2 IQ domain, the domain that is essential for CaBP1 CDI inhibition (Zhou et al., 2004). Experiments using individual CaBP1 lobes in the presence of 1 mM calcium, Ca2+/N-lobeBP and Ca2+/C-lobeBP, revealed that each has a single binding site on the CaV1.2 IQ domain (Figure 3A, B and Table 2). Ca2+/N-lobeBP binding is an endothermic reaction having modest affinity (Kd = 1.11 ± 0.08 μM), whereas Ca2+/C-lobeBP binds ~100-fold stronger via an exothermic reaction that has an affinity (Kd = 10.5 ± 1.9 nM) similar to Ca2+/C-lobeCaM (Van Petegem et al., 2005). Competition experiments in which Ca2+/N-lobeBP was titrated into a pre-formed Ca2+/C-lobeBP-CaV1.2 IQ domain complex demonstrate that Ca2+/C-lobeBP prevents Ca2+/N-lobeBP binding and indicate that the binding sites overlap (Figure 3C). As expected from the affinity differences, Ca2+/C-lobeBP can displace Ca2+/N-lobeBP from the CaV1.2 IQ domain (Figure 3D). The ability of both Ca2+/CaBP1 lobes to bind the CaV1.2 IQ domain at an overlapping site is reminiscent of the behavior of individual Ca2+/CaM lobes (Kim et al., 2008; Van Petegem et al., 2005).
Figure 3.
Characterization of Ca2+/CaBP1-CaV1.2 IQ domain interactions. Exemplar ITC titrations for: A, 75 μM Ca2+/N-lobeBP into 7.5 μM IQ domain. B, 75 μM Ca2+/C-lobeBP into 7.5 μM IQ domain. C, 75 μM Ca2+/N-lobeBP into 7.5 μM CaV1.2 IQ domain and 11.25 μM Ca2+/C-lobeBP. D, 75 μM Ca2+/C-lobeBP into 7.5 μM IQ domain and 50 μM Ca2+/N-lobeBP. E, 75 μM Ca2+/CaBP1 into 5.5 μM IQ domain. F, 75 μM Ca2+/C-lobeBP and 75 μM Ca2+/N-lobeBP into 7.5 μM IQ domain. Red squares denote estimates of enthalpy changes using binding parameters derived from the single lobe binding.G, 75 μM Ca2+/CaBP1 into 7.5 μM IQ domain and 50 μM Ca2+/N-lobeBP. H, Thermodyamic cycle for analysis of the binding of CaBP1 (Ka3) to the CaV1.2 IQ domain. Ka1 describes Ca2+/N-lobeBp binding. Ka2 describes the competition of Ca2+/N-lobeBP by CaBP1. Cartoons depict the observed binding reactions.
Table 2.
Titration calorimetry data
| Cell | Syringe | T (°C) |
N | Kd (nM) | n | ΔH (kcal mol−1) |
ΔS (cal mol−1 K−1) |
Ka/Ka (wt) | ΔΔG (kcal mol−1) |
|---|---|---|---|---|---|---|---|---|---|
| Cav1.2 IQ | Ca2+/N-lobeBP | 15 | 0.82 ± 0.05 | 1,100 ± 80 | 3 | 2.09 ± 0.40 | 34.5 ± 1.5 | - | - |
| Ca2+/C-lobeBP | 15 | 0.90 ± 0.07 | 10.5 ± 1.9 | 4 | −2.88 ± 0.19 | 26.5 ± 0.6 | - | - | |
| Ca2+/C-lobeBP | 25 | 0.98 ± 0.11 | 15.0 ± 0.7 | 2 | −8.63 ± 0.24 | 6.9 ± 0.9 | - | - | |
| Ca2+/C-lobeCaM* | 15 | 0.98 ± 0.18 | 2.63 ± 0.07 | 2 | −6.77 ± 0.24 | 15.75 ± 0.78 | - | - | |
|
| |||||||||
| Cav1.2 IQ Ca2+/N-lobe |
Ca2+/C-lobeBP | 15 | 0.96 ± 0.01 | 16.6 ± 0.7 | 2 | −3.17 ± 0.05 | 24.1 ± 0.7 | - | - |
| Ca2+/CaBP1 | 15 | 0.93 ± 0.12 | 0.288 ± 0.070 | 2 | −4.72 ± 0.07 | 26.6 ± 0.7 | - | - | |
| Ca2+/CaBP1 E94A | 15 | 1.04 ± 0.04 | 0.336 ± 0.097 | 2 | −7.75 ± 0.31 | 15.9 ± 0.5 | 0.86 ± 0.32 | −0.06 ± 0.22 | |
|
| |||||||||
| Cav1.2 IQ F1628A |
Ca2+/C-lobeBP | 25 | 0.88 ± 0.03 | 727 ± 224 | 2 | −4.87 ± 0.11 | 11.8 ± 1.0 | 0.021 ± 0.006 | −2.25 ± 0.18 |
|
| |||||||||
| CaV1.2 IQ F1631A |
Ca2+/C-lobeBP | 15 | 0.83 ± 0.03 | 345 ± 107 | 2 | −2.22 ± 0.20 | 21.9 ± 1.3 | 0.030 ± 0.011 | −1.82 ± 0.21 |
| Ca2+/C-lobeCaM | 15 | 0.80 ± 0.03 | 17.3 ± 4.3 | 2 | −4.87 ± 0.08 | 18.7 ± 0.8 | 0.152 ± 0.038 | −1.22 ± 0.14 | |
Data are expressed as mean values ± standard deviation. ΔΔG is calculated with respect to the wild-type lobe or protein at 15°C with the exception of data for F1628A, which are calculated at 25°C.
indicates values from (Van Petegem et al., 2005). See also Figure S5.
Unlike the simple, individual lobe binding isotherms, titration of full-length Ca2+/CaBP1 into the CaV1.2 IQ domain showed a V-shaped isotherm that could not be attributed to a single binding event (Figure 3E). Because Ca2+/N-lobeBP and Ca2+/C-lobeBP bind to the CaV1.2 IQ domain in a competitive manner, we wondered whether the complex isotherm arose from contributions of each lobe. ITC experiments in which equimolar portions of individual Ca2+/CaBP1 lobes were titrated into the CaV1.2 IQ domain produced a binding isotherm very similar to that of full-length Ca2+/CaBP1 (Figure 3F). Further, when we used parameters from the single lobe experiments to simulate the isotherm in which Ca2+/N-lobeBP and Ca2+/C-lobeBP bind to a single, overlapping IQ domain site, we found excellent correspondence to the measured isotherm (Figure 3F, red triangles). These results indicate that the ‘V’-shaped nature of the isotherm represents a sequence of two events: (1) independent binding of Ca2+/N-lobeBP and Ca2+/C-lobeBP to separate CaV1.2 IQ domains when the IQ domain is in excess, and (2) replacement of Ca2+/N-lobeBP by Ca2+/C-lobeBP as the IQ domain becomes limiting. The ability to dissect the binding reaction this way set the stage for an experiment to determine the thermodynamics of the Ca2+/CaBP1- CaV1.2 IQ domain interaction and test whether the ‘V’-shaped isotherm observed with full-length Ca2+/CaBP1 (Figure 3E) arose from a similar course of events.
Titration of Ca2+/CaBP1 into pre-formed Ca2+/N-lobeBP-CaV1.2 IQ domain complexes yielded a titration isotherm having a single transition (Figure 3G). Analysis using competition ligand binding by displacement (Sigurskjold, 2000) in the context of the Figure 3H thermodynamic cycle reveals that Ca2+/CaBP1 binds the CaV1.2 IQ domain ~40-fold stronger than measured for Ca2+/C-lobeBP alone (Kd= 296 ± 70 pM)(Table 2). This increased affinity is accompanied by a binding enthalpy increase that indicates that Ca2+/N-lobeBP, the interlobe linker, or both contribute to the binding reaction by interacting with the CaV1.2 IQ domain at sites separate from the Ca2+/C-lobeBP binding site. Taken together, the ITC experiments establish that CaBP Ca2+/C-lobe interacts with the CaV1.2 IQ domain in a manner similar to Ca2+/CaM C-lobe, and show that elements from the entire CaBP1 participate CaV1.2 IQ domain binding.
Functional EF-hands not required for CDI inhibition
CaBP1 has four EF hands; however, the importance of metal binding to N-lobe EF-hands is unclear. EF1 has weak Ca2+ affinity (Wingard et al., 2005) and EF2 is non-functional due to the lack of a canonical residue at the ‘z’ position (Figure 1A)(Gifford et al., 2007; Haeseleer et al., 2000). To test whether CaBP1 inhibition of CaV1.2 CDI requires the ability of the CaBP1 EF-hands to bind metal ions, we examined the consequences of introduction of a D➔A mutation at the ‘x’ position of each functional EF hand. This mutation is analogous to those used to dissect CaM EF hand function (Peterson et al., 1999) and should reduce metal-binding ability substantially and. CaBP1 bearing a disrupted EF1 was functionally indistinguishable from wild-type (Figures 4A and B, and Table 1). In contrast, EF3, EF4, and EF34 mutations diminished but did not eliminate the ability of CaBP1 to inhibit CaV1.2 CDI. Thus, the capacity of CaBP1 C-lobe EF hands to bind metal ions is important but not essential for CDI inhibition. This relative insensitivity to EF hand disruption stands in contrast to CaM where functional C-lobe EF-hands are required for CDI (Alseikhan et al., 2002; Peterson et al., 1999). The effects of CaBP1 EF34 are reminiscent of the ability of the CaM EF34 mutant to block CDI (Peterson et al., 1999) and suggest that part of the CaBP1 mechanism may be competition with apo-CaM. In contrast to the minor effects on CDI inhibition, the EF3 and EF4 mutants greatly diminished CaV1.2 CDF (Figure 4C and D) and indicate that CaBP1-mediated CDF requires Ca2+ binding to the C-lobe.
Figure 4.
C-terminal EF-hands are required for CaBP1-mediated CDF but not for inhibition of CDI. A, Representative normalized ICa traces at a test potential of +20 mV for CaV1.2 expressed with CaVβ2a and indicated CaBP1 EF hand mutants (grey), CaBP1 (orange), or CaM (blue). B, ti300 values for data from ‘A’. C, Relative current amplitude changes of ICa in response to 40 +20 mV pulses delivered at 3 Hz for CaV1.2 expressed with CaVβ2a and the indicated chimeras indicated CaBP1 EF hand mutants, CaBP1, or CaM. D, Averaged frequency dependent change of ICa current amplitude for CaV1.2 expressed with CaVβ2a and the indicated chimeras indicated CaBP1 EF hand mutants, CaBP1, or CaM. Parentheses indicate the number of oocytes tested. Error bars indicate s.d. P-values with respect to wild-type were calculated using student’s t-Test: N.S., not significant, P > 0.05; ***, P < 0.005.
CaBP1 crystal structure
To understand how the CaBP1 N-lobe and interlobe linker contribute to function, we crystallized and determined the structure of the CaBP1 functional core, CaBP1(Δ2-15). CaBP1(Δ2-15) crystallized in the I23 space group and diffracted X-rays to 2.9Å (Table 3). Surface entropy reduction (Derewenda and Vekilov, 2006) identified a mutant, CaBP1(Δ2-15) K130A, that did not alter function (Table 1), gave crystals having a different space group, P3121 and improved resolution, 2.4Å, and that enabled solution by MAD (Hendrickson and Ogata, 1997) using selenomethionine-substituted protein. The 2.4Å structure was used for molecular replacement of the I23 crystal form. As there were no major differences between the structures, we used chain A from the 2.4Å structure for analysis.
Table 3.
Crystallographic data collection and refinement statistics
| CaBP1 | CaBP1 K130A | ||||
|---|---|---|---|---|---|
| Data Collection | Native | SeMet | |||
| Space group | I23 | P3121 | P3121 | ||
| Cell dimensions a/b/c (Å) | 206, 206, 206 | 69, 69, 344 | 69, 69, 345 | ||
| α/β/γ (°) | 90, 90, 90 | 90, 90, 120 | 90, 90, 120 | ||
| Resolution (Å) | 50-2.9 (3.06 – 2.90) |
50 - 2.4 (2.53 - 2.40) |
50-2.9 (3.06 - 2.90) |
||
| Peak | Inflection | Remote | |||
| Wavelength | 1.11587 | 1.11587 | 0.97958 | 0.97972 | 1.01987 |
| Rsym (%) | 9.5 (45.1) | 7.2 (43.2) | 9.0 (38.2) | 8.5 (34.9) | 7.9 (28.6) |
| I / σI | 6.4 (1.6) | 7.1 (1.6) | 6.7 (1.8) | 6.9 (2.0) | 7.2 (2.4) |
| Completeness (%) | 97.5 (96.2) |
99.4 (99.0) |
100.0 (100.0) |
100.0 (100.0) |
99.9 (100.0) |
| Redundancy | 4.0 (2.1) | 3.9 (2.8) | 6.9 (7.2) | 6.9 (7.2) | 6.9 (7.2) |
| Unique reflections | 32321 | 43362 | |||
| Average B-factor | 99.7 | 54.4 | |||
| Refinement | |||||
| Rwork/Rfree (%) | 21.3 / 25.0 | 25.2 / 29.5 | |||
| No. of chains in AU | 6 | 6 | |||
| No. of protein atoms | 7093 | 6835 | |||
| No. of ligand atoms | 12 | 42 | |||
| RMSD bond lengths (Å) | 0.020 | 0.016 | |||
| RMSD angles (°) | 1.697 | 1.556 | |||
| Ramachandran best /disallowed regions (%) |
94.9/0.0 | 96.1/0.0 | |||
One crystal was used for each dataset. Values in parentheses are for highest-resolution shell.
CaBP1 has four EF-hands arranged into two lobes. Unexpectedly, a well-ordered interlobe linker (residues 93-100) connects the lobes (Figure 5A). N-lobe EF hands, EF1 and EF2, lack ligands; whereas C-lobe EF hands, EF3 and EF4, each contain a calcium ion. The compact structure is different from what has been observed by NMR, which found independent lobes and a disordered interlobe linker (Li et al., 2009). Comparison with crystal structures of compact (Fallon and Quiocho, 2003) and extended (Wilson and Brunger, 2000) Ca2+/CaM conformations reveals that the CaBP1 lobe orientation is different from both (Figure S3A). Despite the compact conformation, the CaBP1 lobes make few contacts with each other and bury only 270 Å2.
Figure 5.
Structure of CaBP1. A, Cartoon representation of CaBP1. N-lobe (green), linker (red), and C-lobe (blue) are indicated. Calcium ions are depicted as yellow spheres. Secondary structure elements are labeled. B, Superposition of CaBP1 apo-N-lobe structure determined by X-ray (green) and Mg2+-bound form determined by NMR (2K7B)(Li et al., 2009) (light green). Mg2+ ion is a yellow sphere. B-hairpin from the Mg2+-bound form is shown (salmon) and labeled. Equivalent residues in the apo-lobe structure are highlighted in red. C, Superposition of CaBP1 Ca2+/C-lobe structure determined by X-ray (blue) and NMR (grey). D, Superposition of CaBP1 Ca2+/C-lobe (dark blue) and CaM Ca2+/C-lobe (marine) (1EXR)(Van Petegem et al., 2005). Hydrophobic residues involved in CaM Ca2+/C-lobe-IQ domain binding and their CaBP1 Ca2+/C-lobe counterparts are shown as sticks. E, Details of the Glu94/N-lobe cleft interaction. Glu94 and residues from the N-lobe are shown as sticks. N-lobe is colored green and shown in surface representation. Linker is indicated in red. E94 is shown in salmon. Hydrogen bond interaction between E94 and the backbone carbonyls of Met54 is shown as black dashes. See also Figure S3.
N-lobe is in the calcium-free apo-form. Neither EF1 nor EF2 have electron density indicative of a metal ion, despite crystallization conditions containing millimolar Ca2+. Apo-N-lobe shows a number of differences from the Mg2+-bound N-lobe NMR structure (Li et al., 2009) (RMSDCα=2.5Å). The most notable is in the E1 helix (Figure 5B). In the apo-form, E1 extends by an extra turn that encompasses the ‘x’ and ‘y’ ligand-binding residues, Asp35 and Asp37, and clashes with the metal ion position in the Mg2+-bound structure (Li et al., 2009) (Figure 5B). Additionally, apo-N-lobe EF1 and EF2 are further away from each other and lack the short β-hairpin that connects them in the Mg2+-bound structure (Figure 5B). These differences indicate that conformational changes, particularly in EF1, accompany metal binding and are consistent with divalent-ion induced chemical shift changes (Wingard et al., 2005). Although there are some differences in packing, the N-lobe hydrophobic core is buried in both forms (Figure S3B) and does not undergo large changes in exposure seen in the classic apo➔Ca2+/CaM transition (Gifford et al., 2007; Grabarek, 2006) or in the CaBP1 apo➔Ca2+/C-lobe transition (Li et al., 2009). The lack of large divalent ion induced conformational changes agrees with the absence of EF1 D➔A functional effects (Figures 4A and B) and suggests that N-lobe metal binding is unnecessary for CaBP1-mediated CaV1.2 modulation.
CaBP1 Ca2+/C-lobe is similar to the NMR structure (RMSDCα = 1.3Å over 60 Cα positions) (Figure 5C) and resembles the Ca2+/CaM C-lobe ‘open’ conformation. The structural similarity is even better when compared to Ca2+/CaM crystal structures regardless of whether the comparison is made with Ca2+/CaM C-lobe alone (Wilson and Brunger, 2000) or bound to a target such as the CaV1.2 IQ domain (Van Petegem et al., 2005), (RMSDCα =1.0 Å over 60 Cα positions for each) (Figure 5D). This similarity extends to many of the hydrophobic pocket sidechains. Consequently, the CaBP1 Ca2+/C-lobe hydrophobic pocket appears ready to engage the CaV1.2 IQ domain in a manner similar to CaM Ca2+/C-lobe (Van Petegem et al., 2005).
The most striking aspect of the structure is the well-ordered interlobe linker (Figures 5A and S3C), which is clamped to the N-lobe by insertion of Glu94 into a largely hydrophobic N-lobe cleft lined by Tyr56 and Leu91 (Figure 5E). This interaction is absent from the NMR structure (Li et al., 2009). Within the N-lobe cleft, the Glu94 sidechain is within hydrogen bonding distance of the Met54 backbone carbonyl oxygen. Comparison of non-crystallographically related asymmetric unit monomers reveals that the single helical turn in the C-terminal half of the linker (residues 95-99) forms a hinge point for a 5° rotation between the lobes (Figure S3D) and indicates that even in the compact state there is flexibility between the lobes.
Ca2+/C-lobeBP and Ca2+/C-lobeCaM-IQ domain binding sites overlap
MMB and BBM chimeras having swapped CaBP1 and CaM C-lobes retain the properties of the parent, suggesting that CaBP1 and CaM C-lobes are functionally interchangeable (Figure 1). This, together with the strong structural similarity between Ca2+/C-lobeBP and Ca2+/C-lobeCaM bound to the CaV1.2 IQ domain (Figure 5D) and the prior evidence that CaBP1 acts by competing with CaM for the IQ domain (Zhou et al., 2004) led us to wonder whether CaBP1 and CaM Ca2+/C-lobes interact with the CaV1.2 IQ domain at a common site. Titration of Ca2+/C-lobeBP into a pre-formed Ca2+/C-lobeCaM-CaV1.2 IQ domain complex shows that Ca2+/C-lobeCaM blocks binding of Ca2+/C-lobeBP (Figure 6A) and supports the idea that the high affinity CaV1.2 IQ domain binding sites for each overlap. To explore this possibility further, we examined the effects of alanine mutations at F1628 and F1631, two residues that serve as anchors for Ca2+/C-lobeCaM-CaV1.2 IQ domain interactions (Van Petegem et al., 2005).
Figure 6.
Ca2+/C-lobeBP and Ca2+/C-lobeCaM binding sites overlap on the CaV1.2 IQ domain. Exemplar ITC titrations for: A, 75 μM Ca2+/C-lobeBP into a solution of 7.5 μM IQ domain and 11.25 μM Ca2+/C-lobeCaM. B, 75 μM Ca2+/C-lobeBP into 7.5 μM IQ domain F1628A mutant, 15°C. C, 75 μM Ca2+/C-lobeBP into 7.5 μM IQ domain F1628A mutant, 25°C. D, 75 μM Ca2+/C-lobeBP into wild-type IQ domain, 25°C. E, 75 μM Ca2+/C-lobeBP into 7.5 μM IQ domain F1631A mutant. F, 75 μM Ca2+/C-lobeCaM into 7.5 μM IQ domain F1631A mutant. Cartoons depict the observed binding modes for Ca2+/C-lobeBP (dark blue) and Ca2+/C-lobeCaM (light blue). Stars indicate mutant IQ domains.
ITC experiments with Ca2+/C-lobeBP and the F1628A peptide at 15°C, the temperature of our other ITC analyses, gave a small, endothermic signal (Figure 6B). Therefore, we evaluated Ca2+/C-lobeBP binding to the F1628A peptide at 25°C, where the signal was robust (Figure 6C) and compared the results to titration of Ca2+/C-lobeBP into the wild-type peptide at the same temperature (Figure 6D). The data show that CaV1.2 F1628A reduced Ca2+/C-lobeBP affinity and binding enthalpy by ~48-fold and ~4 kcal mol−1, respectively. The enthalpy reduction magnitude is similar to that caused by F1628A for Ca2+/C-lobeCaM (Van Petegem et al., 2005) and is consistent with a loss of hydrophobic interactions. Measurements with the F1631A peptide (Figure 6E) showed a ~33-fold and ~0.6 kcal mol−1 reduction in affinity and binding enthalpy, respectively. The F1631A change also reduces binding affinity and enthalpy for Ca2+/C-lobeCaM (Figure 6F). Taken together, these experiments strongly support the idea that CaBP1 Ca2+/C-lobe interacts with the CaV1.2 IQ domain in a manner akin to Ca2+/CaM C-lobe (Fallon et al., 2005; Van Petegem et al., 2005).
Interlobe linker Glu94 is essential for CDI inhibition but not CaV1.2 IQ domain binding
Given that the CaBP1 interlobe linker is well ordered, makes specific interactions with the N-lobe, and is crucial to CDI inhibition, we used an alanine scan to investigate the functional contribution of each residue. Each alanine mutation diminished CDI inhibition (Figure 7A and B); however, E94A was the only change that completely eliminated CDI inhibition (Figure 7A and B). E94A also caused the appearance of CDI tachyphylaxis (Figure 7C and D) resulting from changes in recovery from inactivation similar to the DMIG linker mutant (Figure S4). To explore the importance of the Glu94 position further, we tested the effects of E94Q and E94D. E94Q almost completely eliminated CDI inhibition (Fig 7A and B) whereas, E94D allowed partial CDI inhibition. These results indicate that the sidechain carboxylate and precise sidechain geometry are essential. To probe the N-lobe/Glu94 interaction further, we replaced Tyr56, which contributes many N-lobe pocket/Glu94 contacts, with alanine. This change substantially diminishes CDI inhibition (Figure 7E and Table 1) and further supports the importance of the N-lobe/Glu94 interaction. The paramount functional role for Glu94 agrees well with the structurally defined N-lobe/Glu94 interaction (Figure 5E).
Figure 7.
Functional examination of the CaBP1 interlobe linker and E94/N-lobe interaction. A, Representative normalized ICa traces at a test potential of +20 mV for CaV1.2 expressed with CaVβ2a and indicated CaBP1 linker mutants (grey), CaBP1 (orange), or CaM (blue). B, ti300 values for data from ‘A’. C, Relative current changes between the last and first pulse for CaV1.2 expressed with CaVβ2a and CaBP1, CaM, or the indicated mutants. D, Averaged frequency dependent change of ICa amplitude for ‘C’. E, and F, Representative normalized ICa traces at a test potential of +20 mV for CaV1.2 expressed with CaVβ2a and indicated CaBP1 mutants (grey), CaBP1 (orange), or CaM (blue). Graphs compare ti300 values. Parentheses indicate the number of oocytes tested. Error bars indicate s.d. See also Figure S4 and Figure S6.
Despite the effects that E94A had on function, ITC experiments revealed that the loss of the N-lobe/Glu94 interaction caused by E94A altered ΔH and ΔS but spared the affinity for the CaV1.2 IQ domain (Kd = 0.336 ± 0.097 nM, Table 2, Figure S5). This result prompted us to test whether the ordered nature of the linker was a key element of CaBP1 function. We made a mutant (4G) that maintained the N-lobe/Glu94 interaction but that converted the C-terminal half of the linker (residues 97-100) to polyglycine. In contrast to the devastating effect of E94A, 4G retained an ability to inhibit CDI that was on par with the single alanine mutants (Figure 7F). Thus, although both the Glu94/N-lobe interaction and interlobe linker length (Figure 2) are essential for CaBP1 function, the order seen in the C-terminal half is not.
CaBP1 and CaM mediated CDF are two distinct processes
CaM-mediated CaV1.2 CDF requires CaVβ (Findeisen and Minor, 2009; Grueter et al., 2006; Hudmon et al., 2005) and CaMKII (Anderson et al., 1994; Grueter et al., 2006; Hudmon et al., 2005; Yuan and Bers, 1994). Although CaMKII activation is not necessary for CaBP1-mediated CDF (Zhou et al., 2004), the extent to which CaV1.2 CaM-mediated CDF (Van Petegem et al., 2005; Zühlke et al., 1999; Zühlke et al., 2000) and CaBP1-mediated CDF (Zhou et al., 2004) share molecular requirements has remained unclear. To test whether CaBP1-mediated CDF necessitates the presence of CaVβ, we used a CaV1.2 mutant, ‘HotA’, that cannot bind CaVβ (Van Petegem et al., 2008) and that eliminates CaM-mediated CDF (Findeisen and Minor, 2009). Unlike the case with CaM, CaBP1 supports CDF when co-expressed with CaV1.2 HotA or wild-type CaV1.2 in the absence of CaVβ2a (Figure 8A and B). Thus, CaBP1-mediated CDF does not require CaVβ.
Figure 8.
Comparison of CaBP1-mediated and CaM-mediated CaV1.2 CDF. A, Relative current changes between the last and first pulse in response to 40 +20 mV pulses delivered at 3 Hz for CaV1.2 and CaV1.2 mutants expressed with indicated CaVβ2a, CaBP1, and CaM combinations. B, Averaged frequency dependent change of ICa amplitude for ‘A’. C, Representative normalized ICa traces at a test potential of +20 mV for CaV1.2 with CaVβ2a and CaBP1 (orange) or CaM (blue) and CaV1.2 TripleA (grey) expressed with CaVβ2a and CaBP1. Parentheses indicate the number of oocytes tested. Error bars indicate s.d.
CaV1.2 CaM-mediated CDF is unmasked by the CaV1.2 IQ domain mutation, I1624A (Van Petegem et al., 2005; Zühlke et al., 1999; Zühlke et al., 2000) (Figure 8A and B). If CaBP1-mediated CDF were similar to CaM-mediated CDF, one might expect that I1624A would enhance CaBP1-mediated CDF. Contrary to this expectation, co-expression of CaV1.2 I1624A with CaBP1 produces CDF having a magnitude indistinguishable from that seen with CaBP1 and CaV1.2 (Figure 8A and B). CaV1.2 IQ domain residues F1618, Y1619, and F1622 are involved in Ca2+/CaM N-lobe-IQ domain interactions that play a role in CaV1.2 CaM-mediated CDF (Hudmon et al., 2005; Van Petegem et al., 2008). The triple alanine mutant, F1618A/Y1619A/F1622A, (‘TripleA’), eliminates CaM-mediated CDF (Van Petegem et al., 2008). In contrast, TripleA had no effect on CaBP1 mediated CDF (Figure 8A and B) or on CaBP1 CDI inhibition (Figure 8C). The insensitivity of CaBP1-medated CDF to manipulations that affect CaM-mediated CDF demonstrates that CaV1.2 CaM-mediated CDF and CaV1.2 CaBP1-mediated CDF are different and indicates that their underlying molecular mechanisms are different.
Discussion
CaBPs belong to a large calcium sensor family found throughout the nervous system (Burgoyne et al., 2004; Haeseleer et al., 2002; Weiss and Burgoyne, 2002) and closely resemble CaM (Haeseleer et al., 2002; Weiss and Burgoyne, 2002). Accordingly, CaBPs interact with diverse classes of CaM-modulated proteins (Haynes et al., 2006): CaVs (Lee et al., 2002; Yang et al., 2002; Zhou et al., 2004), IP3 Receptors (Kasri et al., 2004; White et al., 2006; Yang et al., 2002), TRP channels (Kinoshita-Kawada et al., 2005), and myosin 1c (Tang et al., 2007). CaBP-target interactions impart functional changes distinct from those caused by CaM (Kasri et al., 2004; Lee et al., 2002; Yang et al., 2002; Zhou et al., 2004) and may diversify neuronal responses to calcium signals. CaBPs interact with and reshape the functional properties of certain CaVs (Cui et al., 2007; Haeseleer et al., 2004; Lee et al., 2002; Tippens and Lee, 2007; Yang et al., 2006; Zhou et al., 2004). Because CaVs are prominent in cellular calcium signaling pathways (Clapham, 2007), CaBP remodeling of CaV activity should reach well beyond electrical excitation properties.
CDI is an important type of CaV feedback modulation that limits cellular calcium entry in response to electrical activity. There are a number of situations in which CDI is overridden (Striessnig, 2007). One mechanism is CaBP1 substitution for CaM within CaV multiprotein complexes in retinal (Haeseleer et al., 2004) and auditory (Cui et al., 2007; Yang et al., 2006) neurons. This component change also happens in the brain (Zhou et al., 2004), blocks CDI in CaV1.2 and CaV1.3 (Cui et al., 2007; Yang et al., 2006; Zhou et al., 2004; Zhou et al., 2005), and introduces CDF to CaV1.2 (Zhou et al., 2004). CaBP1 and CaM have two independently-folded EF-hand containing lobes separated by a linker (Li et al., 2009; Masino et al., 2000; Tsalkova and Privalov, 1985). This likeness presents the question of which elements endow CaBP1 with the ability to alter CaV1.2 behavior differently from CaM. Our chimera-based analysis eliminated a role for CaBP1 N-terminal myristoylation, which is important for CaV2.1 modulation (Few et al., 2005), and identified the N-lobe and interlobe linker as the elements that allow CaBP1 to inhibit CDI and enable CDF in CaV1.2 (Figure 1).
The ability of CaBP1 C-lobe to bind calcium was important but not integral to CDI inhibition. This lack of sensitivity would seem to conflict with the reported calcium-dependency of the CaBP1-CaV1.2 IQ domain interaction (Zhou et al., 2004). However, the ability of CaBP1-EF3, EF4, and EF34 mutants to block CDI is reminiscent of the ability of the CaM EF34 mutant to block CDI (Peterson et al., 1999) and suggests that part of the CaBP1 CDI inhibition mechanism could arise from simple competition with apo-CaM binding to CaV1.2. Nevertheless, this effect cannot embody the entire mechanism. Our data show that only the BBM chimera but not BMM or MBM chimeras or the E94A mutant block CDI. If competition were the only effect, these mutants would be potent CDI inhibitors. Thus, the data strongly suggest that besides CaM competition, there must be an active role for the CaBP1 N-lobe/Glu94 module in CaV1.2 CDI inhibition and this role is the dominant contributor.
The CaBP1 crystal structure revealed an unanticipated N-lobe/Glu94 interaction (Figure 5E) that is indispensible for CaBP1 CDI inhibition and CaBP1-mediated CDF of CaV1.2 (Figure 7). It is notable that this position is conserved CaBP2, CaBP4, and CaBP5 (Figure S2), particularly as CaBP4 inhibits CaV1.3 CDI (Cui et al., 2007; Yang et al., 2006). Thus, in addition to serving as part of the CaBP1 module required for CaV1.2 CDI inhibition, the N-lobe/Glu94 interaction may have a role in other CaBPs.
The CaV1.2 IQ domain is the essential CaM modulation site (Peterson et al., 1999; Pitt et al., 2001) and main CaBP1 binding site (Zhou et al., 2004). Our data uncover a strong structural similarity between the CaM and CaBP1 Ca2+/C-lobes (Figure 5D). This structural mimicry suggests that both bind the CaV1.2 IQ domain in a similar manner. This idea is supported by the interchangeability of CaBP1 and CaM C-lobes with respect to parent molecule CaV1.2 modulation (Figures 1D-G) and is consistent with the similar functional sensitivity of both proteins to CaV1.2 IQ domain mutations (Zhou et al., 2004). Further, our ITC studies reveal that Ca2+/C-lobeBP and Ca2+/C-lobeCaM bind with similar affinities to a mutually exclusive CaV1.2 IQ domain site (Figure 6A, C and D, Table 2) and that mutations of residues that form anchors for Ca2+/C-lobeCaM greatly diminish the ability of Ca2+/C-lobeBP to bind (Figure 6C-F). Together, these results suggest that CaBP1 and CaM C-terminal lobes have the same main function, formation of high affinity attachment to the CaV1.2 IQ domain and that this attachment site overlaps.
ITC experiments with full-length CaBP1 show that it binds the CaV1.2 IQ domain ~40-fold stronger than Ca2+/C-lobeBP and that the CaBP1 E94A mutation, which eliminates CDI inhibition, has no effect on binding affinity. Superposition of the Ca2+/CaBP1 and Ca2+/CaM C-lobes in the context of the Ca2+/CaM-CaV1.2 IQ domain complex (Van Petegem et al., 2005) indicates that the compact CaBP1 conformation observed in the crystal structure is incompatible with CaV1.2 IQ domain binding. Many CaBP1 N-lobe/Glu94 atoms clash with the CaV1.2 IQ helix (Figure S6). Therefore, although there is good evidence that the Ca2+/CaBP1 C-lobe binds the CaV1.2 IQ domain in a mode similar to Ca2+/CaM C-lobe, CaBP1 binding must involve a conformation in which the CaBP1 N-lobe is in different position than seen in the CaBP1 structure.
Our ITC experiments indicate that the CaBP1 N-lobe participates in CaV1.2 IQ domain binding. If CaBP1 Ca2+/C-lobe and CaM Ca2+/C-lobe bind the CaV1.2 IQ domain similarly, one candidate CaBP1 N-lobe interaction site would be the CaV1.2 IQ domain residues that comprise the CaM Ca2+/N-lobe binding site (F1618, Y1619, and F1622) (Fallon et al., 2005; Van Petegem et al., 2005). Nevertheless, functional analysis of the CaV1.2 TripleA mutant, which disrupts these anchors and eliminates CaM-mediated CDF (Van Petegem et al., 2005), reveals that neither CaBP1-mediated CDF or CDI inhibition depend on these residues. Therefore, CaBP1 N-lobe must interact with CaV1.2 differently than Ca2+/CaM N-lobe. Conformational differences between the CaBP1 crystal and solution structures (Li et al., 2009) show that CaBP1 can adopt various interlobe conformations. Further, CaBP1 modulation of CaV1.2 tolerates substitutions that render the C-terminal part of the interlobe linker maximally flexible (Figure 7F). Thus, we propose that the CaBP1 binds the CaV1.2 IQ domain using a similar C-lobe position to CaM but an alternate N-lobe pose such as an open structure that would maintain the CaBP1 N-lobe/Glu94 interaction but flex in the C-terminal portion of the interlobe linker. This could allow CaBP1 to interact with the CaV1.2 cytoplasmic N-terminal domain, a channel element that has the capacity to bind CaBP1 and that is required for CaBP1 modulation of CaV1.2 (Zhou et al., 2005).
CaBP1 affects CaV1s and CaV2s differently. CaBP1 inhibits CaV1.2 and CaV1.3 CDI (Cui et al., 2007; Yang et al., 2006; Zhou et al., 2004; Zhou et al., 2005), whereas it accelerates CaV2.1 CDI and changes activation (Few et al., 2005; Lee et al., 2002). Additionally, CaBP1 blocks CaV2.1 CDF (Few et al., 2005; Lee et al., 2002). These CaV2.1 functional changes require CaBP1 N-terminal myristoylation (Few et al., 2005), a modification that we find irrelevant for CaBP1 modulation of CaV1.2. This result is corroborated the ability of a non-myristyolated CaBP1 splice variant, caldendrin, to block CaV1.2 CDI (Tippens and Lee, 2007). The mechanistic basis for the difference between CaV1.2 and CaV2.1 modulation by CaBP1 remains unresolved. CDI and CDF in CaV1s and CaV2s arise from the actions of different CaM lobes (Dunlap, 2007) that are correlated with different Ca2+/CaM binding modes on the CaV1 (Van Petegem et al., 2005) and CaV2 (Kim et al., 2008) IQ domains (however, cf., (Mori et al., 2008)). Similar to Ca2+/CaM (Kim et al., 2008; Van Petegem et al., 2005), we find that both CaBP1 lobes can bind the IQ domain. It is tempting to speculate that the differential effects of CaBP1 on CaV1 and CaV2 CDI and CDF and the differential effects of the CaM lobes with respect to the same processes are related. CaBP1 also binds other CaM binding channel segments (Dick et al., 2008; Ivanina et al., 2000; Lee et al., 1999). It is possible that the differential effects also require interactions with these or other channel domains. Delineating how CaBPs act should not only unravel CaBP-specific mechanisms, but may help dissect the still imperfectly understood action of CaM with respect to CDI and CDF. Our data provide the framework for investigation of these issues.
We have identified the features that allow CaBP1 to modulate CaV1.2 differently from CaM. Our studies reveal an unanticipated essential role for the interlobe linker and show that CaBP1 has two structural modules that encode two separate functions. CaBP1 C-lobe serves as the IQ domain anchor, whereas the N-lobe/linker module contains the elements required for channel modulation. The elements comprising the N-lobe/Glu94 interaction are conserved in many CaBPs that interact with diverse targets. Thus, the structural underpinnings outlined here are likely to aid understanding how CaBPs act on a wide range of targets that have dual modulation by CaM and CaBPs.
Experimental procedures
Expression and Purification
Individual human CaBP1-S (AF169148) lobes were expressed as fusion proteins from a pET28 vector denoted ‘HMT’ (Van Petegem et al., 2004) in E. coli BL21(DE3)pLysS. CaBP1(Δ2-15) and CaBP1(Δ2-15) K130A, and selenium methionine substituted CaBP1(Δ2-15) K130A were expressed a pEGST vector in E. coli BL21(DE3)pLysS. CaV1.2 IQ domains were made as described previously (Van Petegem et al., 2005). Expression and purification details can be found in Supplementary Experimental Procedures.
Crystallization and Refinement
CaBP1(Δ2-15) was crystallized by hanging drop vapor diffusion at 20°C from equal volumes of protein in buffer A and well solution containing 1.9 M (NH4)2SO4, 2-4 % 1,2-propanediol and 0.1 M citrate/sodium citrate, pH 5.5. Native and selenomethionine-substituted CaBP1(Δ2-15) K130A were crystallized similarly except that 1,2-propanediol was replaced by 5-7 % 1,6-hexanediol. Crystals were flash-frozen in well solution plus 20% glycerol. Diffraction data were collected at Beamline 8.3.1 (Advanced Light Source, Lawrence Berkeley National Laboratories) and processed using MOSFLM 7.0.4 (Leslie, 1992). A three-wavelength MAD experiment was performed on crystals of SeMet-substituted CaBP1(Δ2-15) K130A. 64 initial selenium positions were located using SHELXD (Uson and Sheldrick, 1999). After phasing and density modification using DM, the resulting electron density allowed building of an initial model that was improved by iterative cycles of manual building and refinement against native data using Refmac5 (Collaborative Computational Project, 1994). TLS-tensors were used throughout refinement. Tight NCS restraints were employed initially and relaxed in later refinement cycles. Each final model consists of six CaBP1 monomers.
Isothermal titration calorimetry
Titrations were performed at 15°C or 25°C using a VP-ITC Microcalorimeter (MicroCal). Samples were dialyzed overnight at 4°C (Slide-A-Lyzer, 3.5 kDa molecular weight cut-off, Thermo Scientific) against 5 mM KCl, 1 mM CaCl2, 10 mM Hepes, pH 7.4. After centrifugation (30′, 70,000 × g, 4 °C) protein concentrations were determined by absorbance at 280 nm (Edelhoch, 1967). All samples were degassed for 5− prior to loading into the VP-ITC. CaV1.2 IQ domain peptide at concentrations of 5.5 or 7.5 μM were titrated with 75 μM CaBP1 constructs using a schedule of one 4 μl titrant injection followed by 29 injections of 10 μl. For competition experiments, 7.5 μM CaV1.2 IQ domain peptide was preincubated with 50 μM of Ca2+/N-lobeBP and titrated with a solution of 75 μM of CaBP1. To correct the baseline, either heat of dilution from titrations of injectant into buffer was subtracted (Figures 3A and 3D) or the final titration points were used to estimate the baseline (Figures 3B, 3F-H). Data were processed with MicroCal Origin 7.0 using a single site binding model or a competitive binding model, where appropriate.
For binding isotherm analysis of the simultaneous addition of Ca2+/N-lobeBP and Ca2+/C-lobeBP to the CaV1.2 IQ domain, we used a modified version of competition ligand binding by displacement ITC (Sigurskjold, 2000) that takes into account both ligands are added simultaneously. Thus, the stoichiometric concentration of B in the calorimeter cell, equation [26] from (Sigurskjold, 2000), [B]0,i = ft,i[B]0,, is replaced by [B]0,i = [B]2 (1-ft,i), where [B]s is concentration of Ca2+/N-lobeBP in the injection syringe and ft,i is the total dilution factor after the ith injection, respectively.
Electrophysiology
Human CaV1.2 (α1C77, CAA84346) in pcDNA3.1 (Invitrogen), rat CaVβ2a (NP 446303) in pGEMHE, and CaVα2δ-1 (NM_00182276) in pcDNA3.1 were co-expressed with either Homo sapiens CaM (NM_006888) or the short isoform of human CaBP1-S (AF169148) in pGEMHE for Xenopus oocyte two-electrode voltage clamp experiments. Point mutations were introduced using QuikChange (Stratagene). For chimera construction, CaM and CaBP1 were divided into three elements: N-lobe (CaM 1-78 and CaBP1 1-92), linker (CaM 79-82 and CaBP1 93-100 for CaBP1), and C-lobe (CaM 83-149 and CaBP1 101-167). All chimeras having CaBP1 N-lobe include the N-terminal myristoylation domain. Inter-element junctions were framed by introduction of EcoRV sites generated using QuikChange. Elements were excised using EcoRV and either HindIII or XhoI present in the parent vector and combined to produce the chimeras. Following chimera generation, junction sites were returned to the native sequence using QuikChange.
After overnight linearization (XhoI for pcDNA3.1 and NheI for pGEMHE), capped mRNAs were synthesized (T7 mMessenger kit (Ambion)) according to the manufacturer’s protocols. RNA concentrations were determined by A260nm. 50 nl of CaVα1, CaVβ, CaVα2δ-1, and CaBP1 or CaM mRNA at a molar ratio of 1:1:1:20 were injected into defolliculated stage VI Xenopus oocytes using Nanoject II injector (Drummond Scientific). Oocytes were kept at 18 °C in ND96 medium containing penicillin and streptomycin. Two-electrode voltage-clamp experiments were performed 2 to 4 days post-injection using a GeneClamp 500B (Axon Instruments) amplifier controlled by a 1,200 MHz processor computer (Celeron, Gateway) running CLAMPEX 8.2.0.244, and digitized at 1kHz with a Digidata 1332A (Axon Instruments). Immediately prior to recording, oocytes were injected with 47 nl of 100 mM BAPTA to minimize Ca2+-activated Cl− current. Recording solutions contained either 40 mM Ba(OH)2 or 40 mM Ca(NO3)2, 50 mM NaOH, 1 mM KOH, and 10 mM HEPES, adjusted to pH 7.4 using HNO3. Electrodes were filled with 3M KCl and had 0.3-2.0 MΩ resistances. Leak currents were subtracted using a P/4 protocol. Currents were analyzed with Clampfit 8.2 (Axon Instruments). Oocytes were superfused during recording using a Valvelink 16 controller (Automate Scientific). Holding potential for all experiments was −90mV. Results are from at least two independent oocyte batches. ti300 values were calculated from normalized currents at +20 mV and represent the percentage inactivation after 300 milliseconds. Inactivation τ values were calculated at a test potential of +20 mV as described (Findeisen and Minor, 2009). CDF was elicited by a train of 40, 50 ms pulses to +20 mV at 3Hz and calculated as the ratio of the peak current from the last pulse divided by the first. Recovery from inactivation was measured by a protocol with two 450 ms pulses to +20 mV separated by variable time intervals. Consecutive sweeps were separated by 30 s.
Pull-down experiments
Bacterial pellets from a 50 ml culture of CaBP1, CaBP1 mutants, CaM, or CaBP1/CaM chimeras co-expressed with HMT-tagged CaV1.2 IQ domain were resuspended in 1.6 ml lysis buffer (10% sucrose, 150 mM KCl, 5 mM MgCl2, 1 mM CaCl2, 25 μg/ml DNAaseI, 1 mM PMSF, 100 mM Tris, pH 8.8) and lysed by sonication. 100 μl amylose resin in buffer AmyA (250 mM KCl, 1 mM CaCl2, 10 mM Hepes/KOH, pH 7.4) was incubated with the bacterial lysate for 30 minutes at 4 °C with gentle agitation. After four washes with 500 μl AmyA, bound material was eluted in 500 μl AmyA containing 10 mM maltose. Input and eluate fractions were analyzed by SDS-PAGE.
Supplementary Material
Acknowledgments
This work was supported by grants to DLM (NIH-NHLBI R01-HL080050 and the American Heart Association 0740019N) and to FF from the American Heart Association. We thank A. Lee for the CaBP1 clone; D. Palanivelu, A. Tolia, and F. Van Petegem for manuscript comments; J. Holton at ALS Beamline 8.3.1 for data collection assistance. DLM is an AHA Established Investigator. CaBP1 and CaBP1 K130A coordinates and structure factors are deposited with the RCSB under codes 3OX5 and 3OX6, respectively.
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.
References
- Alseikhan BA, DeMaria CD, Colecraft HM, Yue DT. Engineered calmodulins reveal the unexpected eminence of Ca2+ channel inactivation in controlling heart excitation. Proc Natl Acad Sci U S A. 2002;99:17185–17190. doi: 10.1073/pnas.262372999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anderson ME, Braun AP, Schulman H, Premack BA. Multifunctional Ca2+/calmodulin-dependent protein kinase mediates Ca(2+)-induced enhancement of the L-type Ca2+ current in rabbit ventricular myocytes. Circ Res. 1994;75:854–861. doi: 10.1161/01.res.75.5.854. [DOI] [PubMed] [Google Scholar]
- Burgoyne RD, O’Callaghan DW, Hasdemir B, Haynes LP, Tepikin AV. Neuronal Ca2+-sensor proteins: multitalented regulators of neuronal function. Trends Neurosci. 2004;27:203–209. doi: 10.1016/j.tins.2004.01.010. [DOI] [PubMed] [Google Scholar]
- Catterall WA. Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol. 2000;16:521–555. doi: 10.1146/annurev.cellbio.16.1.521. [DOI] [PubMed] [Google Scholar]
- Clapham DE. Calcium signaling. Cell. 2007;131:1047–1058. doi: 10.1016/j.cell.2007.11.028. [DOI] [PubMed] [Google Scholar]
- Collaborative Computational Project, N The CCP4 suite: Programs for protein crystallography. Acta Crystallogr D Biol Crystallogr. 1994;50:760–763. doi: 10.1107/S0907444994003112. [DOI] [PubMed] [Google Scholar]
- Cui G, Meyer AC, Calin-Jageman I, Neef J, Haeseleer F, Moser T, Lee A. Ca2+-binding proteins tune Ca2+-feedback to Cav1.3 channels in mouse auditory hair cells. J Physiol. 2007;585:791–803. doi: 10.1113/jphysiol.2007.142307. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davies A, Hendrich J, Van Minh AT, Wratten J, Douglas L, Dolphin AC. Functional biology of the alpha(2)delta subunits of voltage-gated calcium channels. Trends Pharmacol Sci. 2007;28:220–228. doi: 10.1016/j.tips.2007.03.005. [DOI] [PubMed] [Google Scholar]
- Derewenda ZS, Vekilov PG. Entropy and surface engineering in protein crystallization. Acta Crystallogr D Biol Crystallogr. 2006;62:116–124. doi: 10.1107/S0907444905035237. [DOI] [PubMed] [Google Scholar]
- 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]
- Dolphin AC. beta subunits of voltage-gated calcium channels. J Bioenerg Biomembr. 2003;35:599–620. doi: 10.1023/b:jobb.0000008026.37790.5a. [DOI] [PubMed] [Google Scholar]
- Dunlap K. Calcium channels are models of self-control. J Gen Physiol. 2007;129:379–383. doi: 10.1085/jgp.200709786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Edelhoch H. Spectroscopic determination of tryptophan and tyrosine in proteins. Biochemistry. 1967;6:1948–1954. doi: 10.1021/bi00859a010. [DOI] [PubMed] [Google Scholar]
- Erickson MG, Liang H, Mori MX, Yue DT. FRET two-hybrid mapping reveals function and location of L-type Ca2+ channel CaM preassociation. Neuron. 2003;39:97–107. doi: 10.1016/s0896-6273(03)00395-7. [DOI] [PubMed] [Google Scholar]
- Fallon JL, Halling DB, Hamilton SL, Quiocho FA. Structure of calmodulin bound to the hydrophobic IQ domain of the cardiac Ca(v)1.2 calcium channel. Structure. 2005;13:1881–1886. doi: 10.1016/j.str.2005.09.021. [DOI] [PubMed] [Google Scholar]
- Fallon JL, Quiocho FA. A closed compact structure of native Ca(2+)-calmodulin. Structure. 2003;11:1303–1307. doi: 10.1016/j.str.2003.09.004. [DOI] [PubMed] [Google Scholar]
- Few AP, Lautermilch NJ, Westenbroek RE, Scheuer T, Catterall WA. Differential regulation of CaV2.1 channels by calcium-binding protein 1 and visinin-like protein-2 requires N-terminal myristoylation. J Neurosci. 2005;25:7071–7080. doi: 10.1523/JNEUROSCI.0452-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Findeisen F, Minor DL., Jr. Disruption of the IS6-AID Linker Affects Voltage-gated Calcium Channel Inactivation and Facilitation. J Gen Physiol. 2009;133:327–343. doi: 10.1085/jgp.200810143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gifford JL, Walsh MP, Vogel HJ. Structures and metal-ion-binding properties of the Ca2+-binding helix-loop-helix EF-hand motifs. Biochem J. 2007;405:199–221. doi: 10.1042/BJ20070255. [DOI] [PubMed] [Google Scholar]
- Grabarek Z. Structural basis for diversity of the EF-hand calcium-binding proteins. J Mol Biol. 2006;359:509–525. doi: 10.1016/j.jmb.2006.03.066. [DOI] [PubMed] [Google Scholar]
- Grueter CE, Abiria SA, Dzhura I, Wu Y, Ham AJ, Mohler PJ, Anderson ME, Colbran RJ. L-type Ca2+ channel facilitation mediated by phosphorylation of the beta subunit by CaMKII. Mol Cell. 2006;23:641–650. doi: 10.1016/j.molcel.2006.07.006. [DOI] [PubMed] [Google Scholar]
- Haeseleer F, Imanishi Y, Maeda T, Possin DE, Maeda A, Lee A, Rieke F, Palczewski K. Essential role of Ca2+-binding protein 4, a Cav1.4 channel regulator, in photoreceptor synaptic function. Nat Neurosci. 2004;7:1079–1087. doi: 10.1038/nn1320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haeseleer F, Imanishi Y, Sokal I, Filipek S, Palczewski K. Calcium-binding proteins: intracellular sensors from the calmodulin superfamily. Biochem Biophys Res Commun. 2002;290:615–623. doi: 10.1006/bbrc.2001.6228. [DOI] [PubMed] [Google Scholar]
- Haeseleer F, Sokal I, Verlinde CL, Erdjument-Bromage H, Tempst P, Pronin AN, Benovic JL, Fariss RN, Palczewski K. Five members of a novel Ca(2+)-binding protein (CABP) subfamily with similarity to calmodulin. J Biol Chem. 2000;275:1247–1260. doi: 10.1074/jbc.275.2.1247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haynes LP, Fitzgerald DJ, Wareing B, O’Callaghan DW, Morgan A, Burgoyne RD. Analysis of the interacting partners of the neuronal calcium-binding proteins L-CaBP1, hippocalcin, NCS-1 and neurocalcin delta. Proteomics. 2006;6:1822–1832. doi: 10.1002/pmic.200500489. [DOI] [PubMed] [Google Scholar]
- Hendrickson WA, Ogata CM. Phase Determination from Multiwavelength Anomolous Diffraction Measurements. Meth. Enzymol. 1997;276:494–523. doi: 10.1016/S0076-6879(97)76074-9. [DOI] [PubMed] [Google Scholar]
- Hudmon A, Schulman H, Kim J, Maltez JM, Tsien RW, Pitt GS. CaMKII tethers to L-type Ca2+ channels, establishing a local and dedicated integrator of Ca2+ signals for facilitation. J Cell Biol. 2005;171:537–547. doi: 10.1083/jcb.200505155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ivanina T, Blumenstein Y, Shistik E, Barzilai R, Dascal N. Modulation of L-type Ca2+ channels by gbeta gamma and calmodulin via interactions with N and C termini of alpha 1C. J Biol Chem. 2000;275:39846–39854. doi: 10.1074/jbc.M005881200. [DOI] [PubMed] [Google Scholar]
- Kasri NN, Holmes AM, Bultynck G, Parys JB, Bootman MD, Rietdorf K, Missiaen L, McDonald F, De Smedt H, Conway SJ, et al. Regulation of InsP3 receptor activity by neuronal Ca2+-binding proteins. Embo J. 2004;23:312–321. doi: 10.1038/sj.emboj.7600037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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]
- Kinoshita-Kawada M, Tang J, Xiao R, Kaneko S, Foskett JK, Zhu MX. Inhibition of TRPC5 channels by Ca2+-binding protein 1 in Xenopus oocytes. Pflugers Arch. 2005;450:345–354. doi: 10.1007/s00424-005-1419-1. [DOI] [PubMed] [Google Scholar]
- Lautermilch NJ, Few AP, Scheuer T, Catterall WA. Modulation of CaV2.1 channels by the neuronal calcium-binding protein visinin-like protein-2. J Neurosci. 2005;25:7062–7070. doi: 10.1523/JNEUROSCI.0447-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee A, Westenbroek RE, Haeseleer F, Palczewski K, Scheuer T, Catterall WA. Differential modulation of Ca(v)2.1 channels by calmodulin and Ca2+-binding protein 1. Nat Neurosci. 2002;5:210–217. doi: 10.1038/nn805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee A, Wong ST, Gallagher D, Li B, Storm DR, Scheuer T, Catterall WA. Ca2+/calmodulin binds to and modulates P/Q-type calcium channels. Nature. 1999;399:155–159. doi: 10.1038/20194. [DOI] [PubMed] [Google Scholar]
- Leslie AGW. Recent changes to the MOSFLM package for processing film and image plate data. Joint CCP4 + ESF-EAMCB Newsletter of Protein Crystallography. 1992;26 [Google Scholar]
- Li C, Chan J, Haeseleer F, Mikoshiba K, Palczewski K, Ikura M, Ames JB. Structural insights into Ca2+-dependent regulation of inositol 1,4,5-trisphosphate receptors by CaBP1. J Biol Chem. 2009;284:2472–2481. doi: 10.1074/jbc.M806513200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Masino L, Martin SR, Bayley PM. Ligand binding and thermodynamic stability of a multidomain protein, calmodulin. Protein Sci. 2000;9:1519–1529. doi: 10.1110/ps.9.8.1519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mori MX, Vander Kooi CW, Leahy DJ, Yue DT. Crystal Structure of the Ca(V)2 IQ Domain in Complex with Ca(2+)/Calmodulin: High-Resolution Mechanistic Implications for Channel Regulation by Ca(2+) Structure. 2008;16:607–620. doi: 10.1016/j.str.2008.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peterson BZ, DeMaria CD, Adelman JP, Yue DT. Calmodulin is the Ca2+ sensor for Ca2+ -dependent inactivation of L-type calcium channels. Neuron. 1999;22:549–558. doi: 10.1016/s0896-6273(00)80709-6. [DOI] [PubMed] [Google Scholar]
- Pitt GS. Calmodulin and CaMKII as molecular switches for cardiac ion channels. Cardiovasc Res. 2007;73:641–647. doi: 10.1016/j.cardiores.2006.10.019. [DOI] [PubMed] [Google Scholar]
- Pitt GS, Zühlke RD, Hudmon A, Schulman H, Reuter H, Tsien RW. Molecular basis of calmodulin tethering and Ca2+-dependent inactivation of L-type Ca2+ channels. J Biol Chem. 2001;276:30794–30802. doi: 10.1074/jbc.M104959200. [DOI] [PubMed] [Google Scholar]
- Rocque WJ, McWherter CA, Wood DC, Gordon JI. A comparative analysis of the kinetic mechanism and peptide substrate specificity of human and Saccharomyces cerevisiae myristoyl-CoA:protein N-myristoyltransferase. J Biol Chem. 1993;268:9964–9971. [PubMed] [Google Scholar]
- Sigurskjold BW. Exact analysis of competition ligand binding by displacement isothermal titration calorimetry. Anal Biochem. 2000;277:260–266. doi: 10.1006/abio.1999.4402. [DOI] [PubMed] [Google Scholar]
- Striessnig J. C-terminal tailoring of L-type calcium channel function. J Physiol. 2007;585:643–644. doi: 10.1113/jphysiol.2007.147140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang N, Lin T, Yang J, Foskett JK, Ostap EM. CIB1 and CaBP1 bind to the myo1c regulatory domain. J Muscle Res Cell Motil. 2007;28:285–291. doi: 10.1007/s10974-007-9124-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tippens AL, Lee A. Caldendrin, a neuron-specific modulator of Cav/1.2 (L-type) Ca2+ channels. J Biol Chem. 2007;282:8464–8473. doi: 10.1074/jbc.M611384200. [DOI] [PubMed] [Google Scholar]
- Tsalkova TN, Privalov PL. Thermodynamic study of domain organization in troponin C and calmodulin. J Mol Biol. 1985;181:533–544. doi: 10.1016/0022-2836(85)90425-5. [DOI] [PubMed] [Google Scholar]
- Uson I, Sheldrick GM. Advances in direct methods for protein crystallography. Curr Opin Struct Biol. 1999;9:643–648. doi: 10.1016/s0959-440x(99)00020-2. [DOI] [PubMed] [Google Scholar]
- Van Petegem F, Chatelain FC, Minor DL., Jr. Insights into voltage-gated calcium channel regulation from the structure of the CaV1.2 IQ domain-Ca2+/calmodulin complex. Nat Struct Mol Biol. 2005;12:1108–1115. doi: 10.1038/nsmb1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Petegem F, Clark KA, Chatelain FC, Minor DL., Jr. Structure of a complex between a voltage-gated calcium channel beta-subunit and an alpha-subunit domain. Nature. 2004;429:671–675. doi: 10.1038/nature02588. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Petegem F, Duderstadt KE, Clark KA, Wang M, Minor DL., Jr. Alanine-Scanning Mutagenesis Defines a Conserved Energetic Hotspot in the Ca(V)alpha(1) AID-Ca(V)beta Interaction Site that Is Critical for Channel Modulation. Structure. 2008;16:280–294. doi: 10.1016/j.str.2007.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Petegem F, Minor DL. The structural biology of voltage-gated calcium channel function and regulation. Biochem Soc Trans. 2006;34:887–893. doi: 10.1042/BST0340887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weiss JL, Burgoyne RD. Sense and sensibility in the regulation of voltage-gated Ca(2+) channels. Trends Neurosci. 2002;25:489–491. doi: 10.1016/s0166-2236(02)02247-6. [DOI] [PubMed] [Google Scholar]
- White C, Yang J, Monteiro MJ, Foskett JK. CIB1, a ubiquitously expressed Ca2+-binding protein ligand of the InsP3 receptor Ca2+ release channel. J Biol Chem. 2006;281:20825–20833. doi: 10.1074/jbc.M602175200. [DOI] [PubMed] [Google Scholar]
- Wilson MA, Brunger AT. The 1.0 A crystal structure of Ca(2+)-bound calmodulin: an analysis of disorder and implications for functionally relevant plasticity. J Mol Biol. 2000;301:1237–1256. doi: 10.1006/jmbi.2000.4029. [DOI] [PubMed] [Google Scholar]
- Wingard JN, Chan J, Bosanac I, Haeseleer F, Palczewski K, Ikura M, Ames JB. Structural analysis of Mg2+ and Ca2+ binding to CaBP1, a neuron-specific regulator of calcium channels. J Biol Chem. 2005;280:37461–37470. doi: 10.1074/jbc.M508541200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang J, McBride S, Mak DO, Vardi N, Palczewski K, Haeseleer F, Foskett JK. Identification of a family of calcium sensors as protein ligands of inositol trisphosphate receptor Ca(2+) release channels. Proc Natl Acad Sci U S A. 2002;99:7711–7716. doi: 10.1073/pnas.102006299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang PS, Alseikhan BA, Hiel H, Grant L, Mori MX, Yang W, Fuchs PA, Yue DT. Switching of Ca2+-dependent inactivation of Ca(v)1.3 channels by calcium binding proteins of auditory hair cells. J Neurosci. 2006;26:10677–10689. doi: 10.1523/JNEUROSCI.3236-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yuan W, Bers DM. Ca-dependent facilitation of cardiac Ca current is due to Ca-calmodulin-dependent protein kinase. Am J Physiol. 1994;267:H982–993. doi: 10.1152/ajpheart.1994.267.3.H982. [DOI] [PubMed] [Google Scholar]
- Zhou H, Kim SA, Kirk EA, Tippens AL, Sun H, Haeseleer F, Lee A. Ca2+-binding protein-1 facilitates and forms a postsynaptic complex with Cav1.2 (L-type) Ca2+ channels. J Neurosci. 2004;24:4698–4708. doi: 10.1523/JNEUROSCI.5523-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. J Biol Chem. 2005;280:29612–29619. doi: 10.1074/jbc.M504167200. [DOI] [PubMed] [Google Scholar]
- Zühlke RD, Pitt GS, Deisseroth K, Tsien RW, Reuter H. Calmodulin supports both inactivation and facilitation of L-type calcium channels. Nature. 1999;399:159–162. doi: 10.1038/20200. [DOI] [PubMed] [Google Scholar]
- Zühlke RD, Pitt GS, Tsien RW, Reuter H. Ca2+-sensitive inactivation and facilitation of L-type Ca2+ channels both depend on specific amino acid residues in a consensus calmodulin-binding motif in the(alpha)1C subunit. J Biol Chem. 2000;275:21121–21129. doi: 10.1074/jbc.M002986200. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.








