Abstract
Na+/Ca2+ exchangers (NCXs) are secondary active transporters that couple the translocation of Na+ with the transport of Ca2+ in the opposite direction. The exchanger is an essential Ca2+ extrusion mechanism in excitable cells. It consists of a transmembrane domain and a large intracellular loop that contains two Ca2+-binding domains, CBD1 and CBD2. The two CBDs are adjacent to each other and form a two-domain Ca2+ sensor called CBD12. Binding of intracellular Ca2+ to CBD12 activates the NCX but inhibits the NCX of Drosophila, CALX. NMR spectroscopy and SAXS studies showed that CALX and NCX CBD12 constructs display significant interdomain flexibility in the apo state but assume rigid interdomain arrangements in the Ca2+-bound state. However, detailed structure information on CBD12 in the apo state is missing. Structural characterization of proteins formed by two or more domains connected by flexible linkers is notoriously challenging and requires the combination of orthogonal information from multiple sources. As an attempt to characterize the conformational ensemble of CALX-CBD12 in the apo state, we applied molecular dynamics (MD) simulations, NMR (1H-15N residual dipolar couplings), and small-angle x-ray scattering (SAXS) data in a combined strategy to select an ensemble of conformations in agreement with the experimental data. This joint approach demonstrated that CALX-CBD12 preferentially samples closed conformations, whereas the wide-open interdomain arrangement characteristic of the Ca2+-bound state is less frequently sampled. These results are consistent with the view that Ca2+ binding shifts the CBD12 conformational ensemble toward extended conformers, which could be a key step in the NCXs’ allosteric regulation mechanism. This strategy, combining MD with NMR and SAXS, provides a powerful approach to select ensembles of conformations that could be applied to other flexible multidomain systems.
Significance
The conformational ensemble of CALX-CBD12, the main Ca2+ sensor of Drosophila’s Na+/Ca2+ exchanger, was characterized by a combination of MD simulations with small-angle x-ray scattering and NMR data using the ensemble optimization method approach. This analysis showed that this two-domain construct experiences opening-closing motions, providing molecular information about CALX-CBD12 in the apo state. Ca2+ binding shifts the conformational ensemble toward extended conformers. These findings are consistent with a model according to which Ca2+ modulation of CBD12 plasticity is a key step in the Ca2+-regulation mechanism of the full-length exchanger.
Introduction
High-resolution NMR spectroscopy is a well-established method to investigate structures of small proteins or protein domains in solution at atomic resolution. The major type of structural information obtained from NMR is short 1H-1H internuclear distances (<6 Å) derived from the analysis of nuclear Overhauser effects in NOESY experiments (1,2). Additional sources of structural information, such as residual dipolar couplings (RDCs) measured on weakly aligned samples, rotational diffusion anisotropy determined from the analysis of 15N R2 and R1 relaxation rates, and paramagnetic relaxation enhancements measured in samples labeled with a paramagnetic center, are particularly useful to the structural characterization of complexes and multidomain proteins in solution (3, 4, 5, 6, 7, 8). Small-angle x-ray scattering (SAXS), on the other hand, is a low-resolution method and cannot provide information at atomic resolution (9). From the SAXS scattering profile, it is possible to extract accurate structural parameters such as the average radius of gyration (Rg), the degree of protein folding (Kratky plot), and the interdomain orientation (“finger print”) from the pair-distance distribution function, p(r), which is characteristic of the particle’s shape (10). Therefore, the information derived from SAXS is quite complementary to that obtained from NMR because they provide information in different length scales. Indeed, there are several examples of the simultaneous application of SAXS and NMR data to resolve the structure of large protein-protein complexes or multidomain proteins (11, 12, 13, 14, 15, 16, 17).
Protein motions are essential for function (7,18,19). However, the structural characterization of flexible modular proteins imposes additional challenges. Under these conditions, a unique molecular envelope is not appropriate to describe SAXS data because different protein conformations may exist in solution (10,20). The same is valid for NMR restraints such as RDCs, which are subjected to dynamic averaging because of motions up to the millisecond timescale (21,22). In these cases, one possible approach is to build a large ensemble of different domain orientations representative of the protein dynamics and use the experimental data as a refinement filter of the ensemble (23,24). In this strategy, the principal concern is to avoid overfitting, especially when an ensemble of structures is selected based on a single type of experimental data (23,25,26). The complementarity of NMR and SAXS data can be explored to break the degeneracy of the fitted parameters and prevent overfitting (26,27). Indeed, efforts have been made to calculate structural ensembles of multidomain proteins when different interdomain orientations coexist in solution (5,28), and many of them explore the synergy between NMR and SAXS data (16,27,29, 30, 31, 32). One of the challenges in these approaches is to determine the right ensemble size and to find the appropriate fractions of each model within the ensemble. A minimal representative ensemble may be obtained by optimization of the number of representative models and statistical weights using Bayesian inference (33,34) or genetic algorithms (20,35,36). Alternatively, large ensembles may be calculated using clustering strategies followed by maximal entropy refinement of the statistical weights of each cluster (37).
Na+/Ca2+ exchangers (NCXs) are secondary active transporters that couple the import of Na+ to the extrusion of intracellular Ca2+. The NCXs are essential for the maintenance of intracellular Ca2+ homeostasis, especially in excitable cells (38). They consist of a transmembrane domain, involved in ion translocation across the lipid bilayer, and a large cytosolic loop that is responsible for regulation of the exchanger activity by its substrates (39, 40, 41). The large intracellular loop contains a Ca2+ sensor, CBD12, which consists of two adjacent Ca2+-binding domains, CBD1 and CBD2, separated by a highly conserved linker of only three amino acids (Fig. 1) (42,43). The CBDs are β-sandwich motifs, whose Ca2+-binding sites are located at the interstrand distal loops (44, 45, 46, 47). Therefore, the Ca2+-binding sites of CBD1 are located near the interface with CBD2 (Fig. 1).
CALX-CBD12 is the cytosolic Ca2+ sensor of the Na+/Ca2+ exchanger from Drosophila, CALX (48). Whereas Ca2+ binding to CBD12 activates the mammalian exchangers, CALX is unusual because Ca2+ binding to CBD12 inhibits this exchanger (49). Studies of NMR spectroscopy of the CALX and the NCX-CBD12 constructs showed that the two CBDs are flexibly linked to each other in the apo state, whereas binding of Ca2+ to CBD1 stabilizes a rigid and elongated interdomain arrangement between CBD1 and CBD2 (50,51). The flexibility of different NCX-CBD12 isoforms (NCX1-CBD12, NCX2-CBD12, NCX3-CBD12-AC, and NCX3-CBD12-B) was characterized by SAXS using the ensemble optimization method (EOM) (20,43,52). It was found that a dynamic regime with a wide Rg distribution describes the apo state SAXS profiles, whereas a narrower Rg distribution was selected in the presence of Ca2+. These observations are consistent with the greater interdomain flexibility of the apo state as indicated by NMR (50,51). The same analysis has not been reported for CALX-CBD12.
Despite SAXS and NMR data indicating that CBD12 displays interdomain flexibility (50,51), an atomistic view of the dynamics of the apo state, i.e., the various conformations and interdomain arrangements that CBD12 adopts in solution, is still missing. Considering that an ensemble description of CBD12 in the apo state could give additional insights on the Ca2+ allosteric regulation mechanism of the CALX and the NCX exchangers, this question was addressed in this work using molecular dynamics (MD) simulations to generate an initial CALX-CBD12 ensemble and the EOM approach (20) to select a subensemble that agrees with SAXS and RDC data. In the absence of additional experimental data, it is hard to tell that the selected conformations correspond to the native ensemble; however, they do offer insights on large-scale motions experienced by CALX-CBD12 before Ca2+-binding. The strategy proposed here, combining MD simulations with SAXS and RDC data, might be useful for the analysis of other multidomain systems that also experience significant interdomain flexibility.
Materials and methods
CALX-CBD12 expression and purification
A DNA fragment corresponding to CALX1.1 residues 434–697 was PCR amplified from the CALX1.1 cDNA and cloned in fusion with a N-terminal six-histidine tag in the pAE vector (53) using XhoI and EcoRI restriction sites. Protein expression was performed in Escherichia coli BL21-CodonPlus(DE3)-RIL (Agilent, Santa Clara, CA). Bacterial cell cultures were grown in LB at 37°C to an OD600 of 0.6; then the temperature was reduced to 18°C, and protein expression was induced by the addition of 0.4 mM of isopropyl-β-D-thiogalactopyranosid for 18 h. Cells were harvested by centrifugation, and cell pellets stored at −20°C. Approximately 10 g of cells were suspended in 200 mL of lysis buffer (20 mM Tris (pH 7.0), 200 mM NaCl, 5 mM β-mercaptoetanol, 0.1 μg/mL of pepstatin and aprotinin, 1.0 mM phenylmethanesulfonyl fluoride, and 1 mg/mL lysozyme), and subsequently lysed by sonication using a VCX 500 (Sonics, Newton, CT) instrument. The cell lysate was clarified by centrifugation at 21,000 × g for 1 h, the supernatant was applied into a 5 mL HisTrap affinity column (GE Healthcare, Chicago, IL) pre-equilibrated with buffer A (20 mM sodium phosphate (pH 7.4), 300 mM NaCl, 5 mM β-mercaptoethanol, 10 mM imidazole), washed with 50 mL of buffer A, and eluted during a gradient of 0.01 up to 0.5 M of imidazole. Fractions containing CALX-CBD12 were combined, concentrated to 2 mL using an Amicon concentrator with a 3 kDa cutoff (Millipore, Burlington, MA), and applied into a Superdex 75 (16/600) gel filtration column (GE Healthcare) pre-equilibrated with the gel filtration buffer (20 mM Tris (pH 8.0), 200 mM NaCl, 5 mM β-mercaptoethanol, 1% (V/V) glycerol). Fractions containing CALX-CBD12 were combined and applied into a MonoQ 10/100 (GE Healthcare) ion-exchange column pre-equilibrated with the gel filtration buffer and eluted in a gradient of 1.0 M of NaCl. Subsequently, fractions containing CALX-CBD12 were combined and buffer exchanged to 20 mM Tris (pH 7.4), containing 5 mM of β-mercaptoethanol, 200 mM of NaCl, 1% (V/V) of glycerol, and 20 mM of EDTA. To prepare the SAXS samples, the EDTA was removed by buffer exchange using a 3 kDa Amicon centrifugation device. In the case of the Ca2+-bound sample, a suitable aliquot of a CaCl2 stock solution prepared in the same buffer as the protein was added just before the SAXS measurements. Protein concentration was determined by measuring the absorbance at 280 nm assuming ε280 = 25,900 M1 cm1.
SAXS experiment
SAXS experiments were performed on Xenocs-Xeuss laboratory SAXS equipment (Sassenage, France) placed at the Institute of Physics of the University of São Paulo. This machine uses a microfocus source GENIX 3D with CuKα radiation, wavelength λ = 1.5418 Å, Fox3D focusing mirrors, and two sets of scatterless slits (Xenocs 2.0) for beam collimation. The scattering x ray resulting from the sample-primary beam interaction was recorded on a two-dimensional photon-counting Pilatus 300k detector at a sample-to-detector distance of 0.83 m. The scattering intensity was described as a function of the momentum transfer vector modulus (q) defined by q = 4 πsinθ/λ, where 2 θ is the scattering angle. In the used setup, the typical q range was 0.01 < q < 0.4 Å−1. SAXS data were collected on samples of CALX-CBD12 at a concentration of 5.3 mg ⋅ mL−1 (158 μM) in the absence (apo state) or in the presence of CaCl2 at a molar ratio of 6:1 (Ca2+/CBD12). Under these conditions, the four Ca2+-binding sites in the CBD1 domain must be fully occupied because the total protein concentration is significantly greater than the CALX-CBD12 Ca2+ dissociation constant measured by calorimetry, Kd = 1.6 μM (42). CALX-CBD12 samples were injected in a Xenocs low-noise flow cell, and 900 s of exposures were recorded, generating a total of 20 frames. Pairwise comparison of the collected frames indicated no radiation damage on the samples. The same arrangement was performed for an identically prepared sample but without the protein as a background and a scattering reference (plain water) sample to obtain the final intensity in an absolute scale. The subtraction of the blank, normalization, and absolute scaling were performed over all data using the SUPERSAXS package (C.L.P.O. and J. S. Pedersen, unpublished data). After that, all frames were averaged to obtain the final SAXS profiles. Fitting of the SAXS profile was carried out with GNOM (54) using the indirect Fourier transform (IFT) method to monodisperse systems. From this procedure the pair-distance distribution function p(r), the maximal diameter (Dmax), the radius of gyration (Rg), and the intensity on q = 0 (I(0)) were determined. I(0) was used to calculate the protein molecular weight in solution (55). The Rg of the scattering molecule was determined also by linear fitting of the Guinier region (56) to the following intensity function: ln(I(q)) = ln(I(0) − /3)q2. The calculation of the theoretical SAXS profile from known protein three-dimensional coordinates was carried out using CRYSOL (57). A first assessment of CALX-CBD12 flexibility was performed applying the EOM (20,36). RANCH (36) was used to build the initial CALX-CBD12 pool based on the crystal structure of the Ca2+-bound state (Protein Data Bank, PDB: 3RB5) (42) (Fig. 1) and the protein amino acid sequence. Residues D550–G555, in the linker between CALX-CBD1 and CALX-CBD2, were treated as flexible to generate a random pool of models, and the CBDs were treated as independent rigid bodies. A total of 10,000 models were built without symmetry or distance restraints. After the initial pool generation with RANCH, a total of 100 genetic algorithm (GA) cycles was performed with 1000 generations to select the optimal ensemble as implemented in GAJOE. As many as 50 ensembles were selected by GAJOE, each of them containing 20 or fewer models. Multiple copies of the same conformer in each ensemble were allowed during the optimization. The experimental SAXS data and analysis were deposited in the Small Angle Scattering Biological Data Bank under the accession codes SASBDB: SASDL26 (apo state) and SASBDB: SASDL36 (Ca2+-bound state).
MD simulation
MD trajectories of 3.4 μs were calculated using the AMBER18 software package (58) at the target temperature of 300.0 K using the Berendsen weak coupling thermostat with 5 ps coupling time constant, pressure of 1.0 bar maintained with the Monte Carlo barostat, and the AMBER ff14SB force field (59). Starting structures for the simulations were taken from the crystallographic coordinates of CALX-CBD12 in the Ca2+-bound state (PDB: 3RB5, chain A) (42). The four Ca2+ ions present at the CALX-CBD1 Ca2+-binding sites were removed to build the starting configuration for the apo state. The missing coordinates in the crystallographic models were set up using the LEaP program utility from the AMBER-PDB preparation topology functions (60,61). The two CALX-CBD12 initial structures (with and without the four Ca2+ ions) were solvated in an explicit TIP3P water model using 10 Å dimension water layer around the protein in an octahedral box. To neutralize the box, Na+ and Cl− ions were added as counterions; additional Na+ and Cl− ions were added to create a low salt concentration condition. First, energy minimization was performed during 2000 gradient steps, followed by heating up the system from 100 K up to 300 K within 50 ps in three stages with 25,000 steps of 2 fs time step. All the analyses of structural parameters were carried out using CPPTRAJ routines (62). The interdomain angle between CALX-CBD1 and CALX-CBD2 was measured considering three residues located near the centers of mass of CBD1 (A544), CBD12 (D551), and CBD2 (A688), and then the angle between the inter-residue vectors A554-D551 and D551-A688 was calculated for each MD trajectory snapshot. Structures were visualized with VMD (63).
RDC analysis
Experimental 1H-15N RDCs measured on 2H/15N-labeled CALX-CBD12 samples weakly aligned with Pf1 phages or mixtures of C12E5/n-hexanol (apo state) and with Pf1 phages or compressed polyacrylamide gels (Ca2+-bound state) were previously reported (51). Fittings of the individual MD snapshots to the experimental 1H-15N RDCs were carried out using singular value decomposition implemented in an in-house Python routine using NumPy (64,65). Briefly, each snapshot was aligned applying CPPTRAJ routine (62), with respect to a reference frame taken at 400 ns using all nonhydrogen atoms, after stabilization of the simulation. The Cartesian coordinates of the N-H bond vectors were extracted, normalized, and used to calculate matrix C:
(1) |
Order matrices, A, defined in an arbitrary Cartesian reference system were determined by solving a system of n linear equations (64,65):
(2) |
where n is the number of dipolar couplings, is the vector containing the experimental RDC values (DexpNH), and Aij are the five elements of the traceless order matrix, A. Once A was determined, RDC values were back calculated according to = CACT (64,65). The agreement between and was evaluated using an R-factor (66). The smaller the R-factor, the better the correlation between and , i.e., R = 0 means = . Alternatively, experimental 1H-15N RDCs were fitted to an ensemble of conformations as described by Niu and co-workers (67). In this analysis, population averaged N-H vector coordinates were calculated assuming m different conformations of CALX-CBD12 to build the new coordinate matrix C (Eq. 1) and an average alignment tensor for the ensemble. The six calculated terms , , , , , and now correspond to population averages defined as
(3) |
where wi is the population weight of model i with . The fractions of the different models, wi, were allowed to fluctuate to minimize the R-factor using a Levenberg-Marquart algorithm implemented using in-house Python scripts and the Scipy library (68).
Ensemble analysis combining MD with SAXS and NMR data
Two strategies were implemented to build a minimal structural ensemble of CALX-CBD12 using a Python routine (Fig. S1). In the first approach, snapshots in the production phase of the MD trajectory (from 0.4 μs up to 3.4 μs) were saved every 0.2 ns, resulting in a pool of 15,000 models, which is considered a reasonable sampling size (36,69) (step 1A; first approach) (Fig. S1). This pool was used as an external library to SAXS-EOM (step 2A; first approach), and the agreement of the EOM-selected representative conformers with experimental 1H-15N RDCs was evaluated by averaging the individual conformer coordinates (step 3A, first approach) (Fig. S1). In the second approach, a larger pool of 54,000 models was built saving MD snapshots from the production phase of the trajectory every 60 ps (step 1B; second approach) (Fig. S1). The consistency of each snapshot with experimental 1H-15N RDCs was first evaluated by singular value decomposition, generating an R-factor distribution of the MD trajectory (step 2B; second approach) (Fig. S1). All models with R-factors lower than given thresholds were selected to form external pools to SAXS-EOM (step 3B, second approach) (Fig. S1). A total of six R-factor limits were tested for the apo state (Pf1 data set): 0.54 (15,000 models), 0.53 (10,000 models), 0.52 (5000 models), 0.47 (1000 models), 0.43 (100 models), and 0.42 (50 models). Finally, the SAXS-EOM analysis (step 4B, second approach, Fig. S1) was carried out to select a minimal structural ensemble. CRYSOL (57) was used to compute the theoretical SAXS intensity of each individual model in the pool and GAJOE, using the GA algorithm, to select the optimal ensemble that best describes the SAXS data (36). Briefly, a total of 100 GA cycles were performed with 1000 generations to select the optimal ensemble, as implemented in GAJOE. As many as 50 ensembles were selected by GAJOE, each of them containing 20 or fewer models. Multiple copies of the same conformer in each ensemble were allowed during the ensemble optimization.
Results and discussion
CALX-CBD12 displays conformational heterogeneity in the absence of Ca2+
To investigate whether CALX-CBD12 SAXS scattering profiles are sensitive to the differential interdomain dynamics exhibited by the apo and the Ca2+-bound states, we recorded SAXS intensities of CALX-CBD12 samples prepared in the absence and in the presence of CaCl2 at a molar ratio of 1:6 (CBD12/CaCl2) (Fig. 2 A). The radii of gyration (Rg) obtained from IFT analysis or via Guinier plot are 26.1 ± 0.1 or 25.3 ± 0.2 Å and 28.3 ± 0.1 or 27.1 ± 0.3 Å for the apo and the Ca2+-bound states, respectively (Table 1). These results indicate the predominance of slightly fewer open structures in the apo state, whereas extended conformations are predominant in the Ca2+-bound state. Because a crystal structure of CALX-CBD12 in the Ca2+-bound state is available (PDB: 3RB5) (42) (Fig. 1), we compared the theoretical SAXS intensities with the experimental data using CRYSOL (57). The calculated SAXS intensities showed good agreement with the experimental data obtained in the presence of Ca2+ (χ2 = 2.3) (Fig. 2 D), whereas a poorer agreement was observed in its absence, particularly at low q-values indicating a more compact arrangement for the apo state and at the intermediate q range (χ2 = 10.2). The pair-distance distribution function p(r) observed for the apo state is characteristic of a slightly elongated molecule. In contrast, the p(r) obtained for the Ca2+-bound state shows an additional shoulder at ∼55 Å, which is characteristic of a two-domain protein with larger Dmax relative to the apo state (Fig. 2 B). Further analysis with the Kratky plot suggested that CALX-CBD12 behaves as a rigid body in the Ca2+-bound state, whereas the presence of a broad second shoulder at the q = 0.15–0.20 Å−1 region suggests that it behaves as two domains connected by a flexible linker in the absence of Ca2+ (Fig. 2 C) (70). To provide further evidence for flexibility in the apo state, we built CALX-CBD12 models displaying interdomain angles in the range of 63–145° and computed their SAXS intensities using CRYSOL and GNOM (Fig. S2). These calculations showed that the CALX-CBD12 pair distribution function, p(r), is highly affected by the interdomain angle. The p(r) functions of extended conformers (interdomain angles of 116–145°) display a smaller shoulder at ∼55 Å, which increases in intensity followed by a decrease in Dmax in the case of compact interdomain arrangements (Fig. S2). The p(r) distribution function exhibited by CALX-CBD12 in the apo state displays similar Dmax as the bound state (Table 1), whereas it loses the second maximal correlation at 55 Å, suggesting the presence of conformational heterogeneity (Fig. 2). Altogether, the data are consistent with the view that CALX-CBD12 is best described by an ensemble of conformations in the apo state as previously indicated by NMR (51), but it also suggests that compact interdomain arrangements are sampled more frequently than extended ones.
Table 1.
Sample | Apo | Ca2+ | 3RB5a |
---|---|---|---|
Guinier analysis | |||
Rg (Å) | 25.3 ± 0.2 | 27.1 ± 0.3 | 28 |
I(0) (cm−1) | 0.118 ± 5 × 10−3 | 0.114 ± 4 × 10−3 | 0.121462 |
p(r) analysis | |||
Rg (Å) | 26.1 ± 0.1 | 28.3 ± 0.1 | 28 |
Dmax (Å) | 87 | 93 | 93 |
I(0) (cm−1) | 0.1175 ± 4 × 10−3 | 0.1143 ± 3 × 10−3 | N/A |
MW (kDa) | 33 ± 3 | 33 ± 3 | 30 |
Analysis of the SAXS intensity profiles by rigid body modeling using the EOM approach (20) showed that two Rg distributions consisting of relatively compact models describe the apo state SAXS intensities, whereas a single narrower Rg distribution with the CBDs in extended arrangements was selected for the Ca2+-bound state (Fig. S3). However, the agreement between calculated and experimental SAXS intensities in the apo state remained poor (χ2 = 10.3), suggesting that the experimental data were not fully described by the EOM-selected ensemble (Fig. S3 B). This observation may be rationalized considering that the initial pool of CALX-CBD12 conformers was built using RANCH by randomization of the short interdomain linker, whereas CALX-CBD1 and CALX-CBD2 were treated as rigid bodies. Hence, this initial pool may not capture the complexity of the conformational energy landscape of the two-domain construct. To start from a more realistic library of conformers, we carried out atomistic MD simulations of CALX-CBD12 in the apo and in the Ca2+-bound states and used the MD ensemble as input for EOM.
Sampling CALX-CBD12 energy landscape by atomistic MD simulations
Long 3.4 μs MD trajectories of CALX-CBD12 in the Ca2+-bound and in the apo states were calculated in explicit water starting from the CALX-CBD12 crystal structure as described in the Materials and methods. Analysis of the backbone root mean-square deviation (RMSD) relative to the first frame indicated that the simulations took ∼0.4 μs to equilibrate (Fig. 3, A and B). The Ca2+-bound form deviated from the starting x-ray structure and was confined into a major conformational state characterized by RMSD of ∼5 Å and interdomain angle = 81 ± 7° during most of the trajectory, experiencing frequent jumps to other conformations with RMSDs in the range of 3.5–7 Å (Figs. 3, A and B and S4). In contrast, the apo form visited three major conformational states characterized by averaged RMSDs of ∼3.5, 5.0, and 8.5 Å from the initial structure and mean interdomain angles = 116 ± 8, 87 ± 6, and 81 ± 5° (Figs. 3, A and B and S4).
Comparison of the backbone root mean-square fluctuation (RMSF) along the two trajectories indicated that Ca2+ binding to CALX-CBD1 restricted the backbone motions in the Ca2+-binding region corresponding to the AB, CD, and EF loops of CBD1 and the linker between the two domains (Fig. 3, C and D). Particularly, larger fluctuations were observed at the interstrand loops located opposite to the interdomain interface, AB, CD, and EF in CBD2 and BC, DE, and FG in CBD1. The only exception is the FG loop in CBD2, which displayed large fluctuations despite being located near the interdomain interface. Fluctuations of the EF and FG loops in CBD2 were not affected by occupation of the Ca2+-binding sites in CBD1 (Fig. 3, C and D). However, the CBD2 CD loop, despite of its location opposite to the Ca2+-binding sites, displayed increased flexibility in the Ca2+-bound state than in the apo state. We attempted to investigate whether this observation could be in agreement with previously published NMR data for CALX-CBD12 (51); however, NMR signals in this region were broad, and relaxation rates could not be measured in the Ca2+-bound state. These observations are consistent with the presence of dynamics at the microsecond-to-millisecond timescale in the CBD2 CD loop. Indeed, this loop does not show electron density in the crystallographic model of CALX-CBD12 (Fig. 3 D).
Residues E455 and D552, which coordinate Ca2+ at sites Ca1, Ca2, and Ca3 in CBD1 (42), remained ∼4.3 Å apart during most of the trajectory in the Ca2+-bound state (Fig. 4). Furthermore, the orientations of D552 and D550 side chains, which coordinate Ca2+ at sites Ca1 and Ca2, also remained stable (Fig. S5). These observations indicate that the Ca2+ coordination geometry was preserved during the simulation. In contrast, in the apo state simulation, E455 and D552 became further apart and sampled two major conformational states, indicating that the Ca2+-coordination geometry was disrupted and suggesting greater flexibility at the Ca2+-binding sites (Figs. 4 and S5). This conclusion is consistent with the observation that R584 was more mobile in the apo state than in the Ca2+-bound state simulation (Fig. S6). R584 could play a pivotal role in the stabilization of the extended interdomain arrangement seen in the x-ray structure, as it participates in a hydrogen-bonding network with N615 in CBD2 and D552 that coordinates Ca2+ at sites Ca1 and Ca2 in CBD1 (Fig. 4). Furthermore, the formation of a salt bridge between R584 and D517 in the CBD1 EF loop and a hydrogen bond between D521 in the CBD1 EF loop and R673 in α-helix H2 (Figs. 4 and S6) may help stabilize the extended arrangement between CALX-CBD1 and CALX-CBD2. These two interactions break in more compact interdomain arrangements, in which E521 makes a new hydrogen bond with R681 in the C-terminal end of α-helix H2 (Figs. 4 and S6). Indeed, these analyses are consistent with the observation that CALX-CBD12 deviated from the initial structure and sampled compact interdomain arrangements during most of the trajectory in the Ca2+-bound state (Fig. 3). The crystallographic structure of CALX-CBD12 shows hydrophobic contacts (<4 Å) between F519 in the EF loop with L677 and S678 in α-helix H2 (Fig. 4) (42). When these contacts were examined during the course of the MD simulations, it was found that the distance between F519 and S678 displayed broader bimodal distribution in the apo state relative to the Ca2+-bound state (Fig. 4), consistent with the fact that CALX-CBD12 more frequently sampled a wider range of different interdomain arrangements in the former state.
Using the MD trajectory as initial pool for SAXS-EOM analysis
We investigated whether refinement of an MD-derived pool against SAXS data could improve the agreement between experimental and calculated SAXS intensities. To this end, an external pool corresponding to 15,000 MD snapshots was selected and refined against SAXS data using EOM. The refined ensemble (MD-EOM) yielded a better agreement with experimental SAXS profiles than the rigid body ensemble (RANCH-EOM) as indicated by the χ2, which decreased from 6.6 to 5.2 and from 10.3 to 1.2 for the Ca2+-bound and the apo states, respectively (Figs. S3 and S7). The apo state MD-EOM ensemble consisted of four Rg distributions, with representative models displaying Rg-values of 23.7, 24.9, 24.9, and 26.8 Å and interdomain angles of 75, 88, 89, and 127°. With the exception of the latter, all of them are more compact than the x-ray structure that displays Rg = 28 Å and interdomain angle of 131° (Table S1). It is, therefore, clear that a trend toward compact conformations is observed for the apo state (Fig. S7). In the case of the Ca2+-bound state, two narrower Rg and Dmax distributions were obtained (Fig. S7). The two representative conformers displayed similar interdomain arrangements as indicated by interdomain angles of 126 and 129°, consistent with a relatively rigid two-domain protein in the extended conformation (Fig. S7). It is noteworthy that the apo state SAXS-refined ensemble contained a population of the wide-open conformation characterized by Rg ∼26.7 Å, which is, however, enriched in the Ca2+-bound ensemble. Comparison of the initial MD pool with the MD-EOM ensemble shows that CALX-CBD12 became trapped in a local minimum far from the native structure along most of the MD trajectory in the Ca2+-bound state (Fig. S7).
RDCs measured on protein samples weakly aligned in the magnetic field report on the average orientation of the internuclear bond vector with respect to an alignment tensor. Hence, RDCs are rich sources of information on structure and dynamics of proteins in solution. It was previously reported that 1H-15N RDCs obtained for CALX-CBD12 in the apo state did not agree with the crystallographic structure as indicated by R = 0.57 obtained for RDCs measured in the presence of Pf1 (42,51). However, better agreement was obtained with the coordinates of the separate domains as indicated by R = 0.26 and 0.34 for CALX-CBD1 and CALX-CBD2, respectively (Fig. S8). Fitting the experimental RDCs to the population averaged coordinates of the four apo state MD-EOM representative CALX-CBD12 conformers (Eqs. 2 and 3) improved the R-factor only slightly from 0.57 to 0.44 (Figs. S8 and S9), suggesting that MD sampling of CALX-CBD12 conformational space was not sufficient or that the SAXS data did not contain information on the local geometry of the HN bond vectors. This analysis yielded conformer populations that were highly correlated with each other, indicating that the correct fraction of each model was unreliable. In contrast, RDCs measured for the Ca2+-bound state were in reasonable agreement with the crystal structure (R = 0.35) (51), which is consistent with the rigidity of the Ca2+-bound state (Fig. S8).
Using RDCs to select native-like conformations sampled during MD
During the MD simulations, CALX-CBD12 eventually visited conformations that are consistent with the 1H-15N RDCs. R-factor distributions of all MD conformers were calculated as described in the Materials and methods (Fig. 5). The apo state R-factor distribution was centered at approximately R = 0.6 (Fig. 5, red), indicating that conformational sampling was insufficient to improve the agreement with the experimental RDC data set. However, a few snapshots exhibited R-factors in the range of 0.3–0.4, indicating consistency with the experimental data. The lowest R-factor MD snapshot corresponded to a compact interdomain arrangement (Fig. 5, red), and reproduced well the experimental SAXS curve with χ2 = 1.6 using CRYSOL (Fig. S10). However, the calculated pair distribution function p(r) indicated a more compact conformation than the experimental curve (Fig. S10). Furthermore, the CorMap analysis indicated that the calculated and experimental SAXS intensities differ mainly at the low q region (Fig. S12). Therefore, these snapshots possibly represent compact conformers that are substantially populated in the native apo state ensemble, but they do not fully describe the experimental SAXS data. Similarly, the R-factor distribution of the Ca2+-bound MD ensemble was centered at R = 0.4 (Fig. 5, blue), larger than that obtained with the crystallographic structure (R = 0.35) and consistent with the fact that CALX-CBD12 deviated from the Ca2+-bound native conformation during most of the MD trajectory, visiting more compact interdomain arrangements (Fig. S7). As observed for the apo state, a few MD snapshots displayed improved agreement with the RDCs than the crystal structure, as indicated by R-factors as low as 0.28. The snapshot displaying the lowest R-factor showed an overall agreement with the interdomain arrangement observed in the crystal structure (Fig. 5 B, blue), suggesting that the improved agreement with the experimental RDCs could result from local reorientations of the HN bond vectors rather than changes in the interdomain orientation. Indeed, the local geometry overall quality improved during the simulation as indicated by an increase from 78 to 88% of residues in the most favored regions of the Ramachandran plot in the MD snapshot relative to the x-ray structure. Similar observations were made for RDC data sets measured using either C12E5/n-hexanol mixtures (apo state) or compressed polyacrylamide gels (Ca2+-bound state) as the alignment medium and are shown in Fig. S10.
Combining SAXS and RDCs to refine the apo state CALX-CBD12 MD ensemble
To select a subensemble that agrees with both low-resolution and high-resolution structural information, we used RDC and SAXS experimental data to refine the MD ensemble of CALX-CBD12. We first filtered the external MD pool based on the R-factor limit of 0.54, corresponding to 15,000 snapshots (Figs. 5 and S1). This RDC-filtered pool was then refined with SAXS data using the EOM approach. The refined MD-RDC-EOM subensemble displayed trimodal Rg and Dmax distributions (Figs. 6 and S11). It contains a larger number of compact conformers, whereas those ones characterized by wider interdomain arrangements are less populated (Fig. 6). The selected representative subensemble reproduced the experimental SAXS intensities with χ2 = 1.2 and yielded a better agreement with the experimental data at low and midrange q-values (0.1 < q < 0.2 Å−1) according to two-dimensional analysis using CorMap (Fig. S12), representing an improvement with respect to the fittings obtained using single models. The ensemble description for CALX-CBD12 adopted here is further justified by the observation of interdomain dynamics from the analysis of previously published NMR data (51). The three representative conformers display compact (population p1 = 59%, θ = 79°), semiopen (p2 = 33%, θ = 85°), and wide-open (p3 = 8%, θ = 111°) interdomain arrangements, the latter being similar to the Ca2+-bound state (Figs. 6 and S11). Additional SAXS-EOM runs carried out starting from smaller apo state external pools by assuming lower R-factor thresholds converged to similar subensembles, predominantly composed of semiopen (Rg = 25.5 Å and θ = 92°) and compact (Rg = 23.4 Å and θ = 73°) models with the advantage of consuming less computer time (Table S1). When the flexibility of CALX-CBD12 was quantified based on SAXS data, values of Rflex = 69% and Rσ = 1.1 were obtained for the apo state (Table S1). Despite the smallest subensembles having a reduced Rflex (Rflex decreases from 78% for the largest pool to 56% for the smallest pool), the ratios of the standard deviations of the selected distributions with respect to the pool, Rσ, remained greater than 0.8, near the maximal expected average of 1.0 for a fully flexible system (36) (Table S1).
As a control, the same MD-RDC-EOM analysis was carried out for the Ca2+-bound state using an R-factor threshold of 0.40 corresponding to a MD ensemble of 15,000 snapshots. When the external ensemble was refined using SAXS data, a single and narrow Rg distribution centered around wide-open conformers, which included the crystalographic structure of CALX-CBD12 in the Ca2+-bound state (PDB: 3RB5), was obtained. The representative structure displayed interdomain angle θ = 121°, backbone RMSD of 2.3 Å with respect to 3RB5, and R-factor of 0.37, nearly the same as that displayed by the crystallographic structure (R = 0.35) (Fig. 6). Flexibility analysis based on SAXS data yielded Rflex = 52% and Rσ = 0.12 for the Ca2+-bound state, consistent with a rigid system (Table S1). Therefore, the Ca2+-bound state is best represented by the 3RB5 structure. Similar conclusions were obtained when other 1H-15N RDC data sets, measured either in PEG (apo state) or in compressed polyacrylamide gels (Ca2+-bound state), were used for filtering (Fig. S13).
Conclusion
Understanding the CALX Ca2+ regulation mechanism in atomic detail requires description of the conformational energy landscape of the two-domain Ca2+ sensor, CBD12, in the apo state. Here, this question was addressed using molecular dynamics simulations, SAXS, and RDC experimental data in combination with the EOM approach for the analysis of flexible multidomain proteins. It was found that CALX-CBD12 interconverts between open (130 > θ > 110°) and closed (99 > θ > 70°) interdomain arrangements in the apo state; however, the latter is more frequently sampled. In contrast, a single structure corresponding to the open interdomain arrangement (θ ∼131°) best represents the Ca2+-bound state. Although the MD simulations sampled a range of open and closed conformations both in the Ca2+-bound and the apo states, a preference for more closed conformations was observed. Especially for the Ca2+-bound case, this is not consistent with the experimental result of a dominating extended interdomain orientation. One needs to consider that small inaccuracies of the force field representation by 1–2 kcal/mol are equivalent to relative population changes (e.g., between closed versus open states) of a factor of 10. In summary, these findings highlight the interdomain motions experienced by the apo state. Binding of Ca2+ shifts this equilibrium toward the wide-open state (Fig. 7). This observation is in agreement with the previous proposal that this population shift event will trigger a tension on the linker domains that eventually could be relayed to the transmembrane helices, causing CALX inactivation (51). The overall strategy of combining optimized conformational sampling from MD simulations, high-resolution information from NMR-RDC data, and low-resolution information from SAXS data can be a good guide for the investigation of other types of multidomain proteins.
Author contributions
M.F.d.S.D. collected and analyzed SAXS data and analyzed RDC data. P.A.M.V. prepared protein samples required for SAXS data collection. L.A.A. provided RDC data. M.F.d.S.D. and M.Z. carried out MD simulations and analysis of MD trajectories. M.F.d.S.D., M.Z., M.S., R.K.S., and C.L.P.O. designed the research and wrote the manuscript.
Acknowledgments
This work was supported by grants from the São Paulo Research Foundation (FAPESP 2016/07490-1; 2019/19968-1; 2018/16092-5; 2016/24531-3) and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq 420490/2016-7). P.M.V. received FAPESP PhD fellowship (2017/05614-8). M.F.d.S.D. and L.A.A. received PhD fellowships from Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES #88882.160170/2014-01 and #33002010017P0). R.K.S. and C.L.P.O. receive CNPq research fellowships (#311807/2017-8, #303001/2019-4). This work was supported by the Instituto Nacional de Ciência e Tecnologia em Fluídos Complexos, INCT-GFCx.
Editor: Jill Trewhella.
Footnotes
Supporting material can be found online at https://doi.org/10.1016/j.bpj.2021.07.022.
Contributor Information
Cristiano L.P. Oliveira, Email: crislpo@if.usp.br.
Roberto K. Salinas, Email: rsalinas@usp.br.
Supporting material
References
- 1.Wuthrich K. The development of nuclear magnetic resonance spectroscopy as a technique for protein structure determination. Acc. Chem. Res. 1989;22:36–44. [Google Scholar]
- 2.Markwick P.R.L., Malliavin T., Nilges M. Structural biology by NMR: structure, dynamics, and interactions. PLoS Comput. Biol. 2008;4:e1000168. doi: 10.1371/journal.pcbi.1000168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Bertini I., Kursula P., Yuan J. Accurate solution structures of proteins from X-ray data and a minimal set of NMR data: calmodulin-peptide complexes as examples. J. Am. Chem. Soc. 2009;131:5134–5144. doi: 10.1021/ja8080764. [DOI] [PubMed] [Google Scholar]
- 4.Göbl C., Madl T., Sattler M. NMR approaches for structural analysis of multidomain proteins and complexes in solution. Prog. Nucl. Magn. Reson. Spectrosc. 2014;80:26–63. doi: 10.1016/j.pnmrs.2014.05.003. [DOI] [PubMed] [Google Scholar]
- 5.Simon B., Madl T., Sattler M. An efficient protocol for NMR-spectroscopy-based structure determination of protein complexes in solution. Angew. Chem. Int. Ed. Engl. 2010;49:1967–1970. doi: 10.1002/anie.200906147. [DOI] [PubMed] [Google Scholar]
- 6.Madl T., Güttler T., Sattler M. Structural analysis of large protein complexes using solvent paramagnetic relaxation enhancements. Angew. Chem. Int. Ed. Engl. 2011;50:3993–3997. doi: 10.1002/anie.201007168. [DOI] [PubMed] [Google Scholar]
- 7.Pickford A.R., Campbell I.D. NMR studies of modular protein structures and their interactions. Chem. Rev. 2004;104:3557–3566. doi: 10.1021/cr0304018. [DOI] [PubMed] [Google Scholar]
- 8.Su X.-C., Ozawa K., Otting G. NMR analysis of the dynamic exchange of the NS2B cofactor between open and closed conformations of the West Nile virus NS2B-NS3 protease. PLoS Negl. Trop. Dis. 2009;3:e561. doi: 10.1371/journal.pntd.0000561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Oliveira C. 2011. Current Trends in X-Ray Crystallography.https://www.intechopen.com/books/current-trends-in-x-ray-crystallography [Google Scholar]
- 10.Mertens H.D.T., Svergun D.I. Structural characterization of proteins and complexes using small-angle X-ray solution scattering. J. Struct. Biol. 2010;172:128–141. doi: 10.1016/j.jsb.2010.06.012. [DOI] [PubMed] [Google Scholar]
- 11.Brosey C.A., Tainer J.A. Evolving SAXS versatility: solution X-ray scattering for macromolecular architecture, functional landscapes, and integrative structural biology. Curr. Opin. Struct. Biol. 2019;58:197–213. doi: 10.1016/j.sbi.2019.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Grishaev A., Tugarinov V., Bax A. Refined solution structure of the 82-kDa enzyme malate synthase G from joint NMR and synchrotron SAXS restraints. J. Biomol. NMR. 2008;40:95–106. doi: 10.1007/s10858-007-9211-5. [DOI] [PubMed] [Google Scholar]
- 13.Schwieters C.D., Clore G.M. Using small angle solution scattering data in Xplor-NIH structure calculations. Prog. Nucl. Magn. Reson. Spectrosc. 2014;80:1–11. doi: 10.1016/j.pnmrs.2014.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Gabel F., Simon B., Sattler M. A structure refinement protocol combining NMR residual dipolar couplings and small angle scattering restraints. J. Biomol. NMR. 2008;41:199–208. doi: 10.1007/s10858-008-9258-y. [DOI] [PubMed] [Google Scholar]
- 15.Grishaev A., Wu J., Bax A. Refinement of multidomain protein structures by combination of solution small-angle X-ray scattering and NMR data. J. Am. Chem. Soc. 2005;127:16621–16628. doi: 10.1021/ja054342m. [DOI] [PubMed] [Google Scholar]
- 16.Madl T., Gabel F., Sattler M. NMR and small-angle scattering-based structural analysis of protein complexes in solution. J. Struct. Biol. 2011;173:472–482. doi: 10.1016/j.jsb.2010.11.004. [DOI] [PubMed] [Google Scholar]
- 17.Venditti V., Egner T.K., Clore G.M. Hybrid approaches to structural characterization of conformational ensembles of complex macromolecular systems combining NMR residual dipolar couplings and solution X-ray scattering. Chem. Rev. 2016;116:6305–6322. doi: 10.1021/acs.chemrev.5b00592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Henzler-Wildman K., Kern D. Dynamic personalities of proteins. Nature. 2007;450:964–972. doi: 10.1038/nature06522. [DOI] [PubMed] [Google Scholar]
- 19.Smock R.G., Gierasch L.M. Sending signals dynamically. Science. 2009;324:198–203. doi: 10.1126/science.1169377. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Bernadó P., Mylonas E., Svergun D.I. Structural characterization of flexible proteins using small-angle X-ray scattering. J. Am. Chem. Soc. 2007;129:5656–5664. doi: 10.1021/ja069124n. [DOI] [PubMed] [Google Scholar]
- 21.Blackledge M. Recent progress in the study of biomolecular structure and dynamics in solution from residual dipolar couplings. Prog. Nucl. Magn. Reson. Spectrosc. 2005;46:23–61. [Google Scholar]
- 22.Tolman J.R., Ruan K. NMR residual dipolar couplings as probes of biomolecular dynamics. Chem. Rev. 2006;106:1720–1736. doi: 10.1021/cr040429z. [DOI] [PubMed] [Google Scholar]
- 23.Ravera E., Sgheri L., Luchinat C. A critical assessment of methods to recover information from averaged data. Phys. Chem. Chem. Phys. 2016;18:5686–5701. doi: 10.1039/c5cp04077a. [DOI] [PubMed] [Google Scholar]
- 24.Allison J.R. Using simulation to interpret experimental data in terms of protein conformational ensembles. Curr. Opin. Struct. Biol. 2017;43:79–87. doi: 10.1016/j.sbi.2016.11.018. [DOI] [PubMed] [Google Scholar]
- 25.Schneidman-Duhovny D., Hammel M., Sali A. FoXS, FoXSDock and MultiFoXS: single-state and multi-state structural modeling of proteins and their complexes based on SAXS profiles. Nucleic Acids Res. 2016;44:W424–W429. doi: 10.1093/nar/gkw389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Bonomi M., Heller G.T., Vendruscolo M. Principles of protein structural ensemble determination. Curr. Opin. Struct. Biol. 2017;42:106–116. doi: 10.1016/j.sbi.2016.12.004. [DOI] [PubMed] [Google Scholar]
- 27.Bernadó P., Modig K., Akke M. Structure and dynamics of ribosomal protein L12: an ensemble model based on SAXS and NMR relaxation. Biophys. J. 2010;98:2374–2382. doi: 10.1016/j.bpj.2010.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Sonntag M., Jagtap P.K.A., Sattler M. Segmental, domain-selective perdeuteration and small-angle neutron scattering for structural analysis of multi-domain proteins. Angew. Chem. Int.Engl. 2017;56:9322–9325. doi: 10.1002/anie.201702904. [DOI] [PubMed] [Google Scholar]
- 29.Wang J., Zuo X., Wang Y.X. Determination of multicomponent protein structures in solution using global orientation and shape restraints. J. Am. Chem. Soc. 2009;131:10507–10515. doi: 10.1021/ja902528f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Debiec K.T., Whitley M.J., Gronenborn A.M. Integrating NMR, SAXS, and atomistic simulations: structure and dynamics of a two-domain protein. Biophys. J. 2018;114:839–855. doi: 10.1016/j.bpj.2018.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Deshmukh L., Schwieters C.D., Clore G.M. Structure and dynamics of full-length HIV-1 capsid protein in solution. J. Am. Chem. Soc. 2013;135:16133–16147. doi: 10.1021/ja406246z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Bertini I., Giachetti A., Svergun D.I. Conformational space of flexible biological macromolecules from average data. J. Am. Chem. Soc. 2010;132:13553–13558. doi: 10.1021/ja1063923. [DOI] [PubMed] [Google Scholar]
- 33.Potrzebowski W., Trewhella J., Andre I. Bayesian inference of protein conformational ensembles from limited structural data. PLoS Comput. Biol. 2018;14:e1006641. doi: 10.1371/journal.pcbi.1006641. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Antonov L.D., Olsson S., Hamelryck T. Bayesian inference of protein ensembles from SAXS data. Phys. Chem. Chem. Phys. 2016;18:5832–5838. doi: 10.1039/c5cp04886a. [DOI] [PubMed] [Google Scholar]
- 35.Pelikan M., Hura G.L., Hammel M. Structure and flexibility within proteins as identified through small angle X-ray scattering. Gen. Physiol. Biophys. 2009;28:174–189. doi: 10.4149/gpb_2009_02_174. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Tria G., Mertens H.D.T., Svergun D.I. Advanced ensemble modelling of flexible macromolecules using X-ray solution scattering. IUCrJ. 2015;2:207–217. doi: 10.1107/S205225251500202X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Różycki B., Kim Y.C., Hummer G. SAXS ensemble refinement of ESCRT-III CHMP3 conformational transitions. Structure. 2011;19:109–116. doi: 10.1016/j.str.2010.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Khananshvili D. Sodium-calcium exchangers (NCX): molecular hallmarks underlying the tissue-specific and systemic functions. Pflugers Arch. 2014;466:43–60. doi: 10.1007/s00424-013-1405-y. [DOI] [PubMed] [Google Scholar]
- 39.Reeves J.P., Condrescu M. Ionic regulation of the cardiac sodium-calcium exchanger. Channels (Austin) 2008;2:322–328. doi: 10.4161/chan.2.5.6897. [DOI] [PubMed] [Google Scholar]
- 40.Nicoll D.A., Ottolia M., Philipson K.D. 20 years from NCX purification and cloning: milestones. Adv. Exp. Med. Biol. 2013;961:17–23. doi: 10.1007/978-1-4614-4756-6_2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Hilgemann D.W. Regulation and deregulation of cardiac Na(+)-Ca2+ exchange in giant excised sarcolemmal membrane patches. Nature. 1990;344:242–245. doi: 10.1038/344242a0. [DOI] [PubMed] [Google Scholar]
- 42.Wu M., Tong S., Zheng L. Structural basis of the Ca2+ inhibitory mechanism of Drosophila Na+/Ca2+ exchanger CALX and its modification by alternative splicing. Structure. 2011;19:1509–1517. doi: 10.1016/j.str.2011.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Giladi M., Bohbot H., Khananshvili D. Dynamic features of allosteric Ca2+ sensor in tissue-specific NCX variants. Cell Calcium. 2012;51:478–485. doi: 10.1016/j.ceca.2012.04.007. [DOI] [PubMed] [Google Scholar]
- 44.Hilge M., Aelen J., Vuister G.W. Ca2+ regulation in the Na+/Ca2+ exchanger involves two markedly different Ca2+ sensors. Mol. Cell. 2006;22:15–25. doi: 10.1016/j.molcel.2006.03.008. [DOI] [PubMed] [Google Scholar]
- 45.Hilge M., Aelen J., Vuister G.W. Ca2+ regulation in the Na+/Ca2+ exchanger features a dual electrostatic switch mechanism. Proc. Natl. Acad. Sci. USA. 2009;106:14333–14338. doi: 10.1073/pnas.0902171106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Besserer G.M., Ottolia M., Abramson J. The second Ca2+-binding domain of the Na+ Ca2+ exchanger is essential for regulation: crystal structures and mutational analysis. Proc. Natl. Acad. Sci. USA. 2007;104:18467–18472. doi: 10.1073/pnas.0707417104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Nicoll D.A., Sawaya M.R., Abramson J. The crystal structure of the primary Ca2+ sensor of the Na+/Ca2+ exchanger reveals a novel Ca2+ binding motif. J. Biol. Chem. 2006;281:21577–21581. doi: 10.1074/jbc.C600117200. [DOI] [PubMed] [Google Scholar]
- 48.Zheng L., Wu M., Tong S. Structural studies of the Ca(2+) regulatory domain of Drosophila Na(+)/Ca (2+) exchanger CALX. Adv. Exp. Med. Biol. 2013;961:55–63. doi: 10.1007/978-1-4614-4756-6_6. [DOI] [PubMed] [Google Scholar]
- 49.Hryshko L.V., Matsuoka S., Philipson K.D. Anomalous regulation of the Drosophila Na(+)-Ca2+ exchanger by Ca2+ J. Gen. Physiol. 1996;108:67–74. doi: 10.1085/jgp.108.1.67. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Salinas R.K., Bruschweiler-Li L., Brüschweiler R. Ca2+ binding alters the interdomain flexibility between the two cytoplasmic calcium-binding domains in the Na+/Ca2+ exchanger. J. Biol. Chem. 2011;286:32123–32131. doi: 10.1074/jbc.M111.249268. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Abiko L.A., Vitale P.M., Brüschweiler R. Model for the allosteric regulation of the Na+/Ca2+ exchanger NCX. Proteins. 2016;84:580–590. doi: 10.1002/prot.25003. [DOI] [PubMed] [Google Scholar]
- 52.Giladi M., Lee S.Y., Khananshvili D. Structure-based dynamic arrays in regulatory domains of sodium-calcium exchanger (NCX) isoforms. Sci. Rep. 2017;7:993. doi: 10.1038/s41598-017-01102-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Ramos C.R.R., Abreu P.A.E., Ho P.L. A high-copy T7 Escherichia coli expression vector for the production of recombinant proteins with a minimal N-terminal His-tagged fusion peptide. Braz. J. Med. Biol. Res. 2004;37:1103–1109. doi: 10.1590/s0100-879x2004000800001. [DOI] [PubMed] [Google Scholar]
- 54.Semenyuk A.V., Svergun D.I. GNOM - a program package for small-angle scattering data-processing. J. Appl. Cryst. 1991;24:537–540. [Google Scholar]
- 55.Oliveira C.L.P., Behrens M.A., Pedersen J.S. A SAXS study of glucagon fibrillation. J. Mol. Biol. 2009;387:147–161. doi: 10.1016/j.jmb.2009.01.020. [DOI] [PubMed] [Google Scholar]
- 56.Guinier A., Fournet G. Wiley; New York: 1955. Small-Angle Scattering of X-Rays. [Google Scholar]
- 57.Svergun D., Barberato C., Koch M.H.J. CRYSOL - a program to evaluate x-ray solution scattering of biological macromolecules from atomic coordinates. J. Appl. Cryst. 1995;28:768–773. [Google Scholar]
- 58.Case D.A., Ben-Shalom Y., Kollman P.A. University of California; San Francisco, CA: 2018. AMBER 2018. [Google Scholar]
- 59.Maier J.A., Martinez C., Simmerling C. ff14SB: improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theory Comput. 2015;11:3696–3713. doi: 10.1021/acs.jctc.5b00255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Case D.A., Cheatham T.E., III, Woods R.J. The Amber biomolecular simulation programs. J. Comput. Chem. 2005;26:1668–1688. doi: 10.1002/jcc.20290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Pearlman D.A., Case D.A., Kollman P. AMBER, a package of computer programs for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to simulate the structural and energetic properties of molecules. Comput. Phys. Commun. 1995;91:1–41. [Google Scholar]
- 62.Roe D.R., Cheatham T.E., III PTRAJ and CPPTRAJ: software for processing and analysis of molecular dynamics trajectory data. J. Chem. Theory Comput. 2013;9:3084–3095. doi: 10.1021/ct400341p. [DOI] [PubMed] [Google Scholar]
- 63.Humphrey W., Dalke A., Schulten K. VMD: visual molecular dynamics. J. Mol. Graph. 1996;14:33–38, 27–28.. doi: 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]
- 64.Losonczi J.A., Andrec M., Prestegard J.H. Order matrix analysis of residual dipolar couplings using singular value decomposition. J. Magn. Reson. 1999;138:334–342. doi: 10.1006/jmre.1999.1754. [DOI] [PubMed] [Google Scholar]
- 65.Showalter S.A., Brüschweiler R. Quantitative molecular ensemble interpretation of NMR dipolar couplings without restraints. J. Am. Chem. Soc. 2007;129:4158–4159. doi: 10.1021/ja070658d. [DOI] [PubMed] [Google Scholar]
- 66.Clore G.M., Garrett D.S. R-factor, free R, and complete cross-validation for dipolar coupling refinement of NMR structures. J. Am. Chem. Soc. 1999;121:9008–9012. [Google Scholar]
- 67.Niu X., Bruschweiler-Li L., Chapman M.S. Arginine kinase: joint crystallographic and NMR RDC analyses link substrate-associated motions to intrinsic flexibility. J. Mol. Biol. 2011;405:479–496. doi: 10.1016/j.jmb.2010.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Virtanen P., Gommers R., van Mulbregt P., SciPy 1.0 Contributors SciPy 1.0: fundamental algorithms for scientific computing in Python. Nat. Methods. 2020;17:261–272. doi: 10.1038/s41592-019-0686-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Bernadó P., Svergun D.I. Structural analysis of intrinsically disordered proteins by small-angle X-ray scattering. Mol. Biosyst. 2012;8:151–167. doi: 10.1039/c1mb05275f. [DOI] [PubMed] [Google Scholar]
- 70.Bernadó P. Effect of interdomain dynamics on the structure determination of modular proteins by small-angle scattering. Eur. Biophys. J. 2010;39:769–780. doi: 10.1007/s00249-009-0549-3. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.