Skip to main content
Protein Science : A Publication of the Protein Society logoLink to Protein Science : A Publication of the Protein Society
. 2014 Jun 20;23(9):1247–1261. doi: 10.1002/pro.2506

Structural and thermodynamic characterization of the recognition of the S100-binding peptides TRTK12 and p53 by calmodulin

Lucas N Wafer 1,2, Franco O Tzul 1,2, Pranav P Pandharipande 2,3, Scott A McCallum 2, George I Makhatadze 1,2,*
PMCID: PMC4243996  PMID: 24947426

Abstract

Calmodulin (CaM) is a multifunctional messenger protein that activates a wide variety of signaling pathways in eukaryotic cells in a calcium-dependent manner. CaM has been proposed to be functionally distinct from the S100 proteins, a related family of eukaryotic calcium-binding proteins. Previously, it was demonstrated that peptides derived from the actin-capping protein, TRTK12, and the tumor-suppressor protein, p53, interact with multiple members of the S100 proteins. To test the specificity of these peptides, they were screened using isothermal titration calorimetry against 16 members of the human S100 protein family, as well as CaM, which served as a negative control. Interestingly, both the TRTK12 and p53 peptides were found to interact with CaM. These interactions were further confirmed by both fluorescence and nuclear magnetic resonance spectroscopies. These peptides have distinct sequences from the known CaM target sequences. The TRTK12 peptide was found to independently interact with both CaM domains and bind with a stoichiometry of 2:1 and dissociations constants Kd,C-term = 2 ± 1 µM and Kd,N-term = 14 ± 1 µM. In contrast, the p53 peptide was found to interact only with the C-terminal domain of CaM, Kd,C-term =2 ± 1 µM, 25°C. Using NMR spectroscopy, the locations of the peptide binding sites were mapped onto the structure of CaM. The binding sites for both peptides were found to overlap with the binding interface for previously identified targets on both domains of CaM. This study demonstrates the plasticity of CaM in target binding and may suggest a possible overlap in target specificity between CaM and the S100 proteins.

Keywords: S100 proteins, calmodulin, isothermal titration calorimetry, nuclear magnetic resonance, line-shape analysis, p53, TRTK12, CapZ

Introduction

Calmodulin (CaM) is a small (17 kDa), acidic, calcium-binding protein that is ubiquitous in eukaryotes.1 It has been shown to bind more than 300 targets in response to transient fluctuations of intracellular calcium concentrations and plays critical roles in a number of cellular processes, including metabolism, the immune response, muscle contraction, and apoptosis.1 CaM is also thought to play a role in several neurodegenerative disorders, such as Parkinson's2 and Alzheimer's disease.3 This small protein consists of two distinct N- and C-terminal domains each containing two EF-hand calcium-binding motifs. Calcium binding induces large conformational changes within each globular domain which expose hydrophobic patches that are generally required for protein binding and various CaM functions.4 In the majority of available structures, the alpha helical linker between the N- and C-terminal domains unwinds to allow both domains to wrap around a single, usually α-helical, target peptide sequence.1,5 However, an increasing number of novel structures have been reported where CaM adopts a more open, flexible conformation and interacts with its targets with differing stoichiometries.4,68

Interestingly, CaM is proposed to be functionally distinct from the related, eukaryotic calcium-binding S100 proteins, due to differing subcellular localization, expression, and target recognition.911 In particular, the TRTKIDWNKILS peptide (TRTK12), derived from the C-terminal region (residues 265–276) of the actin-capping protein (CapZ) has been proposed to be a specific binding motif for the S100 proteins and structurally distinct from CaM-binding peptides.1214 TRTK12 was previously identified as a consensus binding sequence for S100B12,13 and subsequently shown to effectively act as a reversible inhibitor for interactions of glial fibrillary acidic protein or the full-length CapZ protein with S100B.12 In addition, the TRTK12 peptide has also been shown to bind S100A1, S100A2, and S100P in a calcium-dependent manner.15 Similarly, the p53 C-terminal peptide, derived from the negative regulatory domain (residues 367–388) of the p53 tumor-suppressor protein, has previously been shown to bind to some members of the S100 protein family, including S100B and S100A1, in a calcium-dependent manner.1618 The p53 protein is a tetrameric transcription activator that regulates cell cycle arrest and apoptosis in response to cellular stress.19 S100 protein binding has been shown to prevent kinase C-dependent phosphorylation of Ser376 and Ser378, which are important for p53 activation and are proposed to stabilize the biologically active, tetrameric form of the protein.14,16,19

To probe the binding specificity of the TRTK12 and p53 peptides to the S100 protein family, the peptides were screened for binding using isothermal titration calorimetry (ITC) against 16 members of the human S100 protein family, as well as CaM. Considering previous reports of distinct functionality of the S100 proteins and CaM,911,20 the latter was intended to serve as a negative control. Unexpectedly, novel interactions were detected for both the TRTK12 and p53 peptides with human CaM. To gain a better understanding of these interactions, the binding of TRTK12 and p53 peptides were further characterized using thermodynamic (ITC), kinetic (NMR), and structural (NMR) methods.

Results and Discussion

Interactions of the p53 peptide with the S100 proteins

The C-terminal p53 peptide was screened for binding against 16 representative members of the human S100 protein family using ITC. Interactions were detected with only four proteins, S100B, S100A1, S100A2, and S100A4, consistent with previous findings by us and others.14,16,18 The thermodynamics of these interactions are summarized in Table I and in the Supporting Information Figure S1. For the S100B and S100A1 proteins, the titration data fit well to the simplest binding model with a 1:1 stoichiometry, where an individual p53 peptide binds each subunit of S100 dimer. In contrast, the p53 peptide appears to bind the S100A2 and S100A4 proteins with a stoichiometry of 1:2, where a single peptide binds the S100 dimer. Such variations in the binding stoichiometry for the S100 proteins and peptides derived from p53 have previously been observed.16,21 The dissociation constants obtained in ITC experiments are between 2 and 200 µM and are modulated by ionic strength (Table I), likely as a result of electrostatic screening of the large number of charged residues in the p53 peptide. Under physiological salt concentrations (120 mM NaCl) the measured binding affinities are consistent with previously reported dissociation constants measured by other methods.16

Table I.

Summary of the Dissociation Constants and the Changes in the Heat Capacity upon Binding of the p53 Peptide to CaM, S100B, S100A1, S100A2, and S100A4 at Different Ionic Strengths

Protein Inline graphicaM) Inline graphicb (kJ mol−1 K−1) Inline graphicc (kJ mol−1 K−1)
S100B 2 ± 1 −1.0 ± 0.1 −1.0 ± 0.1
S100B, 120 mM NaCl 20 ± 5 −1.2 ± 0.3 −1.0 ± 0.1
S100A1 20 ± 5 −1.1 ± 0.1 −1.1 ± 0.1
S100A1, 120 mM NaCl >200d n.d. −1.1 ± 0.1
S100A2, 120 mM NaCl 50 ± 3 −0.8 ± 0.1 n.d.
S100A4, 120 mM NaCl 50 ± 4d n.d. n.d.
CaM 2 ± 1 −0.9 ± 0.1 −1.3 ± 0.2
CaM, 120 mM NaCl 20 ± 1 −0.8 ± 0.1 −1.3 ± 0.2
a

Dissociation constants at 25°C obtained using ITC.

b

The experimentally determined changes in the heat capacity upon binding (Fig. 2).

c

Structure-based calculations [Eq. (2)] of homology models based on one peptide binding to each subunit of the dimer of human s100 proteins.

d

Dissociation constant determined at 35°C.

n.d. not determined.

In addition, Table I also summarizes the heat capacity changes, ΔCp, on p53 binding the S100 proteins, obtained from the temperature dependencies of the enthalpies of binding (Supporting Information Fig. S1). The values of ΔCp, appear to cluster around −1.0 kJ mol−1 K−1. Importantly, for the S100B-p53 interaction, the experimentally determined ΔCp value does not appear to vary with increasing ionic strength (−1.0 ± 0.1 kJ mol−1 K−1 vs. −1.2 ± 0.3 kJ mol−1 K−1). The magnitude of these ΔCp values, as well as their similarity for all S100-p53 interactions, suggest a significant amount of hydrophobic surface area is buried in all S100-p53 complexes.18,22

Thermodynamics of binding of the p53 peptide to CaM

Figure 1(A) shows representative ITC data in which the C-terminal p53 peptide was titrated into CaM in the presence of 5 mM calcium chloride. The titration data fit well to the simplest model, where the p53 peptide binds CaM with a 1:1 stoichiometry (Table I). The dissociation constant obtained in the absence of physiological ionic strength, Kd = 2 ± 1 µM, is an order of magnitude tighter than that obtained in its presence, Kd = 20 ± 2 µM, (Table I). Increasing ionic strength has previously been shown to reduce the affinity of several biologically-relevant targets for CaM23,24 and is also observed, for example, in S100B-p53 interactions (Table I).

Figure 1.

Figure 1

Representative ITC experiments showing the binding of p53 and TRTK12 to human CaM in the presence of calcium. Upper plots in each panel represent the raw heat effects as a function of time and lower plots show the cumulative heat effects (▪) as a function of the molar ratio of peptide to protein, and the fits to the experimental data for binding of p53 to CaM (A), the binding of TRTK12 to CaM at 25°C (B), and the binding of TRTK12 to the CaM N-terminal fragment CaM(1–80) at 25°C (C).

Figure 2(A) shows the temperature dependence of the enthalpy of binding, ΔHcal, of p53 to CaM obtained in the absence and presence of physiological ionic strength. Although there is a change in the dissociation constant on p53 binding to CaM, the enthalpy of binding is independent of ionic strength. The slope of the temperature dependence is linear and corresponds to the heat capacity change on binding, ΔCp. These values are summarized in Table II. The experimentally determined ΔCp value does not appear to depend on ionic strength [Fig. 2(A)], as the heat capacity change in the absence of NaCl (−0.9 ± 0.1 kJ mol−1 K−1) is in excellent agreement with that obtained in the presence of 120 mM NaCl (−0.8 ± 0.1 kJ mol−1 K−1). The similarity between the enthalpies of binding and changes in heat capacity on binding in the absence and presence of physiological ionic strength, as well as the modest change in the dissociation constant, suggest that the interactions are not driven by nonspecific electrostatic interactions. Furthermore, the magnitude of the negative ΔCp suggests the involvement of hydrophobic interactions in the formation of the p53 peptide and CaM complex.

Figure 2.

Figure 2

Temperature dependence of the enthalpies of binding, ΔHcal, for the p53 and TRTK12 peptides to CaM. Panel A: Binding of the C-terminal p53 peptide to CaM in a buffer containing 5 mM calcium chloride and either 0 mM NaCl (red inverted triangles) or 120 mM NaCl (black inverted triangle). Panel B: Binding of the wild-type TRTK12 peptide to human CaM and the N-terminal fragment CaM(1–80) in the presence of 5 mM calcium chloride. The individual enthalpies for peptide binding to the first (inverted black triangles) and second (black triangles) binding sites of full-length human CaM are shown, as well as the total enthalpy of binding (black circles). The enthalpies of TRTK12 binding to CaM(1–80) are overlaid (inverted orange triangles). Solid lines represent linear fits of the ΔHcal temperature dependencies for each peptide-CaM interaction. The slopes of these lines represent the changes in heat capacity, which are summarized in.

Table II.

Summary of the Dissociation Constants, the Changes in the Heat Capacity upon Binding, and the Kinetic Rate Constants of the p53 Peptide and the TRTK12 Wild-Type Peptide and Its Alanine Variants Binding to CaM at 25°C

Peptide Inline graphicaM) Inline graphicaM) Inline graphicbM) Inline graphicc (kJ mol−1 K−1) Inline graphicc (kJ mol−1 K−1) Inline graphicd (kJ mol−1 K−1)
p53 2 ± 1 n.d. −0.9 ± 0.1 −1.3 ± 0.2, n/a
p53, 120 mM NaCl 20 ± 1 n.d. −0.8 ± 0.1 −1.3 ± 0.2 n/a
TRTK12 3 ± 1 14 ± 1 [25 ± 4] 7 ± 1 −0.6 ± 0.1 −0.6 ± 0.1 [−0.7 ± 0.2], −0.9 ± 0.2, −0.6 ± 0.2
TRTKM1 (T1A) 4 ± 1 20 ± 1 [20 ± 1] 5 ± 1 −0.7 ± 0.1 −0.6 ± 0.1 [−0.8 ± 0.2] −0.9 ± 0.2, −0.7 ± 0.2
TRTKM5 (I5A) 10 ± 1 100 ± 3 40 ± 1 −2.0 ± 0.3 −0.7 ± 0.1 −1.1 ± 0.2, −0.7 ± 0.2
TRTKM6 (D6A) 0.3 ± 0.1 6 ± 1 2 ± 1 −1.1 ± 0.1 −0.3 ± 0.1 −1.3 ± 0.2, −0.8 ± 0.2
TRTKM10 (I10A) 20 ± 2 100 ± 5 60 ± 4 −0.7 ± 0.1 −0.6 ± 0.1 −1.0 ± 0.2, −0.8 ± 0.2
TRTKM11 (L11A)e 80 ± 1 n.d.b. 70 ± 10 −1.0 ± 0.1 n.d.b. −1.1 ± 0.2, n/a
a

Dissociation constants obtained using ITC using two independent site binding model. Values in square brackets were obtained for titrations of the peptide into the N-terminal domain fragment of calmodulin CaM(1–80) (see Materials and Methods).

b

Values were obtained using emission fluorescence assays and single site model and thus represent apparent binding constants.

c

The experimentally determined changes in the heat capacity upon binding for each site (Figure 2). Values in brackets were obtained for titrations of the peptide into the N-terminal domain fragment of calmodulin CaM(1–80) (see Materials and Methods).

d

Structure-based calculations (Equation (2)) of homology models based on one peptide (PDB: 1CFF) or two peptides (PDB: 2LLQ/2LLO) binding to human CaM. First value for C-terminal domain and second value for N-terminal domain.

e

ITC Titration data fit best to a single-site binding model for this peptide.

n.d.b., no detectable binding; n/a, not applicable; n.d., not determined.

Thermodynamics of binding of the TRTK12 peptides with CaM

Figure 1(B) shows representative ITC data in which the wild-type TRTK12 peptide was titrated into CaM in the presence of 5 mM calcium chloride. The ITC data fit well to the two independent binding sites model. The 2:1 stoichiometry was confirmed by performing reverse titrations (data not shown). The dissociation constants, Kd,1 and Kd,2, obtained from the fit of the wild-type peptide binding to CaM were determined to be 3 ± 1 µM and 14 ± 1 µM at 25°C (Table II). There is approximately a fivefold difference in affinity between the two binding sites, suggesting the peptide preferentially binds to one of the domains of CaM.

In addition to the wild-type TRTK12, the thermodynamics of binding were measured for five alanine variants of this peptide (see Materials and Methods for sequences and rationale). There is reasonable agreement between the dissociation constants obtained from ITC for the individual sites and the apparent dissociation constants obtained using fluorescence spectroscopy (Table II). All peptides, except TRTKM11 (L11A), were found to bind with a stoichiometry similar to that of wild type, that is, two peptides per CaM. The TRTKM11 (L11A) peptide appears to bind CaM with a stoichiometry of 1:1, suggesting that L11 is a critical residue for binding and that the L11A substitution abolishes binding to one of the two available sites. Interestingly, the same TRTKM11 peptide was found to bind several S100 proteins, including S100B and S100P, with an atypical stoichiometry of one peptide per S100 dimer.15 For the majority of alanine variants, the dissociation constants for each of the two binding sites were found to differ from the wild-type peptides to different extents. For example, the dissociation constants for the first (Kd1 = 10 ± 1 µM) and second (Kd2 = 100 ± 2 µM) binding sites for the TRTKM5 variant are approximately threefold and sevenfold weaker, respectively, than those of wild type. The binding affinities for individual variants suggest that there may be important interactions associated with the hydrophobic side chains of I5, I10, and L11 of the TRTK12 peptide. For example, the binding affinities for the first (Kd1 = 20 ± 2 µM) and second (Kd2 = 100 ± 5 µM) binding sites of the TRTKM10 variant are approximately sevenfold weaker than those of wild type. Furthermore, the L11A substitution in the TRTKM11 peptide abolishes binding for one of the sites and significantly weakens binding for the other (Kd1 = 100 ± 2 µM). Electrostatic interactions also appear to modulate binding as the TRTKM6 (D6A) peptide was found to have a significantly tighter affinity than the wild-type peptide for both binding sites (Table II). This is consistent with the large number of negative charges on the surface of CaM, which may act to electrostatically repel the wild-type peptide.

Structural aspects of TRTK12 peptide binding to CaM

Previously, we have established that the TRTK12 and p53 peptides bind to CaM with a 2:1 and 1:1 stoichiometry, respectively. Here, we use NMR spectroscopy and computational modeling to map the location of the peptide binding sites on CaM. The interactions between TRTK12 and CaM were examined by monitoring binding-induced changes in the chemical shifts of 15N/1H CaM backbone atoms between the peptide-free and peptide-bound states. The TRTK12 peptide appears to bind CaM with a stoichiometry of 2:1, and occupy distinct binding sites on both the N- and C-terminal domains [Fig. 3(B)]. The residues involved in binding the wild-type TRTK12 peptide in the N-terminal domain appear to include Glu14, Phe16, Phe19, Gly33, Val35, Met36, Glu47, Asp50, Ile52, Phe68, Thr70, Met71, Met72, Ala73, Arg74, and Asp80. Significant chemical-shift changes were also observed for residues in the first (Thr26, Ile27, and Thr29) and second (Ala57, Ile63, and Asp64) calcium-binding sites within the N-terminal domain. This is consistent with observations for other CaM-binding peptides, such as melittin,26 and suggests the structural linkage between calcium binding and peptide binding to the N-terminal domain of CaM. The chemical-shift perturbations observed for residues from the C-terminal domain were similar to those involved in binding p53 and include Phe89, Val91, Leu105, Arg106, Asn111, Ala128, Phe141, Val142, Gln143, Met144, Met145, Thr146, Ala147, and Lys148. Significant binding-induced chemical-shift changes were also observed for residues in the third (Asp93 and Ala103) and fourth (Ile130 and Asp133) calcium-binding sites within the C-terminal domain.

Figure 3.

Figure 3

Changes in backbone 1H and 15N chemical shifts for TRTK12 (A) and p53 (B) binding to Ca2+-bound CaM (PDB: 2L7L25). Residues are color coded according to the degree of combined backbone 1H and 15N chemical-shift perturbations (c.s.p.) on TRTK12 (C) and p53 (D) binding. Color assignment was done as follows: c.s.p. 0.3–0.4 (yellow), c.s.p. 0.4–0.5 (orange), and c.s.p. > 0.5 (red). Selected residues are labeled for orientation only.

The 2:1 binding stoichiometry of TRTK12 to CaM suggests that the binding mode is different from canonical one, in which both CaM domains to wrap around a single target peptide sequence. However, a similar 2:1 stoichiometry has been observed for the interactions of the NSCaTE and NtMKP1 peptides with CaM.27,28 A sequence analysis of TRTK12 shows little similarity to the canonical CaM-binding epitopes. The peptide lacks the glycine and glutamine residues that are common to the IQ motif, as well as the hydrophobic residue spacing common to the 1-10, 1-14, and 1-16 motifs. Instead, the TRTK12 peptide aligns best with noncanonical binding sequences (Fig. 5), such as CaD-A and MLC, which have been shown to preferentially interact with the C-terminal domain of CaM.29

Figure 5.

Figure 5

Alignment of the TRTK12 peptide with known CaM-binding peptides. The peptides are aligned according to the position of their charged residues (blue or red) and hydrophobic anchor (green), which is proposed to bind in a hydrophobic pocket of the C-terminal domain of CaM. The location of the N- (N) and C-termini (C) of each peptide is indicated.

To confirm the binding model of TRTK12 to CaM, the interactions of the wild-type peptide and its alanine variants with the isolated N-terminal domain fragment of CaM, CaM(1–80), were also studied. Titrations of the TRTK12 peptides and CaM(1–80) consistently resulted in stoichiometry of 1:1 [Fig. 1(C)] and Kd = 25 ± 4 µM, which compares well with the Kd,2=14 ± 1 µM for the second site of the full-length CaM (Table II). Furthermore, the enthalpies of binding for the TRTK12 peptides and CaM(1–80) are in excellent agreement with those obtained for one of the two binding sites on full-length CaM [Fig. 2(B)]. Likewise, the ΔCp value obtained for the binding of the TRTK12 peptide to CaM(1–80), −0.7 ± 0.1 kJ mol−1 K−1, is similar to that obtained for TRTK12 binding to one of the individual sites on full-length CaM, −0.6 ± 0.1 kJ mol−1 K−1 (Table II). Together, these results confirm that the binding sites are delocalized and there is one for each CaM domain.

While the orientation of the peptide within the N-terminal domain is unknown, it is likely that tryptophan 7 (W7) of TRTK12 is buried in a hydrophobic pocket. This is supported by tryptophan fluorescence experiments performed by titrating CaM(1–80) into a solution containing the wild-type TRTK12 peptide, where both an increase in intensity and a blue shift in λmax were observed (Supporting Information Figure S2). In addition, significant chemical-shift perturbations were observed for a large number of hydrophobic and charged residues in the N-terminal domain that make up the canonical binding pocket of CaM. These residues include Glu14, Phe19, Val35, Met36, Glu47, Ile63, Phe68, Met71, Met72, and Asp80 [Fig. 3(B)]. Homology models based on the CaM-NSCaTE (PDB: 2LQC27) and CaM-ER-α (PDB: 2LLO6) complexes are consistent with the burial of W7. For illustrative purposes, Figure 4 also shows a high-scoring complex of CaM-TRTK12 generated using HADDOCK (see Materials and Methods for details). The generated structures were furthered evaluated and scored based on their ability to recapitulate the experimental data (see Materials and Methods for selection criteria). As with the homology models, the HADDOCK-generated structures bury W7 in the canonical tryptophan-binding pocket. Furthermore, the highest-scoring structures place the peptide in the same conformation as the homology models: an antiparallel orientation with the peptide's C-terminus lying between helices III and IV of CaM. These models suggest D6 from TRTK12 is in a somewhat unfavorable environment, as it lies adjacent to similar charges in N-terminal domain of CaM, including Glu14. This is consistent with the enhanced binding affinities observed for the TRTKM6 (D6A) variant (Table II).

Figure 4.

Figure 4

Panel A: Homology model of the TRTK12 peptide (pink) in complex with the N-terminal domain of CaM (orange) generated using HADDOCK. This structure is superimposed with the complex of the NSCaTE peptide (cyan) binding to the N-terminal domain of CaM (PDB: 2LQC27). Panel B: An alternate view of the structures shown in Panel A. The tryptophan side chain of both peptides is shown as sticks for comparison. Panel C: Homology model of the TRTK12 peptide (pink) in complex with the C-terminal domain of CaM (red) generated using HADDOCK. This structure is superimposed with the complex of the C20W peptide (cyan) binding to the C-terminal domain of CaM (PDB: 1CFF30). Panel D: An alternate view of the structures shown in Panel C. The tryptophan side chain of both peptides is shown as sticks for comparison.

Homology models of TRTK12 in complex with the C-terminal domain of CaM also produced a consistent set of structures. Homology models based on the CaM-C20W (PDB: 1CFF30) and CaM-ER-α (PDB: 2LLQ6) complexes suggest that the TRTK12 peptides lies in the C-terminal domain with the peptide's N-terminus pointing away from the center of the protein and the peptide's C-terminus lying between helices V and IV. This orientation is also consistent with the highest-scoring complexes generated using HADDOCK (see Materials and Methods). In these structures, W7 of TRTK12 is buried in the canonical tryptophan-binding pocket of CaM (Fig. 4), consistent with the significant chemical-shift perturbations observed for this cluster of nonpolar residues in the C-terminal domain. In addition, this arrangement allows for several favorable electrostatic interactions, placing the R2 and K4 residues from TRTK12 within range of several acidic residues from CaM, including Asp80, Glu84, and Glu127. Interestingly, this proposed conformation of the bound state aligns D6 of the peptide near several similar charges on CaM, including Glu123 and Glu127. As with the N-terminal domain, this orientation is consistent with the tighter dissociation constants observed for the TRTKM6 variant, which substitutes D6 with an alanine.

The existence of several charge–charge interactions, as well as the burial of some polar residues from TRTK12, may explain the smaller magnitude of ΔCp (−1.3 to −0.8 kJ mol−1 K−1 Table II) observed for binding to the individual CaM domains as opposed to the canonical binding of 1 peptide per CaM (−3.7 to −2.7 kJ mol−1 K−1 see Ref.8). Furthermore, the available structures of the TRTK12 peptide interacting with the closely-related S100 proteins suggests that only a portion of the peptide directly interacts with its binding partners, leaving the remaining sequence solvent exposed.31,32 Nonetheless, homology models of the CaM-TRTK12 complex suggest that the total amount of surface area buried per domain, 1600 ± 100 Å2, is similar to what has been observed for several other CaM-binding targets, such as the NSCaTE27 and MLC8 peptides. Several bulky, hydrophobic side chains of the TRTK12 peptide may anchor the peptide to CaM, contributing to the significant amount of buried surface area. This is consistent with the significantly weaker dissociation constants observed for the TRTKM5 (I5A), TRTKM10 (I10A), and TRTKM11 (I11A) variants (Table II). It is also consistent with the highest-scoring structures generated by HADDOCK, where I5, I10, and L11 are buried against hydrophobic residues in the binding pockets of both domains (Fig. 4). It is important to note that the conformation of the TRTK12 peptide in the bound form remains to be determined. Homology modeling suggests that the peptide in the bound form should adopt helical conformation, while the HADDOCK model, generated based on available constraints, suggests a more extended conformation.

Structural aspects of p53 peptide binding to CaM

The following CaM residues experienced significant changes in chemical shift on p53 peptide binding: Phe19, Asp50, Thr70, Ala73, Asp80, Ile85, Ala88, Val91, Phe92, Asp93, Asn97, Leu105, Val108, Met109, Thr110, Asn111, Lys115, Val121, Glu127, Ala128, Asp129, Phe141, Val142, Gln143, Met144, Met145, Thr146, Ala147, and Lys148. The vast majority of residues found to interact with p53 appear to structurally cluster to a region in the C-terminal domain of CaM. The only exception is residue F19 from the N-terminal domain, which appears to interact with p53 in the bound complex, possibly through an interaction with the tail portion of the peptide.

The analysis of ITC binding isotherms, together with the chemical-shift data obtained from NMR, strongly suggest that the p53 peptide binds CaM with a 1:1 stoichiometry. Furthermore, these interactions map to the C-terminal domain of CaM (Fig. 3). Previously, it has been shown that peptides bound to a single domain of CaM have a less negative ΔCp than those bound to both domains via canonical binding.8 This is consistent with the relatively small absolute value obtained for the change in the heat capacity on p53 peptide binding (−0.9 ± 0.1 kJ mol−1 K−1) and the ΔCp value for the noncanonical interactions of several peptides with CaM.8

Interestingly, many of the same CaM residues appear to be involved in binding both the p53 and TRTK12 peptides. This is consistent with the available structures of several targets in complex with the C-terminal domain of CaM alone, such as eNOS,33 the ER-α,6 CaMKI25 and suggests that the binding interface often consists of the same key residues. Within the bound complex, many peptides adopt an antiparallel orientation, with their N-termini pointing away from the center of CaM toward the protein's C-terminal end. However, significant heterogeneity has been observed for several peptides in complex with CaM and conformational changes in both CaM and its binding partners appear to be very sensitive to slight differences in the target sequences.26,34,35 Furthermore, several peptides have been observed to bind CaM in both parallel and antiparallel orientations, particularly when they contain so called “palindromic sequences.”36 This may be the case for the p53 peptide, which contains large hydrophobic anchors (L4 and F20) at both its N- and C-termini, but does not contain an intervening tryptophan. In addition, this peptide contains a large number of lysine, arginine, and histidine residues regularly spaced throughout its core. The inherent heterogeneity in the bound complexes suggests that identifying the residues of the binding interface of CaM may be more significant than the particular orientation of a single, bound conformation.

Line-shape analysis and the kinetics of peptide binding

For the TRTK12 peptide titration, residues that underwent significant changes in their backbone chemical shifts could be separated into two groups based on their individual titration profiles. One group includes residues Phe16, Phe19, Gly33, Ala57, Asp64, and Phe68 while the other includes residues Val91, Ala102, Leu105, Ala128, Ile130, Gln143, Ala147, and Lys148. Importantly, the residues from the first group map to a distinct location on the N-terminal domain while residues from the second group map onto the C-terminal domain of CaM (Fig. 3). One-dimensional (1D) line shapes extracted from the series of 2D HSQC spectra were analyzed using LineShapeKin (Material and Methods) to obtain the koff rate constants for the TRTK12 peptides' dissociation from CaM. The quality of the fits from LineShapeKin (see Materials and Methods), as determined by both χ2 analysis and visual inspection, suggests that the dissociation constants for the C-terminal and N-terminal domains are approximately 3 ± 1 and 12 ± 1 µM at 35°C, respectively (see Supporting Information Fig. S3). These are in excellent agreement with the dissociation constants obtained using ITC at 35°C: 5 ± 1 and 15 ± 1 µM, respectively. The kinetic parameters of binding obtained from the fits are as follows: koff1 = 1400 ± 210 s−1 and kon1 = (4.5 ± 1.7) × 108 M−1 s−1 for the C-terminal domain and koff2 = 570 ± 120 s−1 and kon2 = (4.8 ± 1.1) × 107 M−1 s−1 for the N-terminal domain.

For the p53 titration, all CaM residues that experienced significant chemical-shift perturbations exhibited a monophasic dependence on the peptide concentration throughout the titration. This is consistent with the stoichiometry observed in the ITC titration (i.e., 1:1). Binding was observed to occur in the intermediate to fast exchange regime (koff ≥ Δω), which allowed the position-averaged chemical shifts of the peptide free and peptide-bound forms to be easily tracked throughout the titration (see Supporting Information Fig. S3). The apparent dissociation constant calculated from line-shape analysis, using residues Asp93, Leu105, Ala128, Phe141, and Lys148, lies between 0.3 and 0.6 µM and is similar to 2 ± 1 µM obtained from ITC. The fitted data resulted in the same koff value, 200 ± 40 s−1, whether the Kd value was simultaneously fit with koff or whether Kd was held constant using the value obtained from ITC. As the dissociation constants obtained from ITC have been proposed to be more accurate than those directly obtained from line-shape analysis,37 the Kd from ITC was used to calculate the apparent kon rate constant: (9 ± 5) ×107 M−1 s−1.

The koff rate constants determined using line-shape analysis for both TRTK12 and p53 are relatively fast compared to other CaM-binding peptides. For targets with nanomolar binding affinities to CaM, koff is generally observed to lie between 0.1 and 10 s−1 with associated kon rate constants ≥1 × 108 M−1 s−1.29,38 There is limited kinetic data available for peptides that bind CaM with micromolar affinities, but koff for at least one peptide has been observed to be greater than or equal to 400 s−1.39

The fast bimolecular rate constants observed in this study may suggest a mechanism of binding. It is well-established that the TRTK12 and p53 peptides lack significant secondary structure in solution and, therefore, represent intrinsically disordered peptides (IDP's).18 In addition, these peptides likely adopt a helical conformation in the bound state, given that this is the predominant structural motif for CaM-binding targets1 and that these peptides contain significant helical content when in complex with the S100 proteins.31,32,40 Previous work has demonstrated that the coupled binding and folding process for IDP's generally proceeds via one of two mechanisms: conformational selection (binding after folding) or “fly-casting,” (folding after binding).41 It has been suggested that these mechanisms can be discriminated in the case of IDP binding if the apparent kon is sufficiently fast, ≥1 × 107 M−1 s−1, due to the dependence of this rate constant on the folding equilibrium.41 The kon (9 ± 5) ×108 M−1 s−1 for p53 and (4.5 ± 1.7) ×108 M−1 s−1 for TRTK12 binding to the C-terminus of CaM exceed this threshold by more than an order of magnitude, which suggests they may indeed bind via a “fly-casting” mechanism. The presence of certain kinetic intermediates is also considered to be direct evidence of a “folding after binding” mechanism.41,42 Although not observed in this study, a variety of CaM-binding targets have been observed to associate in a biphasic manner.38,42 This has been proposed to include a fast phase, where an initial encounter complex is formed, and a slower phase, where the target rearranges to exclude water, bury several ion pairs, and often forms an alpha helix.42 These data are consistent with previous suggestions that the “fly-casting” mechanism may be the predominant binding mode for IDP's and their binding partners.4345 The results presented here provide additional experimental evidence that a large number of unstructured ligands interact with their protein receptors through a “fly-casting” mechanism.18,43,44

Target overlap between CaM and the S100 proteins

Although more than 300 target sequences are known to bind CaM, the vast majority can be classified as belonging to one of four binding motifs: 1-14, 1-10, 1-16, and the IQ motif.1 The specific spacing of bulky hydrophobic groups, positively charged residues, and/or polar glutamine residues in these motifs is proposed to be crucial for the high affinity often observed for CaM binding partners. In addition, these binding epitopes are generally unstructured in solution, but adopt a helical structure in the bound complex.1 Interestingly, the sequences of TRTK12 and p53 peptides do not fit into any of the aforementioned binding motifs despite their net positive charge, overall hydrophobicity, and lack of structure in solution. A search of the CaM Target Database produced no results, that is, no similar sequences were found in the database of known targets.1 As such, both peptides appear to be novel in vitro binding partners for CaM. Furthermore, both peptides produced a maximum, normalized score of 0 for all residues using the putative CaM binding site analysis.1

Canonical sequence recognition by CaM is characterized by a 1:1 stoichiometry, with both the N- and C-terminal domains wrapping around an α-helical peptide/protein.1 However, several studies have also identified target peptide sequences that interact with CaM by only binding to the N- or C-terminal domain.27,28,30 In addition, peptides have been observed to bind both domains individually, sometimes with micromolar affinities.7,46,47 Importantly, the interactions between CaM and full-length proteins have been shown to depend on the presence of short, linear sequence motifs within its binding partners.1 Therefore, using peptides that represent these sequences to study CaM binding is a well-established approach. The dissociation constants obtained for the TRTK12 (3 ± 1 µM and 14 ± 2 µM) and p53 (2 ± 1 µM) peptides are similar to the dissociation constants for other biologically relevant, CaM-binding partners. Furthermore, the dissociation constants appear to be independent of the binding mode, as they are similar for both canonical (e.g., α-synuclein 35 ± 10 µM48) and noncanonical (e.g., caldesmon 1.4 ± 0.1 µM,8,49 NSCaTE 0.1–2.0 µM7) target interactions. In addition, the binding affinity of the p53 peptide for CaM in the presence of physiological ionic strength is significantly tighter than those observed for several of the S100 proteins, which have been shown to interact with this sequence in vivo.16,50,51 Similarly, the binding affinities of the TRTK12 peptides for CaM are similar to those observed for S100B (Kd= 2 ± 0.1 µM), which has been shown to bind both the TRTK12 peptide and the full-length CapZ protein.12 Lastly, CaM has been shown to bind to proteins that perform similar functions as the p53 and CapZ proteins. These include NAP-22,52 p68,53 and adducin.54 Taken together, this suggests that the interactions of the p53 and TRTK12 peptides with CaM may be physiologically relevant. However, it should be noted that additional studies will be necessary to substantiate these interactions in vivo with the full-length p53 and CapZ proteins.

It is interesting that the TRTK12 and p53 peptides bind not only to CaM but also to several members of the S100 protein family.12,15,16,18 Early work9,55 on these proteins sought to differentiate them from CaM and related calcium-binding proteins. In particular, they were shown to exhibit different subcellular localization and oligomerization, unique responses to differentiation, distinct binding clefts, and to interact with a variety of distinct intracellular and extracellular targets.911,20,56 However, the functional relationship between CaM and the S100 proteins remains unclear. It is now thought that the S100 proteins arose through a gene duplication of C-terminal domain of CaM, with subsequent perturbation of the calcium binding affinity in the remaining N-terminal EF-hand.57 Phylogenetic analysis of the human S100 proteins shows they consist of four subgroups: A2/A3/A4/A5/A6, A13/A14/A16, A7/A8/A9/A12/G, and A1/A10/A11/B/P/Z.57 Pairwise sequence alignments of the S100 proteins and CaM show that the latter subgroup, which consists of A1/A10/A11/B/P/Z, shares the largest degree of sequence similarity with CaM (Fig. 6). This is also evident from comparisons of the binding site residues of these proteins (Supporting Information Table S1), which show that the A1/A10/A11/B/P/Z and A2/A3/A4/A5/A6 subgroups contains a similar percent of hydrophobic, polar, and acidic surface area as CaM, relative to other S100 proteins.20

Figure 6.

Figure 6

Percent identity matrix of 16 representative human S100 proteins and the C-terminal domain of CaM generated using ClustalOmega.58 CaM shares the largest percent identity with members of the ancestral S100 proteins: S100A1, S100A11, S100B, S100P, and S100Z.

These sequence homologies may explain the similarity in plasticity of binding that has been observed for these proteins. In particular, CaM has been shown to interact with its binding partners with different stoichiometries, multiple target orientations, large variations in affinity, and using one or both domains.1,4,7,8,59 Similar binding plasticity has been observed for the S100 proteins.10,18,60,61 For example, a recently published structure indicates that S100A4 is able to interact with nonmuscle myosin IIA tail in an asymmetric manner, where a single peptide simultaneously interacts with both canonical binding sites and stretches across the S100A4 dimer interface.60 Likewise, S100A1, S100A2, and S100A6 have been reported to interact with peptides derived from the MDM2 proteins through a sequential binding model with strong, negative cooperativity.61 Atypical binding has also been observed for the closely-related S100B18 and S100A10/A11 proteins.10 In addition, several studies have noted overlapping target specificity between CaM and the S100 proteins, particularly those derived from the S100B branch of the phylogenetic tree. The overlapping targets include melittin,8,62 caldesmon,8,63 tau,64,65 TRPM3,66 IQGAP1,67,68 bHLH transcription factors,69,70 p37,71,72 NDR,18,73 neuromodulin,14,74 the ryanodine receptor,75 twitchin kinase,76,77 tropomyosin,78 and now the TRTK12 and p53 peptides. Taken together, these results raise the question of whether or not a functional overlap may exist between CaM and the S100 proteins. Whether these interactions occur in vivo remains to be tested.

Materials and Methods

Protein purification

Full-length human CaM and representative members of the human S100 protein family, including S100A1, S100A2, S100A3, S100A4, S100A5, S100A6, S100A7, S100A8, S100A9, S100A10, S100A11, S100A12, S100A13, S100P, S100Z, and S100B, were overexpressed in Escherichia coli and purified as previously described.8,10,18 The molar extinction coefficients at 280 nm (ε280) were calculated as previously described.15 The N-terminal domain of human calmodulin, CaM(1–80), was transformed into BL21 (DE3) cells using the PT7-7 bacterial vector and purified as previously described.79 The sequence of CaM(1–80) contains no tryptophans, tyrosines, or cysteine residues, so the concentration was measured using the extinction coefficient for the peptide backbone determined at 205 nm:80

Peptide purification

All peptides were synthesized at the Penn State College of Medicine Macromolecular Core Facility using standard Fmoc chemistry with their N- and C-termini acetylated and amidated, respectively. The C-terminal p53 peptide [p53(367–388); amino acid sequence, Ac-YSHLKSKKGQSTSRHKKLMFKTE-Am], and the TRTK12 wild-type peptide [Cap-Z(265–276); amino acid sequence, Ac-TRTKIDWNKILS-Am], derived from the actin-binding protein, CapZ, have previously been used in structural studies.31,32,8183 For clarification, the tryptophan located at the seventh position in the TRTK12 peptide sequence [Cap-Z(271)] will be referred to as “W7.” The alanine variants of this peptide sequence, including TRTK12M1 (Ac-ARTKIDWNKILS-Am), TRTK12M5 (Ac-TRTKADWNKILS-Am), TRTK12M6 (Ac-TRTKIAWNKILS-Am), TRTK12M10 (Ac-TRTKIDWNKALS-Am), and TRTK12M11 (Ac-TRTKIDWNKIAS-Am) were previously studied by this lab.15 These variants were designed based on the hydrophobic and electrostatic interactions reported in the literature between the wild-type TRTK12 peptide and the S100B and S100A1 proteins.31,32,81,82 The peptide concentrations were determined using molar extinction coefficients (ε280) of 5500 M−1cm−1 for the TRTK12 wild-type and alanine variants and 1490 M−1 cm−1 for the p53 peptide.84

Isothermal titration calorimetry

ITC measurements were performed using a VP-ITC instrument (MicroCal, Northhampton, MA) as previously described.10,18,62,85 The resulting titration curves were analyzed using the Origin scripts for ITC data analysis provided by MicroCal. The data were analyzed using a variety of binding models, in the order of increasing complexity, as described previously.15 The simplest model that adequately fit the data was accepted.

NMR spectroscopy

NMR spectra were collected at 25 or 35°C on a Bruker AVANCE II 600 or 800 MHz spectrometer equipped with a triple-resonance cryoprobe with z-axis gradients and processed using the Bruker software suite TopSpin (version 2.1). Backbone resonance assignments were confirmed for human CaM in the peptide-free state using 1H-15N HSQC and 15N-separated 3D TOCSY experiments and previously published assignments as input.5,25 Reassignment of the amide groups in the peptide-bound state were determined by monitoring changes in chemical shifts throughout the peptide titrations in a series of 2D 1H-15N HSQC spectra. The experimental buffer consisted of 10 mM Tris, 5 mM CaCl2, 15 mM NaCl, and 0.34 mM NaN3 and 10% D2O (pH 7.5). Chemical-shift mapping experiments were performed by titrating 0–1 mM peptide into 0.1–0.4 mM calcium-bound CaM and monitoring changes in backbone 15N and 1H. Titrations were continued until there were no observable changes in the chemical shifts. The combined perturbation of the proton and nitrogen chemical shifts, ΔCS, was quantified according to the following weighted metric:18,86

graphic file with name pro0023-1247-mu10.jpg (1)

where δN and δH are the changes in chemical shift (parts per million) in the nitrogen and proton dimensions, respectively. Spectra were analyzed using Sparky.87 Chemical-shift perturbations were considered to be significant if they were greater than a weighted average of 0.3.

LinShapeKin analysis

Line-shape analysis was performed using the BiophysicsLab Matlab package,88 as previously described.18 The 1D line shapes in the 15N and 1H dimensions from 2D HSQCs were extracted in Sparky87,89 and the intensities were normalized to account for exchange broadening in the other dimension.88 Titration data for the p53 peptide were fit to a model in which there is one binding site per CaM. Titration data for the TRTK12 peptide were analyzed independently for each of the two binding sites of CaM (see Supporting Information Methods for details). The LineShapeKin software simultaneously solves for both the apparent binding constant, Kd, and the off rate, koff, analytically, which allows the on rate, kon, to be easily calculated. As previously described,18 amide peaks that underwent significant chemical-shift changes during the titration and met the necessary criteria were selected for analysis. Structurally-clustered, residue-specific data were assumed to be reporting on the same binding event. This assumption was validated by comparing the fits of individual residues with each other and the global fit for structurally-clustered residues, which were in good agreement. All reported values are the result of global fits, which include both 15N and 1H dimensions of all relevant amide groups across all ligand concentrations.

Structure-based calculations of ΔCp

Using the available 3D structures of CaM in the calcium bound state and/or in complex with a peptide target, ΔASAtot was calculated as previously described.22 Homology models of the peptide-bound state were generated using Modeller90 and PDB entries 1CDL,91 1CFF,30 2LLO,6 2LLQ,6 and 2LQC.27 Changes in the accessible surface area on binding, ΔASAtotal, were calculated as previously described.10,18,62 Changes in ΔASA were further subdivided into four categories (aliphatic surface area, aromatic surface area, peptide backbone surface area, and polar surface area) and converted into ΔCp using the following empirical relationship:8,10,18,22,92

graphic file with name pro0023-1247-mu11.jpg (2)

where ΔASAalp, ΔASAarm, and ΔASApol are the changes in ASA for aliphatic, aromatic, and polar amino acids, respectively, and ΔASAbb is the change in ASA for the polypeptide backbone.

HADDOCK docking

To further explore the interaction between the p53 and TRTK12 peptides with CaM, a series of docked complexes were generated using HADDOCK.93,94 The p53 (PDB: 1DT7) and TRTK12 (PDB: 1MWN) structures were extracted from existing complexes and docked with calcium-bound, full-length CaM (PDB: 2M55 or 1CFF), or the individual calcium-bound N-terminal (PDB: 2M55, residues 1–78, or PDB: 1CFF, residues 1–78) and C-terminal (PDB: 2LQP) domains. Ambiguous interactions restraints were defined for CaM by the residues that underwent significant backbone chemical-shift perturbations, as monitored by 15N/1H heteronuclear HSQCs, during the peptide titrations. The restrains for the p53 peptide were defined by the putative hydrophobic anchors based on sequence alignments (Supporting Information Fig. S4). For the TRTK12 peptide, they were defined by the residues that appeared to significantly weaken the dissociation constant on alanine mutagenesis (i.e., I5, I10, and L11). The best 200 structures from each run were clustered and scored according to the HADDOCK algorithm using the default options.93,94 All 200 structures were further evaluated based on their ability to recapitulate the experimentally-obtained ΔCp values according to Eq. (2). This is, to the best of our knowledge, the first case when structural thermodynamic information has been included into refinement. The 10 best structures were chosen and further analyzed based on their ability to recapitulate the observed chemical-shift perturbations, the position of the putative hydrophobic anchors, and, for the CaM-TRTK12 complexes, the relative exposure of the peptide's tryptophan (W7). A representative structure for the binding of each TRTK12 peptide to individual domains of CaM was selected for illustrative purposes.

Acknowledgments

The authors thank Dr. Thomas C Squier for the gift of plasmid expressing mammalian calmodulin and Dr. Madeline Shea for providing us with the plasmid for WT rCaM1–80 in the PT7-7 vector. The authors also thank Calvin Chen and Anthony Bishop for their assistance with the HADDOCK docking and automated analysis. Instrumentation at the Core Facilities at the Center of Biotechnology and Interdisciplinary Studies at RPI were used for some of the experiments reported in this paper.

Glossary

CaM

human calmodulin

EF-hand

calcium-binding domain consisting of a helix-loop-helix structure

Fmoc

fluorenylmethyloxycarbonyl chloride

ITC

isothermal titration calorimetry

TCEP

tris(2-carboxyethyl)phosphine; HSQC, Heteronuclear Single Quantum Coherence, MLC, C-terminal fragment of melittin.

Supporting Information

Additional Supporting Information may be found in the online version of this article.

Supplementary Information

pro0023-1247-sd1.pdf (730.3KB, pdf)

References

  • 1.Yap KL, Kim J, Truong K, Sherman M, Yuan T, Ikura M. Calmodulin target database. J Struct Funct Genomics. 2000;1:8–14. doi: 10.1023/a:1011320027914. [DOI] [PubMed] [Google Scholar]
  • 2.Martinez J, Moeller I, Erdjument-Bromage H, Tempst P, Lauring B. Parkinson's disease-associated alpha-synuclein is a calmodulin substrate. J Biol Chem. 2003;278:17379–17387. doi: 10.1074/jbc.M209020200. [DOI] [PubMed] [Google Scholar]
  • 3.O'Day DH, Myre MA. Calmodulin-binding domains in Alzheimer's disease proteins: extending the calcium hypothesis. Biochem Biophys Res Commun. 2004;320:1051–1054. doi: 10.1016/j.bbrc.2004.06.070. [DOI] [PubMed] [Google Scholar]
  • 4.Yamniuk AP, Vogel HJ. Calmodulin's flexibility allows for promiscuity in its interactions with target proteins and peptides. Mol Biotechnol. 2004;27:33–57. doi: 10.1385/MB:27:1:33. [DOI] [PubMed] [Google Scholar]
  • 5.Chou JJ, Li S, Klee CB, Bax A. Solution structure of Ca(2+)-calmodulin reveals flexible hand-like properties of its domains. Nat Struct Biol. 2001;8:990–997. doi: 10.1038/nsb1101-990. [DOI] [PubMed] [Google Scholar]
  • 6.Zhang Y, Li Z, Sacks DB, Ames JB. Structural basis for Ca2+-induced activation and dimerization of estrogen receptor alpha by calmodulin. J Biol Chem. 2012;287:9336–9344. doi: 10.1074/jbc.M111.334797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Taiakina V, Boone AN, Fux J, Senatore A, Weber-Adrian D, Guillemette JG, Spafford JD. The calmodulin-binding, short linear motif, NSCaTE is conserved in L-type channel ancestors of vertebrate Cav1.2 and Cav1.3 channels. PLoS One. 2013;8:e61765. doi: 10.1371/journal.pone.0061765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Brokx RD, Lopez MM, Vogel HJ, Makhatadze GI. Energetics of target peptide binding by calmodulin reveals different modes of binding. J Biol Chem. 2001;276:14083–14091. doi: 10.1074/jbc.M011026200. [DOI] [PubMed] [Google Scholar]
  • 9.Zimmer DB, Van Eldik LJ. Analysis of the calcium-modulated proteins, S100 and calmodulin, and their target proteins during C6 glioma cell differentiation. J Cell Biol. 1989;108:141–151. doi: 10.1083/jcb.108.1.141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Streicher WW, Lopez MM, Makhatadze GI. Annexin I and annexin II N-terminal peptides binding to S100 protein family members: specificity and thermodynamic characterization. Biochemistry. 2009;48:2788–2798. doi: 10.1021/bi8019959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Santamaria-Kisiel L, Rintala-Dempsey AC, Shaw GS. Calcium-dependent and -independent interactions of the S100 protein family. Biochem J. 2006;396:201–214. doi: 10.1042/BJ20060195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ivanenkov VV, Jamieson GA, Jr, Gruenstein E, Dimlich RV. Characterization of S-100b binding epitopes. Identification of a novel target, the actin capping protein, CapZ. J Biol Chem. 1995;270:14651–14658. doi: 10.1074/jbc.270.24.14651. [DOI] [PubMed] [Google Scholar]
  • 13.Ivanenkov VV, Dimlich RV, Jamieson GA., Jr Interaction of S100a0 protein with the actin capping protein, CapZ: characterization of a putative S100a0 binding site in CapZ alpha-subunit. Biochem Biophys Res Commun. 1996;221:46–50. doi: 10.1006/bbrc.1996.0542. [DOI] [PubMed] [Google Scholar]
  • 14.Wilder PT, Lin J, Bair CL, Charpentier TH, Yang D, Liriano M, Varney KM, Lee A, Oppenheim AB, Adhya S, Carrier F, Weber DJ. Recognition of the tumor suppressor protein p53 and other protein targets by the calcium-binding protein S100B. Biochim Biophys Acta. 2006;1763:1284–1297. doi: 10.1016/j.bbamcr.2006.08.024. [DOI] [PubMed] [Google Scholar]
  • 15.Wafer LN, Tzul FO, Pandharipande PP, Makhatadze GI. Novel interactions of the TRTK12 peptide with S100 protein family members: specificity and thermodynamic characterization. Biochemistry. 2013;52:5844–5856. doi: 10.1021/bi400788s. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Fernandez-Fernandez MR, Rutherford TJ, Fersht AR. Members of the S100 family bind p53 in two distinct ways. Protein Sci. 2008;17:1663–1670. doi: 10.1110/ps.035527.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Wilder PT, Rustandi RR, Drohat AC, Weber DJ. S100B (beta beta) inhibits the protein kinase C-dependent phosphorylation of a peptide derived from p53 in a Ca2+-dependent manner. Protein Sci. 1998;7:794–798. doi: 10.1002/pro.5560070330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Wafer LN, Streicher WW, McCallum SA, Makhatadze GI. Thermodynamic and kinetic analysis of peptides derived from CapZ, NDR, p53, HDM2, and HDM4 binding to human S100B. Biochemistry. 2012;51:7189–7201. doi: 10.1021/bi300865g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Ashcroft M, Kubbutat MH, Vousden KH. Regulation of p53 function and stability by phosphorylation. Mol Cell Biol. 1999;19:1751–1758. doi: 10.1128/mcb.19.3.1751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Bhattacharya S, Bunick CG, Chazin WJ. Target selectivity in EF-hand calcium binding proteins. Biochim Biophys Acta. 2004;1742:69–79. doi: 10.1016/j.bbamcr.2004.09.002. [DOI] [PubMed] [Google Scholar]
  • 21.van Dieck J, Fernandez-Fernandez MR, Veprintsev DB, Fersht AR. Modulation of the oligomerization state of p53 by differential binding of proteins of the S100 family to p53 monomers and tetramers. J Biol Chem. 2009;284:13804–13811. doi: 10.1074/jbc.M901351200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Makhatadze GI, Privalov PL. Energetics of protein structure. Adv Protein Chem. 1995;47:307–425. doi: 10.1016/s0065-3233(08)60548-3. [DOI] [PubMed] [Google Scholar]
  • 23.Tsvetkov PO, Protasevich II, Gilli R, Lafitte D, Lobachov VM, Haiech J, Briand C, Makarov AA. Apocalmodulin binds to the myosin light chain kinase calmodulin target site. J Biol Chem. 1999;274:18161–18164. doi: 10.1074/jbc.274.26.18161. [DOI] [PubMed] [Google Scholar]
  • 24.Larsson G, Schleucher J, Onions J, Hermann S, Grundstrom T, Wijmenga SS. A novel target recognition revealed by calmodulin in complex with the basic helix–loop–helix transcription factor SEF2-1/E2-2. Protein Sci. 2001;10:169–186. doi: 10.1110/ps.28401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Gifford JL, Ishida H, Vogel HJ. Fast methionine-based solution structure determination of calcium-calmodulin complexes. J Biomol NMR. 2011;50:71–81. doi: 10.1007/s10858-011-9495-3. [DOI] [PubMed] [Google Scholar]
  • 26.Newman RA, Van Scyoc WS, Sorensen BR, Jaren OR, Shea MA. Interdomain cooperativity of calmodulin bound to melittin preferentially increases calcium affinity of sites I and II. Proteins. 2008;71:1792–1812. doi: 10.1002/prot.21861. [DOI] [PubMed] [Google Scholar]
  • 27.Liu Z, Vogel HJ. Structural basis for the regulation of L-type voltage-gated calcium channels: interactions between the N-terminal cytoplasmic domain and Ca(2+)-calmodulin. Front Mol Neurosci. 2012;5:38. doi: 10.3389/fnmol.2012.00038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Rainaldi M, Yamniuk AP, Murase T, Vogel HJ. Calcium-dependent and -independent binding of soybean calmodulin isoforms to the calmodulin binding domain of tobacco MAPK phosphatase-1. J Biol Chem. 2007;282:6031–6042. doi: 10.1074/jbc.M608970200. [DOI] [PubMed] [Google Scholar]
  • 29.Bayley PM, Findlay WA, Martin SR. Target recognition by calmodulin: dissecting the kinetics and affinity of interaction using short peptide sequences. Protein Sci. 1996;5:1215–1228. doi: 10.1002/pro.5560050701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Elshorst B, Hennig M, Forsterling H, Diener A, Maurer M, Schulte P, Schwalbe H, Griesinger C, Krebs J, Schmid H, Vorherr T, Carafoli E. NMR solution structure of a complex of calmodulin with a binding peptide of the Ca2+ pump. Biochemistry. 1999;38:12320–12332. doi: 10.1021/bi9908235. [DOI] [PubMed] [Google Scholar]
  • 31.Wright NT, Cannon BR, Wilder PT, Morgan MT, Varney KM, Zimmer DB, Weber DJ. Solution structure of S100A1 bound to the CapZ peptide (TRTK12) J Mol Biol. 2009;386:1265–1277. doi: 10.1016/j.jmb.2009.01.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Charpentier TH, Thompson LE, Liriano MA, Varney KM, Wilder PT, Pozharski E, Toth EA, Weber DJ. The effects of CapZ peptide (TRTK-12) binding to S100B-Ca2+ as examined by NMR and X-ray crystallography. J Mol Biol. 2010;396:1227–1243. doi: 10.1016/j.jmb.2009.12.057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Piazza M, Futrega K, Spratt DE, Dieckmann T, Guillemette JG. Structure and dynamics of calmodulin (CaM) bound to nitric oxide synthase peptides: effects of a phosphomimetic CaM mutation. Biochemistry. 2012;51:3651–3661. doi: 10.1021/bi300327z. [DOI] [PubMed] [Google Scholar]
  • 34.Herbst S, Maucher D, Schneider M, Ihling CH, Jahn O, Sinz A. Munc13-like skMLCK variants cannot mimic the unique calmodulin binding mode of Munc13 as evidenced by chemical cross-linking and mass spectrometry. PLoS One. 2013;8:e75119. doi: 10.1371/journal.pone.0075119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Nagulapalli M, Parigi G, Yuan J, Gsponer J, Deraos G, Bamm VV, Harauz G, Matsoukas J, de Planque MR, Gerothanassis IP, Babu MM, Luchinat C, Tzakos AG. Recognition pliability is coupled to structural heterogeneity: a calmodulin intrinsically disordered binding region complex. Structure. 2012;20:522–533. doi: 10.1016/j.str.2012.01.021. [DOI] [PubMed] [Google Scholar]
  • 36.Irene D, Huang JW, Chung TY, Li FY, Tzen JT, Lin TH, Chyan CL. Binding orientation and specificity of calmodulin to rat olfactory cyclic nucleotide-gated ion channel. J Biomol Struct Dyn. 2013;31:414–425. doi: 10.1080/07391102.2012.703069. [DOI] [PubMed] [Google Scholar]
  • 37.Kovrigin EL. 2012. NMR line shape analysis software for studies of protein-ligand interaction kinetics. Milwaukee. http://lineshapekin.net/LineShapeKin_ver3.1/LineShapeKin/Docs/LINESHAPEKIN_ver.3.1.pdf.
  • 38.Brown SE, Martin SR, Bayley PM. Kinetic control of the dissociation pathway of calmodulin-peptide complexes. J Biol Chem. 1997;272:3389–3397. doi: 10.1074/jbc.272.6.3389. [DOI] [PubMed] [Google Scholar]
  • 39.Kleerekoper QK, Putkey JA. PEP-19, an intrinsically disordered regulator of calmodulin signaling. J Biol Chem. 2009;284:7455–7464. doi: 10.1074/jbc.M808067200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Rustandi RR, Baldisseri DM, Weber DJ. Structure of the negative regulatory domain of p53 bound to S100B(betabeta) Nat Struct Biol. 2000;7:570–574. doi: 10.1038/76797. [DOI] [PubMed] [Google Scholar]
  • 41.Kiefhaber T, Bachmann A, Jensen KS. Dynamics and mechanisms of coupled protein folding and binding reactions. Curr Opin Struct Biol. 2011;22:21–29. doi: 10.1016/j.sbi.2011.09.010. [DOI] [PubMed] [Google Scholar]
  • 42.Kranz JK, Flynn PF, Fuentes EJ, Wand AJ. Dissection of the pathway of molecular recognition by calmodulin. Biochemistry. 2002;41:2599–2608. doi: 10.1021/bi011818f. [DOI] [PubMed] [Google Scholar]
  • 43.Sugase K, Dyson HJ, Wright PE. Mechanism of coupled folding and binding of an intrinsically disordered protein. Nature. 2007;447:1021–1025. doi: 10.1038/nature05858. [DOI] [PubMed] [Google Scholar]
  • 44.Trizac E, Levy Y, Wolynes PG. Capillarity theory for the fly-casting mechanism. Proc Natl Acad Sci USA. 2010;107:2746–2750. doi: 10.1073/pnas.0914727107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Shoemaker BA, Portman JJ, Wolynes PG. Speeding molecular recognition by using the folding funnel: the fly-casting mechanism. Proc Natl Acad Sci USA. 2000;97:8868–8873. doi: 10.1073/pnas.160259697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Martin SR, Bayley PM. Regulatory implications of a novel mode of interaction of calmodulin with a double IQ-motif target sequence from murine dilute myosin V. Protein Sci. 2002;11:2909–2923. doi: 10.1110/ps.0210402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Rodney GG, Moore CP, Williams BY, Zhang JZ, Krol J, Pedersen SE, Hamilton SL. Calcium binding to calmodulin leads to an N-terminal shift in its binding site on the ryanodine Receptor. J Biol Chem. 2001;276:2069–2074. doi: 10.1074/jbc.M008891200. [DOI] [PubMed] [Google Scholar]
  • 48.Gruschus JM, Yap TL, Pistolesi S, Maltsev AS, Lee JC. NMR structure of calmodulin complexed to an N-terminally acetylated alpha-synuclein peptide. Biochemistry. 2013;52:3436–3445. doi: 10.1021/bi400199p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Zhou N, Yuan T, Mak AS, Vogel HJ. NMR studies of caldesmon-calmodulin interactions. Biochemistry. 1997;36:2817–2825. doi: 10.1021/bi9625713. [DOI] [PubMed] [Google Scholar]
  • 50.Grigorian M, Lukanidin E. [Activator of metastasis in cancer cells, Mst1/S100A4 protein binds to tumor suppressor protein p53] Genetika. 2003;39:900–908. [PubMed] [Google Scholar]
  • 51.Lin J, Yang Q, Yan Z, Markowitz J, Wilder PT, Carrier F, Weber DJ. Inhibiting S100B restores p53 levels in primary malignant melanoma cancer cells. J Biol Chem. 2004;279:34071–34077. doi: 10.1074/jbc.M405419200. [DOI] [PubMed] [Google Scholar]
  • 52.Odagaki S, Kumanogoh H, Nakamura S, Maekawa S. Biochemical interaction of an actin-capping protein, CapZ, with NAP-22. J Neurosci Res. 2009;87:1980–1985. doi: 10.1002/jnr.22040. [DOI] [PubMed] [Google Scholar]
  • 53.Wang H, Gao X, Yang JJ, Liu ZR. Interaction between p68 RNA helicase and Ca2+-calmodulin promotes cell migration and metastasis. Nat Commun. 2013;4:1354. doi: 10.1038/ncomms2345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kuhlman PA, Hughes CA, Bennett V, Fowler VM. A new function for adducin. Calcium/calmodulin-regulated capping of the barbed ends of actin filaments. J Biol Chem. 1996;271:7986–7991. doi: 10.1074/jbc.271.14.7986. [DOI] [PubMed] [Google Scholar]
  • 55.Tabuchi K, Ohnishi R, Nishimoto A, Isobe T, Okuyama T. Reverse cellular distribution of calmodulin to S-100 protein in primate brain. Brain Res. 1984;298:353–357. doi: 10.1016/0006-8993(84)91436-7. [DOI] [PubMed] [Google Scholar]
  • 56.Heizmann CW, Cox JA. New perspectives on S100 proteins: a multi-functional Ca(2+)-, Zn(2+)- and Cu(2+)-binding protein family. Biometals. 1998;11:383–397. doi: 10.1023/a:1009212521172. [DOI] [PubMed] [Google Scholar]
  • 57.Zimmer DB, Eubanks JO, Ramakrishnan D, Criscitiello MF. Evolution of the S100 family of calcium sensor proteins. Cell Calcium. 2013;53:170–179. doi: 10.1016/j.ceca.2012.11.006. [DOI] [PubMed] [Google Scholar]
  • 58.Sievers F, Wilm A, Dineen D, Gibson TJ, Karplus K, Li W, Lopez R, McWilliam H, Remmert M, Soding J, Thompson JD, Higgins DG. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol Sys Biol. 2011;7:539. doi: 10.1038/msb.2011.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Wang E, Zhuang S, Kordowska J, Grabarek Z, Wang CL. Calmodulin binds to caldesmon in an antiparallel manner. Biochemistry. 1997;36:15026–15034. doi: 10.1021/bi963075h. [DOI] [PubMed] [Google Scholar]
  • 60.Kiss B, Duelli A, Radnai L, Kekesi KA, Katona G, Nyitray L. Crystal structure of the S100A4-nonmuscle myosin IIA tail fragment complex reveals an asymmetric target binding mechanism. Proc Natl Aad Sci USA. 2012;109:6048–6053. doi: 10.1073/pnas.1114732109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.van Dieck J, Lum JK, Teufel DP, Fersht AR. S100 proteins interact with the N-terminal domain of MDM2. FEBS Lett. 2010;584:3269–3274. doi: 10.1016/j.febslet.2010.06.024. [DOI] [PubMed] [Google Scholar]
  • 62.Gribenko AV, Guzman-Casado M, Lopez MM, Makhatadze GI. Conformational and thermodynamic properties of peptide binding to the human S100P protein. Protein Sci. 2002;11:1367–1375. doi: 10.1110/ps.0202202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Polyakov AA, Huber PA, Marston SB, Gusev NB. Interaction of isoforms of S100 protein with smooth muscle caldesmon. FEBS Lett. 1998;422:235–239. doi: 10.1016/s0014-5793(98)00014-3. [DOI] [PubMed] [Google Scholar]
  • 64.Yu WH, Fraser PE. S100beta interaction with tau is promoted by zinc and inhibited by hyperphosphorylation in Alzheimer's disease. J Neurosci. 2001;21:2240–2246. doi: 10.1523/JNEUROSCI.21-07-02240.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Lee YC, Wolff J. Calmodulin binds to both microtubule-associated protein 2 and tau proteins. J Biol Chem. 1984;259:1226–1230. [PubMed] [Google Scholar]
  • 66.Holakovska B, Grycova L, Jirku M, Sulc M, Bumba L, Teisinger J. Calmodulin and S100A1 protein interact with N terminus of TRPM3 channel. J Biol Chem. 2012;287:16645–16655. doi: 10.1074/jbc.M112.350686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Heil A, Nazmi AR, Koltzscher M, Poeter M, Austermann J, Assard N, Baudier J, Kaibuchi K, Gerke V. S100P is a novel interaction partner and regulator of IQGAP1. J Biol Chem. 2011;286:7227–7238. doi: 10.1074/jbc.M110.135095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Ho YD, Joyal JL, Li Z, Sacks DB. IQGAP1 integrates Ca2+/calmodulin and Cdc42 signaling. J Biol Chem. 1999;274:464–470. doi: 10.1074/jbc.274.1.464. [DOI] [PubMed] [Google Scholar]
  • 69.Baudier J, Bergeret E, Bertacchi N, Weintraub H, Gagnon J, Garin J. Interactions of myogenic bHLH transcription factors with calcium-binding calmodulin and S100a (alpha alpha) proteins. Biochemistry. 1995;34:7834–7846. doi: 10.1021/bi00024a007. [DOI] [PubMed] [Google Scholar]
  • 70.Onions J, Hermann S, Grundstrom T. Basic helix-loop-helix protein sequences determining differential inhibition by calmodulin and S-100 proteins. J Biol Chem. 1997;272:23930–23937. doi: 10.1074/jbc.272.38.23930. [DOI] [PubMed] [Google Scholar]
  • 71.Dukhanina EA, Dukhanin AS, Lomonosov MY, Lukanidin EM, Georgiev GP. Spectral studies on the calcium-binding properties of Mts1 protein and its interaction with target protein. FEBS Lett. 1997;410:403–406. doi: 10.1016/s0014-5793(97)00576-0. [DOI] [PubMed] [Google Scholar]
  • 72.Alcazar A, Martin ME, Soria E, Rodriguez S, Fando JL, Salinas M. Purification and characterization of guanine nucleotide-exchange factor, eIF-2B, and p37 calmodulin-binding protein from calf brain. J Neurochem. 1995;65:754–761. doi: 10.1046/j.1471-4159.1995.65020754.x. [DOI] [PubMed] [Google Scholar]
  • 73.Bhattacharya S, Large E, Heizmann CW, Hemmings B, Chazin WJ. Structure of the Ca2+/S100B/NDR kinase peptide complex: insights into S100 target specificity and activation of the kinase. Biochemistry. 2003;42:14416–14426. doi: 10.1021/bi035089a. [DOI] [PubMed] [Google Scholar]
  • 74.Chapman ER, Au D, Alexander KA, Nicolson TA, Storm DR. Characterization of the calmodulin binding domain of neuromodulin. Functional significance of serine 41 and phenylalanine 42. J Biol Chem. 1991;266:207–213. [PubMed] [Google Scholar]
  • 75.Wright NT, Prosser BL, Varney KM, Zimmer DB, Schneider MF, Weber DJ. S100A1 and calmodulin compete for the same binding site on ryanodine receptor. J Biol Chem. 2008;283:26676–26683. doi: 10.1074/jbc.M804432200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Heierhorst J, Kobe B, Feil SC, Parker MW, Benian GM, Weiss KR, Kemp BE. Ca2+/S100 regulation of giant protein kinases. Nature. 1996;380:636–639. doi: 10.1038/380636a0. [DOI] [PubMed] [Google Scholar]
  • 77.Wilmann M, Gautel M, Mayans O. Activation of calcium/calmodulin regulated kinases. Cell Mol Biol. 2000;46:883–894. [PubMed] [Google Scholar]
  • 78.Watson MH, Kuhn AE, Mak AS. Caldesmon, calmodulin and tropomyosin interactions. Biochim Biophys Acta. 1990;1054:103–113. doi: 10.1016/0167-4889(90)90211-u. [DOI] [PubMed] [Google Scholar]
  • 79.Sorensen BR, Faga LA, Hultman R, Shea MA. An interdomain linker increases the thermostability and decreases the calcium affinity of the calmodulin N-domain. Biochemistry. 2002;41:15–20. doi: 10.1021/bi011718+. [DOI] [PubMed] [Google Scholar]
  • 80.Scopes RK. Measurement of protein by spectrophotometry at 205 nm. Anal Biochem. 1974;59:277–282. doi: 10.1016/0003-2697(74)90034-7. [DOI] [PubMed] [Google Scholar]
  • 81.Inman KG, Yang R, Rustandi RR, Miller KE, Baldisseri DM, Weber DJ. Solution NMR structure of S100B bound to the high-affinity target peptide TRTK-12. J Mol Biol. 2002;324:1003–1014. doi: 10.1016/s0022-2836(02)01152-x. [DOI] [PubMed] [Google Scholar]
  • 82.McClintock KA, Shaw GS. A novel S100 target conformation is revealed by the solution structure of the Ca2+-S100B-TRTK-12 complex. J Biol Chem. 2003;278:6251–6257. doi: 10.1074/jbc.M210622200. [DOI] [PubMed] [Google Scholar]
  • 83.Rustandi RR, Drohat AC, Baldisseri DM, Wilder PT, Weber DJ. The Ca(2+)-dependent interaction of S100B(beta beta) with a peptide derived from p53. Biochemistry. 1998;37:1951–1960. doi: 10.1021/bi972701n. [DOI] [PubMed] [Google Scholar]
  • 84.Wilkins MR, Gasteiger E, Bairoch A, Sanchez JC, Williams KL, Appel RD, Hochstrasser DF. Protein identification and analysis tools in the ExPASy server. Methods Mol Biol. 1999;112:531–552. doi: 10.1385/1-59259-584-7:531. [DOI] [PubMed] [Google Scholar]
  • 85.Lopez MM, Makhatadze GI. Isothermal titration calorimetry. Methods Mol Biol. 2002;173:121–126. doi: 10.1385/1-59259-184-1:121. [DOI] [PubMed] [Google Scholar]
  • 86.Rule GS, Hitchens TK. Fundamentals of Protein NMR Spectroscopy. NY: Springer; 2005. [Google Scholar]
  • 87.TD Goddard, DG Kneller. SPARKY 3, San Francisco University of California.
  • 88.Kovrigin EL, Loria JP. Enzyme dynamics along the reaction coordinate: critical role of a conserved residue. Biochemistry. 2006;45:2636–2647. doi: 10.1021/bi0525066. [DOI] [PubMed] [Google Scholar]
  • 89.McCallum SA, Hitchens TK, Torborg C, Rule GS. Ligand-induced changes in the structure and dynamics of a human class Mu glutathione S-transferase. Biochemistry. 2000;39:7343–7356. doi: 10.1021/bi992767d. [DOI] [PubMed] [Google Scholar]
  • 90.Eswar N, Webb B, Marti-Renom MA, Madhusudhan MS, Eramian D, Shen MY, Pieper U, Sali A. Comparative protein structure modeling using MODELLER. Curr Protoc Protein Sci Chapter 2. 2007 doi: 10.1002/0471140864.ps0209s50. Unit 2.9. [DOI] [PubMed] [Google Scholar]
  • 91.Meador WE, Means AR, Quiocho FA. Target enzyme recognition by calmodulin: 2.4 A structure of a calmodulin-peptide complex. Science. 1992;257:1251–1255. doi: 10.1126/science.1519061. [DOI] [PubMed] [Google Scholar]
  • 92.Wafer LN, Streicher WW, Makhatadze GI. Thermodynamics of the Trp-cage miniprotein unfolding in urea. Proteins. 2010;78:1376–1381. doi: 10.1002/prot.22681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Dominguez C, Boelens R, Bonvin AM. HADDOCK: a protein-protein docking approach based on biochemical or biophysical information. J Am Chem Soc. 2003;125:1731–1737. doi: 10.1021/ja026939x. [DOI] [PubMed] [Google Scholar]
  • 94.de Vries SJ, van Dijk AD, Krzeminski M, van Dijk M, Thureau A, Hsu V, Wassenaar T, Bonvin AM. HADDOCK versus HADDOCK: new features and performance of HADDOCK2.0 on the CAPRI targets. Proteins. 2007;69:726–733. doi: 10.1002/prot.21723. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary Information

pro0023-1247-sd1.pdf (730.3KB, pdf)

Articles from Protein Science : A Publication of the Protein Society are provided here courtesy of The Protein Society

RESOURCES