Skip to main content
The Journal of Biological Chemistry logoLink to The Journal of Biological Chemistry
. 2016 Mar 31;291(21):11348–11358. doi: 10.1074/jbc.M116.720250

Thermodynamic Basis of Selectivity in the Interactions of Tissue Inhibitors of Metalloproteinases N-domains with Matrix Metalloproteinases-1, -3, and -14*

Haiyin Zou 1, Ying Wu 1,1, Keith Brew 1,2
PMCID: PMC4900279  PMID: 27033700

Abstract

The four tissue inhibitors of metalloproteinases (TIMPs) are potent inhibitors of the many matrixins (MMPs), except that TIMP1 weakly inhibits some MMPs, including MMP14. The broad-spectrum inhibition of MMPs by TIMPs and their N-domains (NTIMPs) is consistent with the previous isothermal titration calorimetric finding that their interactions are entropy-driven but differ in contributions from solvent and conformational entropy (ΔSsolv, ΔSconf), estimated using heat capacity changes (ΔCp). Selective engineered NTIMPs have potential applications for treating MMP-related diseases, including cancer and cardiomyopathy. Here we report isothermal titration calorimetric studies of the effects of selectivity-modifying mutations in NTIMP1 and NTIMP2 on the thermodynamics of their interactions with MMP1, MMP3, and MMP14. The weak inhibition of MMP14 by NTIMP1 reflects a large conformational entropy penalty for binding. The T98L mutation, peripheral to the NTIMP1 reactive site, enhances binding by increasing ΔSsolv but also reduces ΔSconf. However, the same mutation increases NTIMP1 binding to MMP3 in an interaction that has an unusual positive ΔCp. This indicates a decrease in solvent entropy compensated by increased conformational entropy, possibly reflecting interactions involving alternative conformers. The NTIMP2 mutant, S2D/S4A is a selective MMP1 inhibitor through electrostatic effects of a unique MMP-1 arginine. Asp-2 increases reactive site polarity, reducing ΔCp, but increases conformational entropy to maintain strong binding to MMP1. There is a strong negative correlation between ΔSsolv and ΔSconf for all characterized interactions, but the data for each MMP have characteristic ranges, reflecting intrinsic differences in the structures and dynamics of their free and inhibitor-bound forms.

Keywords: calorimetry, conformational change, enzyme inhibitor, matrix metalloproteinase (MMP), protein dynamic, protein engineering, protein-protein interaction, thermodynamics, tissue inhibitor of metalloproteinase (TIMP), zinc

Introduction

Isothermal titration calorimetry can dissect the sources of the free energy changes for protein interactions with other macromolecules and ligands (1). Measurements of the heat released or absorbed during the titration of a protein with a binding partner can quantify the enthalpy of binding (ΔH), association constant Ka, and stoichiometry (N) (1), allowing the calculation of the changes in free energy (ΔG) and entropy (ΔS) of binding. Titrations in buffers with different enthalpies of ionization quantify ionization changes linked to binding, whereas the heat capacity change for binding, ΔCp, measured as the temperature dependence of ΔH, can be used to estimate the change in solvation entropy on binding (ΔSsolv). These parameters can be correlated with the character of the interaction interface and differences in dynamics and solvation between the free and bound conformer populations (24).

Previously, we used isothermal titration calorimetry to elucidate the thermodynamic profiles of interactions between the N-terminal inhibitory domains of tissue inhibitors of metalloproteinases-1 and -2 (NT1 and NT2)3 and the catalytic domains of matrix metalloproteinases -1 and -3 (MMP1c and MMP3c) (5, 6). TIMP1 and TIMP2 are two of the four human TIMPs, endogenous, broad spectrum, slow binding, high affinity inhibitors of the 23 human MMPs and several disintegrin-metalloproteinases (7). The MMPs catalyze the proteolysis of all polypeptide components of the extracellular matrix and are crucial for tissue remodeling, wound healing, embryo implantation, cell migration, shedding of cell surface proteins, and release of bioactive peptides (8, 9); the unregulated activities of various MMPs have been linked to many disease processes including arthritis, heart disease, and tumor metastasis (8).

The two domains of vertebrate TIMPs are each cross-linked by three disulfide bonds. The larger N-terminal domain (NTIMP) carries the MMP inhibitory activity, and the C-terminal domain mediates interactions with other proteins, including some pro-MMPs (7). Recombinant forms of NTIMPs are fully active as MMP inhibitors (10) and have been extensively used in crystallographic (1112), mutational (1318), and NMR studies (10, 1922) of TIMP·MMP interactions. The known structures of complexes between full-length TIMPs and MMP catalytic domains show that the main interaction interface of the TIMP is provided by the N-domain with a few peripheral contacts from the C-domain (11, 12, 2327). N-TIMPs have an oligonucleotide/oligosaccharide-binding fold, a 5-stranded β-barrel structure with two small helices. The core of the MMP-binding site of TIMP is a surface ridge formed by the N-terminal five residues, 1C(T/S)C(V/A/S)P5 and the connector between β-strands C and D, residues that are linked by the Cys-1 to Cys-70 disulfide bond (7, 2325). As shown in Fig. 1, this region is oriented in the active site of the MMP so that the α-amino group and carbonyl oxygen of Cys-1 coordinate the catalytic zinc (11, 12, 2328). When the α-amino group is chemically modified or extended by an N-terminal alanine, the MMP inhibitory activity is essentially eliminated (17, 2930). The side chain of TIMP residue 2 (Ser or Thr) sits over the mouth of the S1′ subsite (the “specificity pocket”) of the MMP active site (11, 12, 2328), and eliminating the side chain by substitution with glycine greatly weakens the affinity for most, but not all, MMPs (13, 14). Residues 3–5 interact with the S2′ and S3′ subsites, and the C-D β-strand connector interacts with the MMP S2 and S3 subsites (Fig. 1). Loops between β-strands A and B and strands E and F and the C-terminal end of β-strand D make variable contributions in different TIMP·MMP complexes (11, 12, 2328). The A-B loop of NT2 includes seven more residues than that of NT1 and has multiple interactions with the MMP in its complexes with MMP14 (MT1·MMP), MMP13c, and MMP10c (24, 25, 27, 28) that are absent from complexes of TIMP-1 with MMP1c, MMP3c, or MMP10c (11, 23, 27).

FIGURE 1.

FIGURE 1.

A schematic view of the interactions between the core of the NT1 interaction ridge and the active site and substrate binding subsites of MMP1. The N-terminal five residues of TIMP-1 are colored cyan, residues 67–70 and 98–99 are gray, the Cys-1–Cys-70 and Cys-3–Cys-99 disulfide bonds are yellow, oxygen atoms are red, and nitrogen atoms are blue. Residues 214–221 of MMP1 are represented by the blue ribbon, and the catalytic Zn2+ is represented by a purple sphere.

TIMP-3 inhibits more metalloproteinases than the other TIMPs, including disintegrin-metalloproteinases such as ADAM-17 (a disintegrin and metalloproteinase-17), ADAMTS-4 (a disintegrin and metalloproteinase with thrombospondin domains), and ADAMTS-5 (7), and these interactions are affected differently from those with MMPs by mutations in its reactive site (17, 31). The specificity of N-TIMPs for MMPs can also be modified by protein engineering to produce variants with enhanced or reduced selectivity for MMPs (1318) and have potential applications treating diseases linked to excess MMP activities. Previously, we identified NT2 variants that inhibit MMP1 preferentially over MMP3 using phage display in conjunction with positive and negative selection with MMP1c and MMP3c, respectively (18). MMP1 (fibroblast collagenase) and MMP3 (stromelysin 1) have been validated as a target and anti-target for cancer therapy mediated by MMP inhibition (32). The most selective mutant, S2D/S4A, has a nanomolar Ki for MMP-1 but does not inhibit MMP3c or MMP14c and weakly inhibits MMP2c, MMP7c, MMP8c, and MMP13c (18). Although other TIMPs are high affinity inhibitors of all MMPs, TIMP-1 is a poor inhibitor of the membrane-type MMPs (MMP14, MMP16, and MMP24) and MMP19. The T98L substitution in NT1 has been shown to enhance its affinity for MMP14 (2–3-fold), and additional substitutions, V4A and P6V, increase the affinity, reflected in a 10-fold improvement of the Ki in the triple mutant (15, 14). A crystallographic structure of the complex of the triple mutant of NT1 with MMP14 has been determined (26).

Here, we have investigated how the mutations in NT1 and NT2 that modify MMP selectivity affect the thermodynamics of N-TIMP·MMP interactions, focusing on the T98L mutation in NT1 with MMP3c and MMP14c and the S2D/S4A double substitution in NT2 (with MMP1c). The thermodynamic profiles of the interactions of these MMPs with WT NT1 and NT2 (6) show that all are driven by entropy increases. Entropy-driven binding has been observed in other proteins that bind to multiple targets, including thioredoxin and protein kinase A (33, 34). The solvent entropy change for the interaction, ΔSsolv, can be estimated from the heat capacity change, ΔCp (35). Using ΔSsolv and ΔSint, the conformation entropy change for the interaction (ΔSconf) can be determined to provide information about differences in conformational dynamics between the free and bound forms of the two proteins. The results highlight the contributions of changes in conformational dynamics and solvent entropy (the hydrophobic effect) to differences in binding to different MMPs.

Experimental Procedures

Construction, Expression, Purification, and Folding of NT2 Variants, MMP1c, and MMP3c

Reagents and cells were from the same sources as in previous studies (5, 6, 2427, 29, 31). The NT1 and NT2 mutants were generated using QuikChange II Site-Directed Mutagenesis kits (Agilent Technologies). Primers were designed using web-based primer design software program (Agilent Technologies). PCR reactions were carried out at 95 °C for 30 s, 55 °C for 1 min, and 68 °C for 5 min for 30 cycles after a 3-min hot start at 95 °C. The PCR products were cloned back into pET-42b vector (Novagen) for expression.

NT2 variants were extracted from inclusion bodies and folded as described previously (6). The native proteins were purified by ion exchange followed by gel filtration with columns (2.5 × 35 cm) of Superdex-75 (Amersham Bioscience), equilibrated, and eluted with 20 mm HEPES buffer, pH 7.4, containing 250 mm NaCl and 20 mm CaCl2. The eluate was collected in 6-ml fractions at a flow rate of 0.5 ml/min. Fractions containing folded N-TIMPs were identified by polyacrylamide gel electrophoresis, pooled, and concentrated using Centriplus YM-3 centrifugal filter devices (Millipore). MMP1c and MMP3c were expressed as inclusion bodies in Escherichia coli BL21-CodonPlus® Competent Cells, folded and purified as in previous studies (6).

Fluorescence Assays for N-TIMP Activity

The inhibition of MMPs by NT2 and NT1 variants was measured by assaying MMP activities for hydrolysis of fluorogenic substrates as described previously (6, 13, 14). Assays were conducted in HEPES buffer (20 mm), pH 7.4, containing 250 mm NaCl, 10 mm CaCl2, and 50 μm ZnCl2, which was used also for the dilution of MMP and TIMP samples. The Kiapp of N-TIMP variants for MMP-1c and MMP-3c at 25 °C (298 K) were determined as described previously (14, 16).

Isothermal Titration Calorimetry of the Interactions of NT1 and NT2 Variants with MMPs

Protein solutions were dialyzed extensively against various buffers at 20 mm concentrations containing 250 mm NaCl, 10 mm CaCl2, and 50 μm ZnCl2 at pH 7.4 and degassed before use. N-TIMPs (12–30 μm) were titrated with the MMP (120–300 μm) at different temperatures using a MicroCal VP-ITC microcalorimeter as described previously (5, 6). The instrument was programmed to carry out 14 injections of 20 μl each over 40 s, spaced at 300-s intervals (see Figs. 2 and 3). The stirring speed was 300 rpm. The data were analyzed by the software package Origin 5.0 from Microcal Inc., which was used to calculate the enthalpy changes (ΔH) and stoichiometry (N) using a single-site binding model. As in previous studies, only ∼44% of the recombinant NT1 and NT2 variants were active (5, 6) reflecting inactivation by N-acetylation the N terminus by the bacterium (30). The heat capacity change (ΔCp), intrinsic enthalpy change (ΔHint), and ionization change (NH+) for each interaction were calculated as described below.

FIGURE 2.

FIGURE 2.

Titration of different N-TIMPs by MMP14c in HEPES buffer, pH 7.4, at 310 K. A, aliquots (20 μl) of MMP14c (120 μm) were injected into NT1 (23 μm). B, aliquots (20 μl) of MMP14c (120 μm) were injected into NT1 T98L mutant (22 μm). C, aliquots (20 μl) of MMP14c (148 μm) were injected into NT2 (16 μm). The heats of binding were measured as described under “Experimental Procedures.”

FIGURE 3.

FIGURE 3.

Plot of observed enthalpies of binding against buffer ionization enthalpies for the interactions of MMP1c and MMP14c with different N-TIMPs. A, titrations of NT2 (circles), NT1 (squares), and NT1 T98L mutant (triangles) with MMP-4c. B, titrations of NT2 S2D/S4A mutant (circles) and S2D mutant (squares) with MMP1c. Data for the titration of WT NT2 with MMP1c, taken from Wu et al. (6), are represented by the line with no data points. Titrations were generally not repeated because of the large amounts of purified proteins required for each experiment (∼1500 μg of MMP and 700 μg of N-TIMP). However, data from four titrations, conducted in different buffers, were analyzed by linear regression analysis to determine ΔHinto and NH+. Also data from four titrations at different temperatures were similarly analyzed to determine ΔCp for each MMP/inhibitor pair so that the two key parameters, ΔHinto and ΔCp, for each NTIMP·MMP interaction were derived from four titrations.

Correlation of Thermodynamics with Structure

As discussed previously (5, 6) the ΔCpo value for a protein-protein interaction is generally considered to be related to changes in nonpolar and polar-accessible surface area, ΔASAnp and ΔASApol (Equations 1 and 2) on complex formation (where surface burial has a negative sign),

graphic file with name zbc02116-4444-m01.jpg

The parameterizations of the coefficients for changes in nonpolar and polar surface used here were a = 0.28 ± 0.12 cal mol−1 K−1−2 and b = −0.09 ± 0.30 cal mol−1 K−1−2 (36). The enthalpy of binding (ΔHo) at 60° (35) was calculated using the relationship,

graphic file with name zbc02116-4444-m02.jpg

where c is −7.27cal mol−1−2, and - is 29.16 cal mol−1−2 (4). ΔHo at 25 °C is then calculated using the calculated value for ΔCp. Polar and apolar surface areas in the interfaces of N-TIMP·MMP complexes were measured from atomic coordinates using InterProSurf (37).

Results and Discussion

Effect of the T98L Mutation in NT1 on Its Interactions with MMP3 and MMP14

For the interactions of NT1, NT2, and the T98L mutant of NT1 with MMP14, enthalpies of binding (ΔHobs) were determined by isothermal titrations at 291 K (Fig. 2) in buffers with different enthalpies of ionization (Pipes, Hepes, Mops, Bes, and Aces). These values were analyzed by linear regression analysis of the plot of ΔHobs against ΔHion using the relationship

graphic file with name zbc02116-4444-m03.jpg

where ΔHion is the enthalpy of ionization of the buffer, NH+ is the number of protons taken up (positive values) or released to the buffer during the protein-protein interaction, and ΔHinto is the enthalpy change independent of buffer (5, 6). These analyses (Fig. 3A) indicate that there is negligible release of protons for the interactions of WT NT1 and NT2 with MMP14 (NH+ of +0.06 ± 0.00 and +0.08 ± 0.01, respectively), whereas the interaction with the NT1 T98L mutant releases one proton (NH+ of 1.02 ± 0.10). The interactions all have highly unfavorable ionization-independent enthalpy changes, ranging from 9 kcal/mol for NT1 to 14.7 kcal/mol for the NT1 T98L mutant (Table 1). Their free energies of binding (ΔG) were calculated from the Ki values determined by inhibition kinetics at 298 using ΔGo = RTln(1/Ki) (see Table 3). ΔHobs values from titrations in HEPES buffer at 291, 298, 303, and 310 K (Table 2) were used to determine ΔCpo values using the relationship,

graphic file with name zbc02116-4444-m04.jpg
TABLE 1.

Enthalpies of binding for the interactions of WT NT1, NT1 T98L, and NT-2 with MMP3c and NT1 T98L with MMP14c in buffers of different enthalpies of ionization

Titrations were carried out at 291 K. ΔHint, the intrinsic enthalpy change NH+ and ionization change were calculated by linear regression.

Buffer ΔHion ΔHobs
MMP3c·NT1 T98L MMP14c·NT2 WT MMP14c·NT1 WT MMP14c·NT1 T98L
kcal/mol
Pipes 2.71 4.65 ± 0.08 10.17 ± 0.10a 11.77 ± 0.37 14.83 ± 0.37
Hepes 4.94 2.63 ± 0.03 10.38 ± 0.09 13.82 ± 0.24 14.96 ± 0.25
Mops 5.20 2.48 ± 0.03
Bes 6.01 10.44 ± 0.11 14.61 ± 0.62 15.04 ± 0.25
Aces 7.55 1.51 ± 0.20 10.51 ± 0.06 16.81 ± 0.30 15.10 ± 0.20
NH+ −0.65 ± 0.11 0.08 ± 0.01 1.02 ± 0.10 0.06 ± 0.003
ΔHint° 6.11 ± 0.62 10.00 ± 0.05 8.87 ± 0.54 14.68 ± 0.02

a Standard errors for fitting data to a single site model using Origen software.

TABLE 3.

Thermodynamic profiles for interactions of NT1 variants and NT2 with MMP3c and MMP14c, at 298 K

Parameter MMP3c·NT2 WTa MMP3c·NT1 WTa MMP3c·NT1 T98L MMP14c·NT2 WT NT1 WT·MMP14c NT1T98L/ MMP14c
ΔHint (kcal/mol) 5.5 ± 0.2b 6.5 ± 0.3b 6.9 ± 0.6 8.0 ± 0.1 6.1 ± 0.5 10.7 ± 0.02
Ka 2.3 × 108 3.5 × 108 1.6 × 109 1.8 × 1010 3.1 × 106 6.7 × 106
ΔG1M (kcal/mol) −11.1 ± 0.2 −11.7 ± 0.1 −12.3 −13.7 −8.6 −9.1
TΔSint (kcal/mol) 16.6 ± 0.2 18.2 ± 0.4 19.2 ± 0.6 21.7 ± 0.05 14.7 ± 0.5 19.8 ± 0.0
ΔCpo (cal/mol/K) −47.2 ± 3.7 −50 ± 6 (−51)c +130 ± 7 (+139)c −288 ± 17 (−289)c −398 ± 11 (−409)c −575 ± 29 (−576)c
TΔSsolv (kcal/mol) 3.6 3.8 −10.6 23.5 ± 1.4 32.4 ± 0.9 46.8 ± 2.4
Estimated TΔSconf d 16.0 17.4 32.8 1.2 ± 1.4 −14.7 −24.0

a Data were from Wu et al. (6).

b Calculated from values at 291K (6).

c Corrected for buffer ionization dependence (6).

d Estimated by assuming a cost of 3 kcal/mol for decrease in entropy of translational and rotational degrees of freedom arising from the protein-protein interaction.

TABLE 2.

Enthalpies of binding for the interactions of NT1 T98L with MMP3c and NT1, NT1 T98L, and N-cT2 with MMP14c at different temperatures

ΔCp was determined by linear regression of ΔHobs, and K and was used to estimate ΔSsolv at 298K as described under “Results and Discussion.”

Temperature (K) ΔHobs
MMP3c·NT1 T98L MMP14c ·NT2 WT NT1 WT/· MMP14c NT1 T98L·MMP14c
kcal/mol
291 2.63 ± 0.03 10.38 ± 0.09 13.82 ± 0.24 14.96 ± 0.25
298 3.38 ± 0.04 8.69 ± 0.04 10.77 ± 0.09 11.28 ± 0.18
303 4.09 ± 0.08 7.27 ± 0.03 9.04 ± 0.09 8.70 ± 0.10
310 5.09 ± 0.08 4.90 ± 0.02 6.18 ± 0.07 3.96 ± 0.14
ΔCpo (cal/mol/K) 130 ± 7 −288 ± 17 (−289)a −398 ± 11 (−409)a −575 ± 29 (−576)a
ΔSsolv (cal/mol) −33 74 102 148

a Corrected for buffer ionization dependence (6).

The value of ΔCpo is provided by the slope of a plot of ΔHobs against temperature (Fig. 4A). The ΔGo values were determined at 298 K, but the ΔHint values were measured at 291 K. Therefore, to obtain a set of parameters at the same temperature, ΔCp for the interaction was used to calculate the ΔHint at 298 K from the value at 291 K (Table 3). TΔSint values for the mutants were calculated from ΔG and ΔHint at 298 K (Table 3). These results indicate that the weaker binding of NT1 to MMP14 relative to NT2 reflects a 7 kcal/mol lower TΔSint of binding. The increased affinity arising from the T98L mutation results from an increase in TΔSint that is partly offset by an (unfavorable) increase in the ΔHint of binding (Table 3).

FIGURE 4.

FIGURE 4.

Plot of observed interaction enthalpies of binding against temperature (K) for different N-TIMP interactions with MMPs. A, data for the titration of NT2, NT1, and the T98L mutant of NT1 with MMP14c are displayed as circles, squares, and triangles, respectively. B, results for the titrations the NT2 S2D/S4A mutant (circles), NT2 S2D mutant (squares), and NT2 (no symbols and taken from Wu et al. (6)) with MMP1c. Results for the titration of MMP3c by the NT1 T98L are represented by filled circles. The lines were generated by linear regression analysis.

Further analysis clarified the source of the entropy changes that affect the binding of these N-TIMP variants with MMP14. The overall entropy change, ΔSint, includes contributions from the interacting proteins (ΔSprotein) and solvent (ΔSsolv), of which the latter can be estimated from the ΔCpo for the interaction using the relationship

graphic file with name zbc02116-4444-m05.jpg

where Ts* is the reference temperature (385 K) at which the hydrophobic contribution to ΔS is zero (4, 35). ΔSprotein includes negative, unfavorable effects arising from the loss of translational and rotational freedom T(ΔStrans + ΔSrot) on complex formation, which has been estimated to be ∼3 (±2.4) kcal/mol for the NT1/MMP3cd interaction (5) and is expected to be similar for other N-TIMP·MMP interactions (6). It also includes the change in conformational entropy, ΔSconf, that encompasses an unfavorable loss of entropy resulting from increased rigidity of the interaction sites of the two proteins in their complex. However, it may also contain favorable entropy increases resulting from increased dynamics in regions distant from the interaction sites in the bound protein populations (5, 6). ΔCpo values for the interactions with MMP-14 are large and negative and lead to estimates of 32–47 kcal/mol for increases in TΔSsolv for the interactions with NT1, NT2, and NT1 T98L (Table 3). Using these values we estimate that the binding of either NT1 or NT1 T98L to MMP-14c results in energetic costs reflecting conformational entropy changes (TΔSconf) of −14.7 and −24.0 kcal/mol for NT1 and NT1 T98L, respectively (Table 3). The interaction with NT2 is associated with an insignificant change in TΔSconf (1.2 ± 1.4 kcal/mol). These results show that the weak binding of NT1 to MMP-14c arises from a large conformational entropy penalty. Surprisingly, the T98L mutation increases this penalty, but this increase is exceeded by the enhanced hydrophobic effect (TΔSsolv), consistent with the hydrophilic to hydrophobic T98L substitution in the NT1 reactive site.

The T98L mutation in NT1 also increases the affinity for MMP-3 by a factor of 4 (15), but the thermodynamic origins of this are different. A more unfavorable ΔHint (8.2 versus 6.0 kcal/mol) is exceeded by an increase in TΔSint (Table 3). Unexpectedly, ΔHobs for binding to MMP3c increases with temperature (Fig. 4B), indicating a positive ΔCp value of 124 cal/mol/K. From this we estimate a value of −10.6 kcal/mol for TΔSsolv, an unfavorable contribution to ΔG that is compensated by an estimated increase in TΔSconf of 33 kcal/mol. In contrast, the interaction of MMP3c with WT NT1, characterized previously, has a ΔCp of −50, a favorable TΔSsolv of 3.8 kcal/mol, and a smaller TΔSconf of 17.4 kcal/mol (5).

Interaction of the NT2 S2D and S2DS4A Mutants with MMP-1c

The interactions of these two mutants were characterized with only MMP1c because they do not inhibit MMP3c or MMP14c (18). Linear regression analysis of ΔHobs values for isothermal titrations at 291 K in buffers with different enthalpies of ionization (Table 4; Figs. 3 and 5) indicates fractional release and uptake of protons, (NH+) of −0.47 and +0.46, respectively, for the S2D and S2D/S4A mutants compared with a negligible uptake of +0.14 previously determined for WT NT2. This analysis also shows that the ionization-independent ΔHint for the single site mutant, S2D, is 9.5 kcal/mol, much more unfavorable than that for the double mutant, S2D/S4A (1.5 kcal/mol; Table 4), contributing to the weaker binding of S2D shown by ΔG values (calculated from the Ki values) of −10.2 kcal/mol for S2D and −14.5 kcal/mol for S2D/S4A (see Table 6). As previously discussed, ΔHobs values from titrations at different temperatures (Table 5, Fig. 4B) were used to determine ΔCpo. This in turn was used to calculate the ΔHint at 298 K and to estimate ΔSsolv. Values of TΔSint for both mutants, calculated from ΔG and ΔHint values at 298 K, show that the entropy contribution for S2D is ∼2 kcal/mol greater than for the S2D/S4A mutant. The thermodynamic parameters of S2D/S4A are similar to those for WT NT2, but the weaker binding of the S2D mutant reflects a larger enthalpy penalty that is only partly compensated by the increase in TΔSint (Table 6).

TABLE 4.

Enthalpies of binding for the interactions of variants of NT2 with MMP1c in buffers of different enthalpies of ionization at 291 K

Titrations were carried out at 291 K. ΔHint, the intrinsic enthalpy change NH+ and ionization change were calculated by linear regression analysis.

Buffer ΔHion ΔHobs
NT2a NT2 S2D/S4A NT2 S2D
kcal/mol
Pipes 2.71 2.83 ± 0.02 8.26 ± 0.11
Hepes 4.94 3.82 ± 0.03 7.36 ± 0.10
Bes 6.01 4.21 ± 0.04 6.53 ± 0.08
Aces 7.55 5.06 ± 0.04 6.09 ± 0.12
NH+ 0.14 0.46 ± 0.03b −0.47 ± 0.05
ΔHint° (kcal/mol) 2.54 1.54 ± 0.17 9.53 ± 0.28

a Data are from Ref.6.

b S.E. for fitting data to Equation 3.

FIGURE 5.

FIGURE 5.

Isothermal calorimetric titration of NT2 S2D/S4A mutant with MMP1c in different buffers at 291K. Left, aliquots (20 μl) of MMP1c (150 μm) were injected into NT2 (30 μm) in ACES buffer, pH 7.4. Right, aliquots (20 μl) of MMP1c (150 μm) were injected into NT2 (30 μm) in HEPES buffer, pH 7.4. The heats of binding were measured as described under “Experimental Procedures.”

TABLE 6.

Thermodynamic profiles for interactions of NT2 variants and NT1 with MMP1c, at 298 K

Parameter NT2 WTa NT2 S2D/S4A NT2 S2D NT1 WTa
ΔHint (kcal/mol) 0.58 0.79 ± 0.2b 8.3 ± 0.3b −0.86
Ka 2.8 × 1010 4 × 1010 2.9 × 107 4.3 × 109
ΔG1M (kcal/mol) −14.2 −14.5 −10.2 −13.4
TΔSint (kcal/mol) 14.8 16.2 ± 0.2 18.5 ± 0.3 12.5
ΔCpo (cal/mol/ K) −278 ± 10 −102 ± 25 (−107)c −165 ± 25 (−160)b −189
Estimated TΔSsolvc (kcal/mol) 21.2 8.2 ± 1.9 12.2 ± 1.8 14.8
Estimated TΔSconfd −3.4 11.0 9.3 0.7

a Values were taken from Wu et al. (6).

b Values calculated at 298 K using ΔCp.

c Corrected for buffer ionization dependence (6).

d Estimated by assuming a cost of 3 kcal/mol for decrease in entropy of translational and rotational degrees of freedom arising from the protein-protein interaction.

TABLE 5.

Enthalpies of interaction of NT2 variants with MMP1c at different temperatures

ΔCp was determined by as described in the legend to Table 3.

Temperature (K) ΔHobs
NT2a NT2 S2D/S4A NT2 S2D
kcal/mol kcal/mol
291 3.82 ± 0.03 7.36 ± 0.10
298 2.94 ± 0.04 5.57 ± 0.09
303 2.40 ± 0.04 5.01 ± 0.09
310 1.88 ± 0.03 4.15 ± 0.16
ΔCpo (cal/mol/K) −278 −102 ± 25 (107)b −165 ± 25
ΔSsolv (cal/mol) 71 27 42

a Data were from Ref.6.

b Corrected for buffer ionization dependence (6).

The source of the entropy driving the interactions WT NT2 and the mutants with MMP1c was determined as described in the previous section using solvent entropy changes (ΔSsolv) estimated from the heat capacity changes for the interactions. The values of TΔSconf at 298 K indicate that conformational entropy increases contribute −9.3 and −11.0 kcal/mol to ΔG for interactions of S2D and S2DS4A with MMP1c, contrasting with WT NT2, which has an unfavorable TΔSconf of −3.4 kcal/mol, presumably derived from increased rigidity in the interaction interface. The thermodynamic parameters previously determined for the interaction of NT1 with MMP1c are also given in Table 6.

Relationship of Thermodynamic Profiles to Structures

Like previously studied N-TIMP·MMP interactions (5, 6) the present interactions have minimal to strongly positive ΔHint values and are driven by large entropy increases. This helps to explain how TIMPs inhibit numerous matrixins whose active sites are adapted to binding diverse biological targets.

Crystallographic structures relevant to the interactions with MMP14 are the TIMP2·MMP14c complex, pdb 1BUV, and the complex of the V4A/P6V/T98L of NT1 with MMP14c, pdb 3MA2 (24, 26). A model of the T98L NT1 mutant with MMP3c was generated from pdb 1UEA (23) by truncating residues 126–180, changing Thr-98 to Leu followed by energy minimization. The V4A, P6V, and T98L mutations in NT1 individually lower the Ki for MMP14c 2–3 fold, suggesting that they contribute additively to the ∼10-fold increase in affinity of the triple mutant (14) without a major conformational change, supporting the use of the 3MA2 structure (26) as a template. The areas of apolar and polar surface in the interfaces of the models were analyzed to predict the values of ΔCp and ΔH (at 25 °C) for the interaction (27, 28) for comparison with those determined experimentally (Table 7). These values show reasonable agreement for the MMP14c complexes, but the experimentally determined ΔCp values for the MMP3c complexes stand out as being far less negative than those calculated from the structures.

TABLE 7.

Comparison of the ΔCp and ΔHint (298 K) values calculated from the polar and apolar interface surface areas of NTIMP·MMP complexes, as described under “Experimental Procedures,” with those determined experimentally

Complex Parameters
Buried surface areas
ΔCp
ΔH
Polar Apolar Calculated Experimental Calculated Experimental
Å2 cal/mol/K kcal/mol
MMP1·NT1 564 971 −221 −194 −0.9 +0.5
MMP3·NT1 514 1525 −381 −51 +9.4 +6.5
MMP3·NTI T98L 558 1532 −388 +139 +10.1 +6.9
MMP14·NT1 512 1205 −292 −398 +4.1 +6.1
MMP14·NT1 T98L 505 1218 −296 −575 +4.5 +10.7
MMP14·NT2 659 1661 −391 −289 +3.9 +8.0

We previously developed models of complexes of NT2 and its S2DS4A mutant with MMP1c (18). These suggested that the double mutation results in more extensive contacts of NT2 Asp-2 with the S1′ pocket of MMP1c relative to Ser2 in WT NT2 (18). The selective inhibition of MMP1c by the S2D/S4A and S2D mutants appears to result from the unique presence of R114 in MMP1c (see Fig. 1), which is replaced by uncharged residues in other human MMPs except for MMP21 (Lys) and MMP23 (His). In contrast, the model of the complex of the NT2 S2D/S4A mutant with MMP-3c shows increased separation in the S1′ pocket, suggesting that charge repulsion between Asp-2 of the inhibitor and the catalytic site Glu-202 of MMP-3c is responsible for the extremely weak binding (18). In the complex of the S2D/S4A mutant with MMP1c, favorable electrostatic interactions between Asp-2 of the NT2 mutant and Arg214 in MMP-1 allows high affinity binding. The greater polarity of Asp as compared with S2 of WT NT2 also results in less negative ΔCp values for the interactions of this mutant with MMP1. This leads to lower (estimated) TΔSsolv values and increased values for TΔSconf implying that the reduced hydrophobic effect is compensated by enhanced conformational dynamics. We propose that structural adjustments in the interaction interface resulting from the insertion of the negatively charged Asp-2 are transmitted through the protein structures to more distant sites, reducing stability and increasing conformational dynamics (5).

Conformational Dynamics and Selectivity in MMP Inhibition

The disagreement between the experimental and calculated values for ΔCp for the NT1/MMP3c interaction was previously attributed to large conformational differences the structures of the both proteins in their free and bound states (19, 22, 23). As a result, the areas of polar and apolar surface quantified from the interaction interface in the complex differ from those buried during the interaction process (5, 6). In contrast, in the NT1·MMP1c complex, the structure of MMP1c shows only minor differences when compared with the structure of the MMP1c (11). These analyses are misleading because they treat the crystallographic structures as static models, whereas they are the average structures of assemblies of conformers (38, 39). The process of complex formation results in the selection of populations of conformers that differ from those in the free state; the selected populations differ in conformational dynamics, solvent interactions, and structure. Fig. 6 compares the structures of the NT1 components from the crystallographic structures of various MMP complexes and with two chains from the solution NMR structure of free NT1 (11, 22, 23, 26, 27) showing large differences, particularly in the loops between β-strands A and B and B and C where missing electron density indicates local unfolding. This may expose apolar groups, making a positive contribution to ΔCp. The core of the reactive site, including the N-terminal five residues is similar in the different complexes but differs from the NMR structure of the free protein (Fig. 6).

FIGURE 6.

FIGURE 6.

Superimposed ribbon structures of NT1 extracted from crystallographic structures of different complexes and solution NMR structures of the free protein. The structures from the complexes are colored as follows: pink, MMP1c (PDB code 2jot); cyan, MMP3c (PDB code 1uea); blue, MMP14c (PDB code 3ma2); purple, MMP10c (PDB code 3v96). The free NT1 structures were chains 1, 20 and 29 from PDB code 1d2b and are colored light gray. The structures were superimposed and displayed using CHIMERA (43).

Eftink et al. (40) have shown that if an interacting protein has two conformational states that both interact with a binding partner, indicating non-mandatory coupling between binding and conformational change, ΔCp can have positive or negative values depending on parameters relating to the conformational transition (40). This may explain the “anomalous” ΔCp values for N-TIMP interactions with MMP3c, including the positive ΔCp for the interaction of MMP3c with the NT1 T98L mutant.

Previously reported dynamics simulations with NT1 and its T98L mutant suggested that the mutation reduced the flexibility of its interface for MMP-14, including the highly flexible N-terminal region, leading to a reduced ΔSconf penalty for binding (26). Our experimental data indicate that WT NT1 binding carries a large entropic penalty that reduces TΔSconf by 12 kcal/mol. The NT1 T98L mutation increases this penalty by 8 kcal/mol, and the increased Ka comes from an increase in solvent entropy (TΔSsolv), estimated from the more negative ΔCp. The reduced flexibility in free NT1 indicated by dynamics simulations appears to be negated by a greater loss of conformational entropy in the NT1·MMP14 complex that may arise from additional interactions between Leu-98 of the NTIMP and the MMPs that are not present in the complex with WT NT1. Such interactions lead to a more negative ΔCp, reflecting an increase in the hydrophobic effect.

Fig. 7A shows the strong negative correlation between ΔSsolv (derived from ΔCp) and ΔSconf for all currently characterized N-TIMP·MMP interactions (adjusted R2 = 0.975). The combined data for the interactions of three MMPs with WT and mutant forms of NT1 and NT2 do not include all possible mutant N-TIMP·MMP combinations. However, the data for the interactions of the WT N-TIMPs with the three MMPs show a similar correlation between ΔSconf andΔSsolv. One explanation of ΔSsolvSconf compensation is that ΔSconf is calculated from ΔSsolv and ΔSint; TΔSint for the 10 currently characterized interactions has a low variance (57.7 ± 9.4 cal/mol) so that the apparent compensation between ΔSsolv and ΔSconf is not unexpected. In Fig. 7A, the data points for interactions with each MMP are grouped together, and the graphical summary of thermodynamic parameters in Fig. 7B shows that interactions involving MMP3 have the highest ΔSconf and lowest ΔSsolv values, whereas those involving MMP14 have the lowest (mostly negative) ΔSconf and highest ΔSsolv. Interactions between NTIMPs and MMP1 have intermediate values of ΔSconf and ΔSsolv and the least unfavorable ΔHint. This is consistent with our previous observation, based on less data, that the character of the MMP has a major influence on the proportions of ΔSconf and ΔSsolv (6), apparently reflecting intrinsic differences in structure and flexibility between the MMPs. Interactions with MMP14 bury more nonpolar surface than those with MMP1, resulting in a more positive ΔSsolv. MMP3c has a higher conformational flexibility than MMP1c and MMP14c (41), reflected in conformational changes in complexes with inhibitors and TIMP-1 (23, 42). Possibly, these differences between MMPs might be exploited in designing more selective TIMP mutants or other inhibitors, but this could be limited by the observed compensation between ΔSconf and ΔSsolv.

FIGURE 7.

FIGURE 7.

Contributions of ΔHint, ΔSsolv, and ΔSconf to N-TIMP·MMP interactions. A, bar chart of values for different interactions. Black, ΔHint; white, ΔSsolv; gray, ΔSconf. B, linear inverse relationship between ΔSsolv and ΔSconf for all characterized interactions. Data for interactions with MMP1c are denoted by circles, MMP3c by squares, and MMP14c by triangles; WT N-TIMPs are represented by open symbols and engineered N-TIMPs by filled symbols. The units of ΔS are cal/mol. The line, generated by regression analysis, represents the relationship: ΔSconf = 67.6 − ΔSsolv.

Conclusions

Specific protein-protein interactions are pivotal for the control of numerous biological processes including those mediated by regulated proteolysis. Although the four human TIMPs are high affinity metalloproteinase inhibitors, they show little selectivity in their interactions with the 23 human MMPs. This is consistent with entropy-driven binding, which encompasses free energy contributions from the relatively nonspecific hydrophobic effect (solvent entropy changes) and/or changes in conformational entropy. The mutually compensating ΔSsolv and ΔSconf, estimated using the key thermodynamic parameter, ΔCp (44), are important in modulating NTIMP·MMP interactions. The character of the MMP strongly affects their proportions, and the magnitude of ΔHint (Fig. 7B). This is illustrated by the T98L mutation in NT1, which enhances binding to MMP-3 by increasing ΔSconf and to MMP-14 by enhancing ΔSsolv. The highly conserved N-terminal regions of TIMPs have a high level of flexibility in the free protein (22), a common feature of multi-specific protein interaction sites (45). Mutations in the N terminus have major effects on avidity and selectivity for MMPs. In previous studies, NTIMP mutants identified as having the greatest selectivity against MMP-1 are the NT1 T2R (14) and T2R/V4I mutants (16), whereas that with the greatest selectivity for MMP-1 is the NT2 S2D/S4A mutant (18). These mutations alter electrostatic interactions between the NTIMP and the unique Arg-214 in the MMP S1′ site, suggesting that the N-terminal region of TIMPs and unique active site features of MMPs are both keys to developing selective inhibitors.

Author Contributions

K. B. devised and supervised the study and wrote the paper. H. Z. and Y. W. designed and performed the experiments and analyzed the data.

*

This work was supported by National Institutes of Health Grant RO1 AR40994. The authors declare that they have no conflicts of interest with the contents of this article. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

3
The abbreviations used are:
NT1
N-terminal domain of TIMP-1
NT2
N-terminal domain of TIMP-2
MMP
matrix metalloproteinase
MMP1c
collagenase 1 catalytic domain
MMP3c
stromelysin 1 catalytic domain
MMP14c
membrane-type metalloproteinase-1 catalytic domain
TIMP
tissue inhibitor of metalloproteinase
N-TIMP
N-terminal inhibitory domain of TIMP
Kiapp
apparent inhibition constant
Bes
2-[bis(2-hydroxyethyl)amino]ethanesulfonic acid
Aces
2-[(2-amino-2-oxoethyl)amino]ethanesulfonic acid.

References

  • 1. Velázquez Campoy A., and Freire E. (2005) ITC in the post-genomic era….? Priceless. Biophys. Chem. 115, 115–124 [DOI] [PubMed] [Google Scholar]
  • 2. Murphy K. P., Freire E. (1992) Thermodynamics of structural stability and cooperative folding behavior in proteins. Adv. Protein Chem. 43, 313–361 [DOI] [PubMed] [Google Scholar]
  • 3. Spolar R. S., Livingstone J. R., and Record M. T. Jr. (1992) Use of liquid hydrocarbon and amide transfer data to estimate contributions to thermodynamic functions of protein folding from the removal of nonpolar and polar surface from water. Biochemistry 31, 3947–3955 [DOI] [PubMed] [Google Scholar]
  • 4. Baker B. M., and Murphy K. P. (1998) Prediction of binding energetics from structure using empirical parameterization. Methods Enzymol. 295, 294–315 [DOI] [PubMed] [Google Scholar]
  • 5. Arumugam S., Gao G., Patton B. L., Semenchenko V., Brew K., and Van Doren S. R. (2003) Increased backbone mobility in β-barrel enhances entropy gain driving binding of N-TIMP-1 to MMP-3. J. Mol. Biol. 327, 719–734 [DOI] [PubMed] [Google Scholar]
  • 6. Wu Y., Wei S., Van Doren S. R., and Brew K. (2011) Entropy increases from different sources support the high-affinity binding of the N-terminal inhibitory domains of tissue inhibitors of metalloproteinases to the catalytic domains of matrix metalloproteinases-1 and -3. J. Biol. Chem. 286, 16891–16899 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Brew K., and Nagase H. (2010) The tissue inhibitors of metalloproteinases (TIMPs): an ancient family with structural and functional diversity. Biochim. Biophys. Acta 1803, 55–71 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Murphy G., and Nagase H. (2008) Progress in matrix metalloproteinase research. Mol. Aspects Med. 29, 290–308 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Mott J. D., and Werb Z. (2004) Regulation of matrix biology by matrix metalloproteinases. Curr. Opin. Cell Biol. 16, 558–564 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Huang W., Suzuki K., Nagase H., Arumugam S., Van Doren S. R., and Brew K. (1996) Folding and characterization of the amino-terminal domain of human tissue inhibitor of metalloproteinases-1 (TIMP-1) expressed at high yield in E. coli. FEBS Lett. 384, 155–161 [DOI] [PubMed] [Google Scholar]
  • 11. Iyer S., Wei S., Brew K., and Acharya K. R. (2007) Crystal structure of the catalytic domain of matrix metalloproteinase-1 in complex with the inhibitory domain of tissue inhibitor of metalloproteinase-1. J. Biol. Chem. 282, 364–371 [DOI] [PubMed] [Google Scholar]
  • 12. Wisniewska M., Goettig P., Maskos K., Belouski E., Winters D., Hecht R., Black R., and Bode W. (2008) Structural determinants of the ADAM inhibition by TIMP-3: crystal structure of the TACE-N-TIMP-3 complex. J. Mol. Biol. 381, 1307–1319 [DOI] [PubMed] [Google Scholar]
  • 13. Meng Q., Malinovskii V., Huang W., Hu Y., Chung L., Nagase H., Bode W., Maskos K., and Brew K. (1999) Residue 2 of TIMP-1 is a major determinant of affinity and specificity for matrix metalloproteinases but effects of substitutions do not correlate with those of the corresponding P1′ residue of substrate. J. Biol. Chem. 274, 10184–10189 [DOI] [PubMed] [Google Scholar]
  • 14. Hamze A. B., Wei S., Bahudhanapati H., Kota S., Acharya K. R., and Brew K. (2007) Constraining specificity in the N-domain of tissue inhibitor of metalloproteinases-1: gelatinase-selective inhibitors. Protein Sci. 16, 1905–1913 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Lee M.-H., Rapti M., Knaüper V., and Murphy G. (2004) Threonine 98, the pivotal residue of tissue inhibitor of metalloproteinases (TIMP)-1 in metalloproteinase recognition. J. Biol. Chem. 279, 17562–17569 [DOI] [PubMed] [Google Scholar]
  • 16. Wei S., Chen Y., Chung L., Nagase H., and Brew K. (2003) Protein engineering of the tissue inhibitor of metalloproteinase 1 (TIMP-1) inhibitory domain. In search of selective matrix metalloproteinase inhibitors. J. Biol. Chem. 278, 9831–9834 [DOI] [PubMed] [Google Scholar]
  • 17. Wei S., Kashiwagi M., Kota S., Xie Z., Nagase H., and Brew K. (2005) Reactive site mutations in tissue inhibitor of metalloproteinase-3 disrupt inhibition of matrix metalloproteinases but not tumor necrosis factor-α-converting enzyme. J. Biol. Chem. 280, 32877–32882 [DOI] [PubMed] [Google Scholar]
  • 18. Bahudhanapati H., Zhang Y., Sidhu S. S., and Brew K. (2011) Phage display of tissue inhibitor of metalloproteinases-2 (TIMP-2): identification of selective inhibitors of collagenase-1 (metalloproteinase 1 (MMP-1)). J. Biol. Chem. 286, 31761–31770 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Arumugam S., Hemme C. L., Yoshida N., Suzuki K., Nagase H., Berjanskii M., Wu B., and Van Doren S. R. (1998) TIMP-1 contact sites and perturbations of stromelysin 1 mapped by NMR and a paramagnetic surface probe. Biochemistry 37, 9650–9657 [DOI] [PubMed] [Google Scholar]
  • 20. Williamson R. A., Carr M. D., Frenkiel T. A, Feeney J., and Freedman R. B. (1997) Mapping the binding site for matrix metalloproteinase on the N-terminal domain of the tissue inhibitor of metalloproteinases-2 by NMR chemical shift perturbation. Biochemistry 36, 13882–13889 [DOI] [PubMed] [Google Scholar]
  • 21. Williamson R. A., Muskett F. W., Howard M. J., Freedman R. B., and Carr M. D. (1999) The effect of matrix metalloproteinase complex formation on the conformational mobility of tissue inhibitor of metalloproteinases-2 (TIMP-2). J. Biol. Chem. 274, 37226–37232 [DOI] [PubMed] [Google Scholar]
  • 22. Wu B., Arumugam S., Gao G., Lee G. I., Semenchenko V., Huang W., Brew K., and Van Doren S. R. (2000) NMR structure of tissue inhibitor of metalloproteinases-1 implicates localized induced fit in recognition of matrix metalloproteinases. J. Mol. Biol. 295, 257–268 [DOI] [PubMed] [Google Scholar]
  • 23. Gomis-Ruth F.-X., Maskos K., Betz M., Bergner A., Huber R., Suzuki K., Yoshida N., Nagase H., Brew K., Bourenkov G. B., Bartunik H., and Bode W. (1997) Mechanism of the inhibition of the human matrix metalloproteinase stromelysin-1 by TIMP-1. Nature 389, 77–81 [DOI] [PubMed] [Google Scholar]
  • 24. Fernandez-Catalan C., Bode W., Huber R., Turk D., Calvete J. J., Lichte A., Tschesche H., and Maskos K. (1998) Crystal structure of the complex formed by the membrane type 1-matrix metalloproteinase with the tissue inhibitor of metalloproteinases-2, the soluble progelatinase A receptor. EMBO J. 17, 5238–5248 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Maskos K., Lang R., Tschesche H., and Bode W. (2007) Flexibility and variability of TIMP binding: x-ray structure of the complex between collagenase-3/MMP-13 and TIMP-2. J. Mol. Biol. 366, 1222–1231 [DOI] [PubMed] [Google Scholar]
  • 26. Grossman M., Tworowski D., Dym O., Lee M. H., Levy Y., Murphy G., and Sagi I. (2010) The intrinsic protein flexibility of endogenous protein inhibitor TIMP-1 controls its binding interface and affects its function. Biochemistry 49, 6184–6192 [DOI] [PubMed] [Google Scholar]
  • 27. Batra J., Robinson J., Soares A. S., Fields A. P., Radisky D. C., and Radisky E. S. (2012) Matrix metalloproteinase-10 (MMP-10) interaction with tissue inhibitors of metalloproteinase TIMP-1 and TIMP-2; binding studies and crystal structure. J. Biol. Chem. 287, 15935–15946 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Batra J., Soares A. S., Mehner C., and Radisky E. S. (2013) Matrix metalloproteinase-10/TIMP-2 structure and analyses define conserved core interactions and diverse exosite interactions in MMP/TIMP complexes. Plos ONE 8, e75836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Higashi S., and Miyazaki K. (1999) Reactive site-modified tissue inhibitor of metalloproteinases-2 inhibits the cell-mediated activation of progelatinase A. J. Biol. Chem. 274, 10497–10504 [DOI] [PubMed] [Google Scholar]
  • 30. Van Doren S. R., Wei S., Gao G., DaGue B. B., Palmier M. O., Bahudhanapati H., and Brew K. (2008) Inactivation of N-TIMP-1 by N-terminal acetylation when expressed in bacteria. Biopolymers 89, 960–968 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Lim N. H., Kashiwagi M., Visse R., Jones J., Enghild J. J., Brew K., and Nagase H. (2010) Reactive site mutants of N-TIMP-3 that selectively inhibit ADAMTS-4 and ADAMTS-5: biological and structural implications. Biochem. J. 431, 113–122 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Overall C. M., and Kleifeld O. (2006) Towards third generation matrix metalloproteinase inhibitors for cancer therapy. Br. J. Cancer 94, 941–946 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Palde P. B., and Carroll K. S. (2015) A universal entropy-driven mechanism for thioredoxin-target recognition. Proc. Natl. Acad. Sci. U.S.A. 112, 7960–7965 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Chang C. E., McLaughlin W. A., Baron R., Wang W., and McCammon J. A. (2008) Entropic contributions and the influence of the hydrophobic environment in promiscuous protein-protein association. Proc. Natl. Acad. Sci. U.S.A. 105, 7456–7461 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Baldwin R. L. (1986) Temperature dependence of the hydrophobic interaction in protein folding. Proc. Natl. Acad. Sci. U.S.A. 83, 8069–8072 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Spolar R. S., and Record M. T. Jr (1994) Coupling of local folding to site-specific binding of proteins to DNA. Science 263, 777–784 [DOI] [PubMed] [Google Scholar]
  • 37. Negi S. S., Schein C. H., Oezguen N., Power T. D., and Braun W. (2007) InterProSurf: a web server for predicting interacting sites on protein surfaces. Bioinformatics 23, 3397–3399 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Kumar S., Ma B., Tsai C. J., Sinha N., and Nussinov R. (2000) Folding and binding cascades: dynamic landscapes and population shifts. Protein Sci. 9, 10–19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Grünberg R., Nilges M., and Leckner J. (2006) Flexibility and conformational entropy in protein-protein binding. Structure 14, 683–693 [DOI] [PubMed] [Google Scholar]
  • 40. Eftink M. R., Anusiem A. C., and Biltonen R. L. (1983) Enthalpy-entropy compensation and heat capacity changes for protein-ligand interactions: general thermodynamic models and data for the binding of nucleotides to ribonuclease A. Biochemistry 22, 3884–3896 [DOI] [PubMed] [Google Scholar]
  • 41. Liang X., Arunima A., Zhao Y., Bhaskaran R., Shende A., Byrne T. S., Fleeks J., Palmier M. O., and Van Doren S. R. (2010) Apparent tradeoff of higher activity in MMP-12 for enhanced stability and flexibility in MMP-3. Biophys. J. 99, 273–283 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Chen L., Rydel T. J., Gu F., Dunaway C. M., Pikul S., Dunham K. M., and Barnett B. L. (1999) Crystal structure of the stromelysin catalytic domain at 2.0 Å resolution: inhibitor-induced conformational changes. J. Mol. Biol. 293, 545–557 [DOI] [PubMed] [Google Scholar]
  • 43. Pettersen E. F., Goddard T. D., Huang C. C., Couch G. S., Greenblatt D. M., Meng E. C., and Ferrin T. E. (2004) UCSF Chimera-a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 [DOI] [PubMed] [Google Scholar]
  • 44. Prabhu N. V., and Sharp K. A. (2005) Heat capacity in proteins. Annu Rev. Phys. Chem. 56, 521–548 [DOI] [PubMed] [Google Scholar]
  • 45. Erijman A., Aizner Y., and Shifman J. M. (2011) Multispecific recognition: mechanism, evolution, and design. Biochemistry 50, 602–611 [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Biological Chemistry are provided here courtesy of American Society for Biochemistry and Molecular Biology

RESOURCES