Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Feb 23.
Published in final edited form as: Biochemistry. 2021 Apr 1;60(15):1165–1177. doi: 10.1021/acs.biochem.0c00865

Stability of the Retinoid X Receptor-α Homodimer in the Presence and Absence of Rexinoid and Coactivator Peptide

Zhengrong Yang 1, Donald D Muccio 2, Nathalia Melo 3, Venkatram R Atigadda 4, Matthew B Renfrow 5
PMCID: PMC9949482  NIHMSID: NIHMS1867554  PMID: 33792309

Abstract

Differential scanning calorimetry and differential scanning fluorimetry were used to measure the thermal stability of human retinoid X receptor-α ligand binding domain (RXRα LBD) homodimer in the absence or presence of rexinoid and coactivator peptide, GRIP-1. The apo-RXRα LBD homodimer displayed a single thermal unfolding transition with a Tm of 58.7 °C and an unfolding enthalpy (ΔH) of 673 kJ/mol (12.5 J/g), much lower than average value (35 J/g) of small globular proteins. Using a heat capacity change (ΔCp) of 15 kJ/(mol K) determined by measurements at different pH values, the free energy of unfolding (ΔG) of the native state was 33 kJ/mol at 37 °C. Rexinoid binding to the apo-homodimer increased Tm by 5 to 9 °C and increased the ΔG of the native homodimer by 12 to 20 kJ/mol at 37 °C, consistent with the nanomolar dissociation constant (Kd) of the rexinoids. GRIP-1 binding to holo-homodimers containing rexinoid resulted in additional increases in ΔG of 14 kJ/mol, a value that was the same for all three rexinoids. Binding of rexinoid and GRIP-1 resulted in a combined 50% increase in unfolding enthalpy, consistent with reduced structural fluidity and more compact folding observed in other published structural studies. The complexes of UAB110 and UAB111 are each more stable than the UAB30 complex by 8 kJ/mol due to enhanced hydrophobic interactions in the binding pocket because of their larger end groups. This increase in thermodynamic stability positively correlates with their improved RXR activation potency. Thermodynamic measurements are thus valuable in predicting agonist potency.

Graphical Abstract

graphic file with name nihms-1867554-f0001.jpg


Human retinoid X receptors (RXRα, RXRβ, and RXRγ) play a key role in many signaling processes which control cellular proliferation, differentiation, and growth in epithelial tissue as well as maintain proper lipid and glucose homeostasis.1 The three RXR isotypes each display different functions in modulating gene transcription of target genes.2 The proteins are approximately 51 kDa and contain two folded domains and three flexible, largely unstructured regions. The DNA binding domain (DBD) binds to specific DNA sequences at the start of target genes3 using two Zn-finger motifs. The ligand binding domain (LBD) at the c-terminal end of RXR is responsible for ligand recognition as well as the associated conformational changes that promote recruitment of coactivator proteins (with replacement of corepressor proteins) important for stabilizing RNA polymerase II to start transcription. The LBD of RXR is connected to the DBD by a short unstructured linker region largely separating the two functions of RXR: specific recognition of target genes and ligand-induced activation of the target gene. Both DBD and LBD contribute to the dimerization interface of the receptor.

The X-ray crystal structures of RXR LBD homodimers alone (apo-RXR LBD) or bound to a variety of ligand agonists (holo-RXR LBD) and coactivator peptides revealed the significant conformational changes that allow for agonist-induced transcriptional activation.414 In apo-RXRα LBD, helix 12 extends away from the core of the domain (Figure 1D) whereas, in the complexes, helix 12 reorients and folds back toward the ligand binding pocket (LBP). Molecular interactions, which occur between the LXXLL15 motif of the coactivator peptide and helices 3, 4, and 12, stabilize the active conformation only observed in the presence of agonists (Figure 1D). The tertiary structural changes induced by coactivator peptide binding also extend to helix 11 residues that interact directly with the agonists. Hydrogen–deuterium exchange mass spectrometry (HDX-MS) and fluorescence polarization studies strongly support that helix 12 is dynamic in both apo- and holo-homodimers and only stabilized when both the coactivator peptide and the agonist are present.8,9,1618 Recent X-ray and HDX-MS studies1923 of nuclear receptor heterodimers containing the entire RXR bound to double stranded DNA with a target recognition site support the validity of those structural and dynamical conclusions drawn from studies on RXRα LBD homodimers.

Figure 1.

Figure 1.

Structures of RXRα LBD, rexinoids, and coactivator peptide. (A) Crystal structure of RXRα LBD homodimer in complex with UAB30 and GRIP-1 (PDB ID: 4K4J). Secondary elements of each dimer are colored according to their positions in the three-layer helical sandwich: helices 1 and 3 are blue; helices 4, 5, 8, and 9 and the β-sheets orange; and helices 6, 7, 10, and 11 teal. Helices 12, which change conformation upon GRIP-1 binding, are red. GRIP-1 peptides are pink. UAB30 rexinoids are displayed as green sticks. (B) Chemical structures of UAB30, UAB110, and UAB111. (C) Peptide sequence of GRIP-1 NR-Box II coactivator peptide. (D) Conformations of helices 3 (blue), 11 (teal), and 12 (red) in the complex, in comparison with their positions in apo-RXRα LBD (gray, PDB ID: 6HN6). (E) Conformations of UAB30 (green), UAB110 (orange), and UAB111 (deep pink) in the ligand binding pocket of RXRα LBD. The rexinoids are displayed as sticks. The amino acid residues lining the pocket are displayed as thin lines.

Our group has designed and studied many rexinoids for the use in cancer treatment and prevention.13,14,2427 UAB30 (Figure 1B) is a novel rexinoid developed by our group currently in phase II clinical trials. Compared to Targretin, the only clinically used rexinoid approved by the FDA,28 UAB30 exhibits high efficacy in cancer prevention without inducing hyperlipidemia which is the dose-limiting toxicity of Targretin.25,2932 UAB110 and UAB111 (Figure 1B) are two rexinoids belonging to another structural class of UAB rexinoids other than UAB30. Each of these three UAB rexinoids bind the RXRα LBD with nanomolar affinities. UAB110 and UAB111 exhibit potencies approximately 40-fold higher than UAB30 in RXR transcriptional activation.14 However, UAB111 induces hyperlipidemia at its effective dose, while UAB110 does not increase triglyceride levels.

To design more potent and nontoxic rexinoids with greater clinical potential and selectivity, we need to fully understand the interactions between the rexinoids and RXRα LBD, and their effects on the structure and dynamics of the RXR homodimers and heterodimers with other nuclear receptors. X-ray crystal structures of RXRα LBD in complex with each UAB rexinoid and a coactivator peptide, GRIP-1 (e.g., RXRα LBD:UAB30:GRIP-1 complex in Figure 1), revealed that only subtle structural differences occurred in the LBP, but there was essentially no difference in the overall backbone fold of the homodimer (Figure 1E).8,9,13,14,33 Using isothermal titration calorimetry (ITC), we showed that GRIP-1 binds to holo-RXR LBD homodimers in complex with each UAB rexinoid with nearly identical affinity and similar enthalpic and entropic signatures.8,9,13,14 The binding thermodynamics of rexinoids to RXR LBD in the absence of coactivator peptide have not been determined, because of limitations to the applicability of ITC in such systems of high-affinity, hydrophobic ligands with small binding enthalpy changes.34,35 Differential scanning calorimetry (DSC) provides an alternative method to obtain thermodynamic data on ligand binding. From the differences in the unfolding temperature and enthalpy with and without a ligand, the thermodynamic parameters of binding of the ligand are determined.36,37

Using DSC, we conducted in this study a complete thermodynamic analysis of the thermal unfolding of human RXRα LBD homodimer in the absence or presence of rexinoid and the coactivator peptide, GRIP-1. The goal was to understand how rexinoids and coactivator peptides modulate the energetics of this protein domain, and to gain knowledge on the thermodynamic parameters of interaction between rexinoids and apo-RXRα LBD. Differential scanning fluorimetry (DSF), a complementary thermal unfolding method to DSC, was used to inform on key aspects of the unfolding of the domain and to rapidly scan unfolding conditions. We obtained the full panel of unfolding thermodynamic parameters on apo-RXRα LBD, including unfolding temperature (Tm), enthalpy (ΔH), entropy (ΔS), and heat capacity change (ΔCpu). These parameters are essential for analyzing binding energetics of agonists and coactivators to RXR but have never been reported before. From the DSC data obtained in the presence of UAB rexinoids and/or GRIP-1, we calculated the thermodynamic parameters of UAB rexinoids binding to apo-RXRα LBD homodimer, which clearly demonstrated for the first time that rexinoid binding was an entropically driven process and correlated well with rexinoid potency as agonists. We also calculated thermodynamic parameters of GRIP-1 binding to holo-RXRα LBD homodimers containing the rexinoids. They corresponded closely to those determined by ITC, which demonstrated the validity of this approach. Thermal unfolding of RXRα LBD homodimer was not completely reversible, and we provided a general guideline on how to deal with this complication to gather meaningful thermodynamic values of ligand binding.

MATERIALS AND METHODS

Protein Purification.

Overexpression and purification of human RXRα LBD (amino acid residues T223–T462) were performed as described by Xia et al.9 Briefly, the His6-tagged RXRα LBD fusion protein was purified using a HiTrap Ni-chelating column (GE Healthcare, Piscataway, NJ), and the His6-tag was removed by incubating with human α-thrombin overnight. The complete removal of His6-tag was confirmed by MALDI-TOF mass spectrometry. The tagless RXRα LBD was further purified on a HiLoad Superdex 75 size exclusion column (SEC, GE Healthcare, Piscataway, NJ). The fractions containing RXRα LBD homodimers or tetramers (Figure S1A) were pooled separately. SDS-PAGE and MALDI were used to establish a purity of >95% and a mass of the monomers at 27 283 Da.

Fluorescence-Based Binding Affinity Assay.

The dissociation constant (Kd) of the rexinoids binding to apo-RXRα LBD was determined in a fluorescence-based binding assay as previously described.24 The fluorescence quenching data were plotted in Origin 7 (OriginLab, LLC.), and nonlinear least-squares curve fitting was performed to obtain an apparent Kd based on the following equation:38

%quenching=100(1Z)(2Rt)/{(Nt+Rt+Kd)+[(Nt+Rt+Kd)24NtRt]1/2} (1)

where Rt was the total concentration of the test rexinoid, Nt was the total concentration of apo-RXRα LBD in monomer unit, and Z represented the residual unquenched % fluorescence after the binding site was completely saturated. Nt was determined from the binding curve of UAB111 which had the highest binding affinity. This value was fixed for fitting of UAB30 and UAB110 binding curves.

Differential Scanning Calorimetry.

Calorimetry experiments were performed using the VP-Capillary DSC system (Malvern Instruments, Westborough, MA) in 0.130 mL cells at a heating rate that varied between of 0.5 and 4 °C/min. An external pressure of 2.0 atm was maintained to prevent possible degassing of the solutions upon heating. Purified apo-RXRα LBD homodimer freshly eluted from SEC was dialyzed in 2 L of DSC buffer (10 mM sodium phosphate, pH 7.0, 50 mM NaCl, 0.5 mM EDTA, and 1 mM TCEP; sodium phosphate was replaced by sodium borate for pH 8.5–9.5) for >4 h and used within 24 h. Holo-RXRα LBD homodimers were prepared by adding rexinoids from a 100-fold concentrated stock in DMSO into apo-RXRα LBD homodimer. The samples were incubated at 22 °C for 10 min and protected from light. GRIP-1 was added into the holo-RXRα LBD homodimers from a 10 mM stock solution in water. To prepare the buffer control, identical amounts of rexinoid and GRIP-1 were added into the dialysis buffer. DSC data analyses were performed using the built-in analysis module in Origin 7 provided by the DSC manufacturer.39 Briefly, after subtraction of the buffer scan, and normalization to molar heat capacity (Cp), a cubic baseline was subtracted from the molar Cp curve to set the pre- and post-transition baselines to zero (Figure S2). The thermal unfolding temperature (Tm) was determined from the maximum of the Cp curve. The calorimetric enthalpy (ΔHc) was obtained by integrating the Cp curve. The apparent van’t Hoff enthalpy (ΔHv) was obtained by fitting the curve to a built-in model, MN2STATE, which fit ΔHv independently of ΔHc (Figure S3). Equations for the calculation of thermodynamic parameters are listed in Appendix I of the Supporting Information.

Differential Scanning Fluorimetry.

Using the Prometheus NT.48 instrument (NanoTemper Technologies, LLC, South San Francisco, CA) with 48 capillary chambers, each sample was excited at 290 nm, and emission was detected simultaneously at 330 and 350 nm. The first derivative of the fluorescence signal at 350 nm relative to 330 nm (F350/F330) versus temperature produced a DSC-like thermogram. The first derivative of fluorescence intensity at each wavelength exhibited the same thermal unfolding profiles. Because the UV–vis absorption of the rexinoids interfered with the detection of unfolding using F350/F330, DSF data of fluorescence intensity at 350 nm were used to obtain Tm. Sample preparation for DSF was the same as for DSC. Each capillary required 10 μL to load. Duplicate measurements were performed on each sample. Tm of each sample was automatically determined by the built-in analysis software and tabulated in an Excel output file. Global curve fitting of the apo-RXRα LBD DSF data obtained at different scan rates was performed using previously published equations40 written in Origin C.

Circular Dichroism.

CD spectra were recorded in an OLIS CD spectrophotometer (OLIS, Athens, GA). Quartz cuvettes with a path-length of 0.02 cm were used. The protein concentration was 12 μM dimer. The CD cell holder was heated using an external water bath to the desired temperatures. CD spectra were recorded from 260 to 190 nm. Buffer baselines were recorded at the same temperatures and subtracted from the protein spectra. Total data collection time was ~5 min for each spectrum. The secondary structural content was determined by the CDNN program (CDNN: CD Spectra Deconvolution, Version 2.1, Universität—Böhm, 1997). For the completely denatured sample in guanidine–HCl, the sample was prepared by adding guanidine–HCl as a solid into the thermally unfolded protein solution to reach a final concentration of 5 M. The sample was then cooled in the cell holder, and the CD spectra were collected at 25 °C. The final protein concentration was adjusted based on the volume change after the addition of the denaturant. Optical interference from the denaturant prevented scans below 210 nm.

Isothermal Titration Calorimetry.

ITC experiments were performed on an Auto-iTC200 system (Malvern Instruments, Westborough, MA). The buffer used in ITC was the DSC buffer (pH 7.0) with 1% w/v DMSO. For titration of UAB30 with apo-RXRα LBD, the ITC sample cell contained 5 μM UAB30 (titrand), and the syringe contained 20 μM apo-RXRα LBD homodimer (titrant). For titration of RXRα LBD:rexinoid with GRIP-1, the titrand was 12.5 μM RXRα LBD dimer and 35 μM test rexinoid, and the titrant was 350 μM GRIP-1. Each titration experiment consisted of 16 injections of 2.5 μL of titrant into titrand at 10, 20, or 30 °C. Background mixing heat was determined from injections of titrant into the same buffer without titrand. Data analysis was performed using the built-in analysis module in Origin 7 provided by the ITC manufacturer. Monomeric protein concentration was used for analysis to obtain the stoichiometry of binding to each RXRα LBD monomer.

RESULTS AND DISCUSSION

I. Thermal Unfolding of Apo-RXRα LBD Homodimers.

The ΔG of apo-RXRα LBD unfolding at 30 °C was previously determined by Harder et al.41 using isothermal chemical denaturation. However, the enthalpic and entropic changes of the unfolding process were not obtained, nor was the unfolding heat capacity change which contributes significantly to the temperature dependence of ΔG.

1.1. Partial Reversibility of Transition.

To determine the stability of apo-RXRα LBD homodimer at 37 °C and other temperatures, thermal unfolding of this homodimer was measured by DSC from 35 to 70 °C using a scan rate (v) of 4 °C/min. A single endotherm was observed centered at the maximum thermal unfolding temperature (Tm) of 58.2 ± 0.1 °C, with a calorimetric unfolding enthalpy (ΔHc) of 665 ± 4 kJ/mol (Figure 2, trace 1). Upon rescan, no thermal unfolding endotherm was observed, indicating that the unfolding of the homodimer was irreversible when heated to 70 °C (Figure 2, trace 2). To determine whether partial reversibility occurred during the unfolding process, the extent of reversibility at various points through the DSC endotherm was examined by consecutive scans of the same homodimer sample from 35 °C to various ending temperatures (Figure S4). DSC endotherms were also obtained on homodimers systematically heated for 1 min at various temperatures and then rapidly cooled in an ice bath. When the incubation temperature was 57.3 °C, which corresponded to 25% unfolding [the extent of unfolding at temperature T was determined by the ratio of integrated area from 35 °C to T, divided by the total area under the DSC curve], an endotherm was observed with the same Tm, but with approximately 80% the intensity of the first scan (Figure 2, trace 3). As the incubation temperature was increased to 58.2 °C (50% unfolding), the ΔHc was decreased to 50% (Figure 2, trace 4). The ΔHc was further decreased to 30% when the incubation temperature was increased to 60.0 °C (75% unfolding, Figure 2, trace 5). These data support that the unfolding of the homodimer was partially reversible through the endotherm.

Figure 2.

Figure 2.

Thermal unfolding of apo-RXRα LBD is partially reversible. DSC molar heat capacity profile of 1.5 μM apo-RXRα LBD homodimer. (1) First scan; (2) rescan; (3) after heating at 57.3 °C for 1 min; (4) after heating at 58.6 °C for 1 min; and (5) after heating at 60.0 °C for 1 min.

The scan rate dependence of the thermal unfolding of apo-RXRa LBD was next examined. The scan rate, v, of the DSC measurement was decreased from 4.0 to 0.5 °C/min. As displayed in Figure 3A and summarized in Table 1, Tm decreased systematically as v decreased, reaching 55.6 ± 0.1 °C when v was 0.5 °C/min. ΔHc remained nearly constant at 670 kJ/mol regardless of v. An apparent van’t Hoff enthalpy (ΔHv) was determined by nonlinear least-squares curve fitting to a two-state model, whereby a natively folded protein, N, unfolds reversibly to U: N ↔ U.42 ΔHv systematically increased from 815 ± 13 kJ/mol when v was 4 °C/min to 1208 ± 4 kJ/mol when v was 0.5 °C/min.

Figure 3.

Figure 3.

Equilibrium unfolding parameters of apo-RXRα LBD obtained by extrapolation to infinite scan rate. (A) DSC molar heat capacity profiles for 1.5 μM apo-RXRα LBD homodimer at different scan rates, v (from left to right): 0.5, 1.0, 2.0, 3.0, and 4.0 °C/min. (B) Tm, ΔHc, and ΔHv as a function of v−1. The Tm and ΔHv values were fitted linearly to v−1. The equilibrium unfolding parameters were obtained by extrapolation to zero v−1.

Table 1.

Scan Rate Dependence of the Thermal Unfolding Parameters of apo-RXRα LBD

scan rate DSC DSFa
v (°C/min) Tm (°C) ΔHc (kJ/mol) ΔHv (kJ/mol) Tm (°C)
0.5 55.6 ± 0.1 681 ± 4 1208 ± 4 55.2 ± 0.1
1 56.1 ± 0.1 669 ± 1 1133 ± 4 55.6 ± 0.1
2 57.2 ± 0.1 677 ± 13 966 ± 8 56.4 ± 0.1
3 57.8 ± 0.02 677 ± 1 874 ± 13 56.7 ± 0.1
4 58.2 ± 0.1 665 ± 4 815 ± 13 57.0 ± 0.1
infiniteb 58.7 ± 0.2 673 ± 8 736 ± 25 57.4 ± 0.1
a

Tm measured from DSF monitored at 350 nm.

b

Extrapolation to v−1 is 0 yielded the equilibrium unfolding parameters.

The thermal unfolding of apo-RXRα LBD was also monitored by measuring the changes in intrinsic fluorescence at 330 and 350 nm (Figure S5). The changes in tryptophan (Trp) emission intensity at 330 or 350 nm reflected the change in Trp environments during unfolding. The differential scanning fluorimetry (DSF) curves of apo-RXRα LBD, which contains two Trp residues, displayed only one unfolding transition that was similar to the DSC endotherms. The trends in the Tm at different scan rates matched those determined by DSC (Table 1), suggesting that both DSC and DSF detected the same global unfolding event. The Tm values determined by DSF were systematically lower than the Tm measured by DSC by ~1 °C. The lower Tm measured by DSF relative to those by DSC was observed for other protein unfolding processes.4345 Trp residues in folded states may sense changes in their local environments at a slightly lower temperature than onset of the global unfolding measured calorimetrically.46

The lack of reversible unfolding transition in DSC rescans and the dependence of Tm on v indicated that the thermal unfolding of RXRα LBD was not in complete equilibrium throughout the heating process. DSC irreversibility is a common phenomenon, especially for multidomain or oligomeric proteins.47 Often protein aggregation occurs at higher temperatures, which prevents reversibility.48 However, these irreversible DSC transitions yield equilibrium data if the unfolding follows the Lumry–Eyring model: whereby N unfolds reversibly to U, which converts to a denatured state (D) slowly and irreversibly: N ↔ U → D.4951 If the irreversible step occurs slower than the rate of protein unfolding and refolding, then thermodynamic data for the reversible step are obtained by extrapolation of measured data to infinite scan rate.47,5054 For apo-RXRα LBD, the Tm and ΔHv were plotted versus v−1 (Figure 3B); each thermal parameter was linearly dependent on v−1 for v faster than 0.5 °C/min. The extrapolated equilibrium unfolding parameters (Table 1) were as follows: Tmeq = 58.7 ± 0.2 °C, ΔHveq = 736 ± 25 kJ/mol, and ΔHceq = 673 ± 8 kJ/mol (using the average of the apparent ΔHc at different v values).

1.2. Two-State Unfolding without Dimer Dissociation.

The ratio of the van’t Hoff enthalpy to the calorimetric enthalpy, ΔHveqHceq, was essentially 1 using the extrapolated parameters. A cooperative unit of 1 when both ΔHceq and ΔHveq were determined based on per mole of dimer indicated that the native dimer was the cooperative unfolding unit. The ratio of 1 for ΔHvHc was not expected if the dimeric protein dissociated during unfolding: N2 ↔ 2U. For this unfolding equilibrium, the unfolding endotherm is asymmetrical about its midpoint, and the apparent ΔHvHc is about 0.754 (see Figure S6 for a simulated DSC transition based on the N2 ↔ 2U model, in comparison to the N ↔ U model). The asymmetry and low ΔHvHc were not observed in any of the DSC traces obtained in this study, suggesting that the native dimer unfolded without significant dissociation to monomers.5557

To examine this further, native gels were used to determine the aggregation state of the homodimer at three different temperatures: 57 °C which was below Tm, 58.7 °C which was the extrapolated equilibrium Tm, and 65 °C which was above the completion of the endotherm. As displayed in Figure 4, apo-RXRα LBD homodimer migrated near 50 kDa, consistent with a dimer, whereas the tetramer migrated near 100 kDa. When the dimer was heated at 57 °C for 1, 2, and 5 min and rapidly cooled in an ice bath, the band corresponding to the dimer was observed, but the intensity decreased as heating time increased. In addition, a species that migrated significantly above that of the unheated dimer emerged and increased in intensity, consistent with an aggregate of much higher molecular weight. When the heating temperature increased to 58.7 °C, less dimer and more aggregated species were observed. When heated to 65 °C for these times, the dimer disappeared, and only a highly aggregated species was present in the native gels. A band consistent with monomer was not observed in any lane.

Figure 4.

Figure 4.

Changes in apo-RXRα LBD aggregation state during thermal unfolding. Native PAGE of 5 μM apo-RXRα LBD homodimer incubated at different temperatures for 1, 2, and 5 min, and then rapidly cooled in an ice bath. Lane 1: Fraction from SEC containing a mixture of tetramer and dimer that was not heated. Lane 2: Fraction from SEC containing only dimer that was not heated. Lanes 3–5: dimer heated at 57 °C (~25% unfolding) for 1, 2, and 5 min. Lanes 6–8: dimer heated at 58.7 °C (50% unfolding) for 1, 2, and 5 min. Lanes 9–11: dimer heated at 65 °C (past 100% unfolding) for 1, 2, and 5 min.

Circular dichroism (CD) and fluorescence spectroscopy were measured at several temperatures to inform on changes in secondary and tertiary structure during the endotherm. The X-ray crystal structure of apo-RXRα LBD contained 12 helices composing about 66% of its secondary structure.4,7 The 222 and 208 nm negative CD signals (Figure 5A), which are significant in helical protein structures, were present in CD spectra at temperatures up to 56 °C.9 The magnitude of the 208 and 222 nm negative bands decreased ~20% when heated to temperatures near to Tm measured by DSC. These signals decreased ~40% when the dimer was heated to 75 °C, slightly higher than the temperature of the end of the DSC endotherm. No further decrease in CD signals was observed when the sample was incubated at 75 °C for 10 more minutes. This indicated that the secondary structure of apo-RXRα LBD was not completely unfolded throughout the DSC endotherm. Based on these CD spectra, the helical content of the native homodimer was 57%, and 33% remained after heating to 75 °C. In contrast, the 222 and 208 nm signals were completely lost for the homodimer in the presence of 5 M guanidine–HCl at 25 °C, which completely denatured and dissociated the dimer into monomers.41

Figure 5.

Figure 5.

Spectroscopic changes in apo-RXRα LBD during thermal unfolding. (A) Circular dichroic spectra of 12 μM apo-RXRα LBD homodimer at different temperatures and in the presence of 5 M guanidine–HCl at 25 °C. (B) Fluorescence spectra of 0.05 μM apo-RXRα LBD homodimer at different temperatures and in the presence of 5 M guanidine–HCl at 25 °C. The excitation wavelength was 280 nm.

Each monomer of apo-RXRα LBD contains two Trp: W282 in helix 2 and W305 in helix 4, which is near the rexinoid binding site. An intense fluorescence band centered at 335 nm was observed at 35 °C, consistent with the hydrophobic environment of the indole group of the two Trp in a folded state (Figure 5B). Upon heating the protein to higher temperatures, the Trp fluorescence intensity decreased due to thermal quenching,58 and the emission maximum wavelength (λmax) gradually red-shifted to 344 nm at 70 °C. A red-shift of λmax to 355 nm was observed for the completely denatured homodimer in 5 M guanidine–HCl. The much less red-shifted signals for the thermally denatured apo-RXRα LBD indicated that either one or both of the Trp were still in a partially hydrophobic environment when heated to 70 °C. Taken together, these thermodynamic and spectral data are most consistent with the native homodimer being converted to a partially unfolded and aggregated apo-RXRα LBD species above Tm.

1.3. Thermodynamic Unfolding Parameters at 37 °C.

Using a v of 4 °C/min for DSC measurements, the time needed to unfold the protein was less than 3 min, which minimized the effects of the irreversible processes during the unfolding transition. This is reflected in the fact that the unfolding parameters obtained at 4 °C/min were very similar to the extrapolated equilibrium parameters (Table 1 and Figure S3). Therefore, all subsequent DSC experiments were conducted at 4 °C/min. To modulate unfolding Tm and enthalpy for determining the unfolding heat capacity change (ΔCpu), the pH of the buffer was changed. Due to the rapidity of the measurement and lower sample amounts, DSF was first used to survey a range of pH values between 5 and 9 (Figure S7). DSF showed that apo-RXRα LBD was most stable at pH 7.0, and it required an increase of 2 pH units to induce significant decreases in Tm.

DSC was performed on apo-RXRα LBD homodimer at four different pH values: 7.0, 8.6, 9.0, and 9.5 (Figure 6A). To ensure that buffer ionization enthalpy did not contribute to the observed ΔHc, inorganic buffers with low ionization enthalpies were used. Additional DSC experiments were performed using buffers with higher ionization enthalpies. The observed ΔHc displayed no dependence on buffering components (Figure S8). ΔHc obtained in inorganic buffers was fitted linearly to Tm (Figure 6B). From the slope of the line, ΔCpu was determined to be 15 ± 1 kJ/(mol K). Using this ΔCpu and the extrapolated values of Tmeq and ΔHceq from the DSC measurements (Table 1), ΔH of unfolding for the apo-RXRα LBD homodimer was calculated to be 347 ± 21 kJ/mol at 37 °C. ΔS of unfolding was 1.01 ± 0.06 kJ/(mol K); −TΔS was −314 ± 17 kJ/mol, and ΔG of unfolding was 33 ± 3 kJ/mol at 37 °C. A ΔG of 32 ± 3 kJ/mol was calculated using parameters obtained at a v of 4 °C/min, which was within experimental error of the ΔG calculated from the extrapolated parameters. Apo-RXRα LBD homodimer was most stable at 20 °C, with a ΔG of 43 ± 3 kJ/mol.

Figure 6.

Figure 6.

Determination of the unfolding heat capacity change (ΔCpu) of apo-RXRα LBD homodimers. (A) DSC molar heat capacity profiles for 1.5 μM apo-RXR-LBD homodimer using a scan rate of 4 °C/min at different pH values (from left to right): 9.5, 9.0,8.6, and 7.0. (B) ΔCpu was determined by a linear regression of ΔHc versus Tm. The R value for the linear fit was 0.99.

According to Harder et al.,41 the chemical denaturation of apo-RXRα LBD in guanidine–HCl is reversible and follows a three-state mechanism: N2 ↔ 2I ↔ 2U, wherein “I” is a monomeric, partially unfolded, intermediate. Although the ΔG change of the first step accounts for ~80% of the total unfolding ΔG, spectroscopy data support that the monomeric intermediate retains significant native secondary structure, and the two Trp residues remain partially buried. It is during the second step of unfolding where this unfolding intermediate, I, loses all secondary structures, and the Trp are completely solvent exposed. The authors estimate that the ΔG of chemical denaturation for the first step is 35 ± 0.8 kJ/mol at 30 °C. From our study, the estimated thermal unfolding ΔG for apo-RXRα LBD homodimers at 30 °C was 39 ± 2 kJ/mol. Likewise, the spectral data on the thermally unfolded protein (Figure 5) are similar to those estimated for the monomeric intermediate in guanidine–HCl. Guanidine–HCl is well-known for its capability to prevent aggregation of partially or fully unfolded proteins because of its chaotropic and ionic properties.5961 It is possible that, in our study, DSC detected the unfolding of the native dimer to a partially unfolded intermediate very similar to that detected by Harder et al. However, because of the lack of chaotropes such as guanidine–HCl to stabilize a monomer, the unfolded state was thermally aggregated without a significant heat signature.

The data presented strongly support the model N2 ↔ I2 → D, where I2 is a partially unfolded, dimeric intermediate. This model40 was used to globally fit a set of four DSF data at different scan rates (1 to 4 °C/min, Figure S5). The best fitted parameters for the reversible first step were Tm = 57.5 °C and ΔHv = 710 kJ/mol. These values were very similar to those obtained by extrapolation to infinite scan rate (Table 1). The second, irreversible step was best fitted with a single rate-constant, k, of 2.1 min–1. To gain more understanding on how the rate of the irreversible step affects the extrapolated thermodynamic parameters obtained by fitting the DSC to a reversible model, a series of simulations were performed using k values of 0.5, 2.1, and 8 min−1 (Figure S9A). The simulated curves were fitted to the two-state reversible model, N2 ↔ I2, and the resulting thermodynamic parameters, Tm, ΔHv, and ΔHc, were plotted against v−1 (Figure S9B). When k was 0.5 min−1, which was relatively slow compared to the scan rates, the extrapolated Tmeq and ΔHveq were essentially identical to the input values for the simulation (Figure S9B, green circles on the axes). When k was 8 min−1, which was relatively fast compared to the scan rates, Tmeq and ΔHveq were different from the input values. Tmeq was 1.5 °C lower, and ΔHveq was 23% higher. In contrast, the observed ΔHc value was constant and equal to the input value. When k was 2.1 min−1, which was our fitted value using the Lumry–Eyring model, Tmeq was lower by only 0.5 °C, and ΔHveq was higher by 9%. Again, ΔHc was scan-rate-independent and equal to the input value. In short, if partial reversibility was established, and the rate of the irreversible step was no faster than the highest scan rate, then Tmeq obtained by extrapolation to infinity v was within 1% of the input value. ΔHveq could be different from the input value, but ΔHceq was equal to the input value. Therefore, ΔHceq became the logical choice to use for the calculation of other thermodynamic parameters, such as ΔG.

The ΔG of unfolding for apo-RXRα LBD was unusually low compared to other dimeric proteins, which may suggest a weak dimeric interface ready to dissociate into monomers at sub-micromolar concentrations to facilitate the formation of heterodimers with other receptors.41 However, our thermal unfolding data indicated that the dimer interface was relatively stable as it did not dissociate throughout the unfolding endotherm. This was more consistent with the DNA binding properties and transcriptional activities observed for the full-length RXR homodimer.3,62 Moreover, the thermal unfolding of apo-RXRα LBD homodimer was accompanied by a relatively low unfolding ΔHc of 673 kJ/mol at Tm, i.e., a low specific heat of 12.5 J/g, which was substantially lower than the specific heat of a typical soluble globular protein at the same Tm (29–38 J/g according to Murphy and Freire55). The low specific heat may be caused by three factors. First, it may be due to incomplete unfolding as demonstrated by the CD and fluorescence spectroscopy data on the thermally unfolded protein. Second, it may be due to homodimers generally unfolding with lower specific heats than single domain proteins.63 Third, it may indicate that the native apo-homodimer is less compactly folded and more dynamic than typical, well-folded globular proteins. Such structural fluidity likely relates to the lack of rexinoid bound to its hydrophobic pocket and the dynamic helix 12 region of the domain.8,9

II. Thermal Unfolding of Holo-RXRα LBD Homodimers Bound with Rexinoids.

The stability of RXRα LBD bound with rexinoid (holo-RXRα LBD) was examined next to determine the relationship between the rexinoid structure and holo-RXRα LBD unfolding energetics. Rexinoids UAB30, UAB110, and UAB111 quenched more than 90% of the protein fluorescence signal of apo-RXRα-LBD at 337 nm at 25 °C when the ratio of the protein to the rexinoids reached 1:1. The dissociation constants (Kd) determined from the binding isotherm based on fluorescence quenching was 38 ± 14 nM for UAB30, 22 ± 6 nM for UAB110, and 2.4 ± 0.4 nM for UAB111.

DSF was first used to determine the rexinoid concentration required to saturate the binding sites at Tm because Kd was expected to be dependent on temperature. For UAB30, the DSF unfolding curve continuously shifted to higher Tm (Figure S10A) when the ligand concentration was increased from 1.25 to 25 μM using a dimer protein concentration of 2.5 μM. There was no measurable increase in Tm when the rexinoid concentration was doubled to 50 μM, indicating that effective saturation was reached at rexinoid concentrations above 25 μM (5:1 molar ratio, based on one binding site/monomer). An increase in Tm of 4.4 ± 0.1 °C in the presence of UAB30 at a 2:1 rexinoid/monomer molar ratio was consistent with UAB30 strongly binding to the native homodimer at 25 °C and at elevated temperatures. For UAB110 and UAB111, the shifts in Tm in the presence of rexinoids at a 2:1 rexinoid/monomer molar ratio were 7 and 7.5 °C, respectively, which were substantially higher than the shifts caused by UAB30 (Figure S10B) and consistent with their higher binding affinities.

DSF was also used to examine the effect of scan rate on the Tm of the holo-RXRα LBD homodimer bound with each rexinoid. The linear regression of Tm values of each complex with respect to v−1 yielded slopes almost identical to that of apo-RXRα LBD (Figure 7). This suggested that the rate of the irreversible step in the Lumry–Eyring unfolding model was similar for both holo-homodimers and the apo-homodimer.

Figure 7.

Figure 7.

DSF Tm values as a function of v−1 for RXRα LBD homodimers with and without rexinoids or coactivator peptide. DSF Tm values of 2.5 μM RXRα LBD homodimer at pH 7.0, in the presence of no rexinoid or GRIP-1 (Apo), 0.4 mM GRIP-1 (Apo + GRIP), 30 μM UAB30 (UAB30), 30 μM UAB30 and 0.4 mM GRIP-1 (UAB30 + GRIP), 10 μM UAB110 (UAB110), 10 μM UAB110 and 0.4 mM GRIP-1 (UAB110 + GRIP), 10 μM UAB111 (UAB111), or 10 μM UAB111 and 0.4 mM GRIP-1 (UAB111 + GRIP). The scan rate, v, was 4.0 °C/min. The lines are linear regressions of Tm values with respect to v−1. All samples contained 1% DMSO.

To gather thermodynamic data on holo-RXRα LBD dimers, DSC was performed. The effect of DMSO, a cosolvent necessary for solubilizing the rexinoids, was first examined. Using a v of 4 °C/min, the Tm, ΔHc, and ΔHv values of apo-RXRα LBD homodimer determined by DSC in the presence of 1% w/v DMSO (Table 2) were within experimental errors of those values obtained in the absence of DMSO (Table 1). This sample was used as the point of reference (DMSO control) for comparison with holo-RXRα LBD dimers containing rexinoids.

Table 2.

Thermodynamic Parameters of Unfolding of Apo-RXRα LBD and Holo-RXRα LBD with and without GRIP-1

Tm (°C) ΔHc (kJ/mol) ΔHv (kJ/mol) ΔCpu (kJ/(mol K)) ΔH (37 °C) (kJ/mol) TΔS (37 °C) (kJ/mol) ΔG (37 °C) (kJ/mol)
apo-RXRα LBDa 58.7 ± 0.2 673 ± 8 736 ± 25 15 ± 1 347 ± 21 −314 ± 17 33 ± 3
DMSO controlb 57.9 ± 0.4 690 ± 8 803 ± 4 15 ± 1 376 ± 17 −342 ± 17 33 ± 2
+GRIP-1 59.3 ± 0.2 690 ± 29 811 ± 13 14 ± 2 389 ± 52 −353 ± 48 36 ± 7
+UAB30 63.6 ± 0.1 803 ± 4 836 ± 13 17 ± 1 371 ± 26 −326 ± 23 45 ± 4
+UAB110 66.6 ± 0.5 849 ± 13 811 ± 17 16 ± 2 362 ± 37 −310 ± 32 52 ± 8
+UAB111 66.9 ± 0.3 840 ± 21 823 ± 21 17 ± 1 362 ± 24 −309 ± 20 53 ± 5
+UAB30/GRIP-1 67.5 ± 0.3 932 ± 33 924 ± 25 17 ± 1 374 ± 35 −317 ± 30 58 ± 8
+UAB110/GRIP-1 69.9 ± 0.5 995 ± 17 924 ± 8 18 ± 1 393 ± 35 −327 ± 30 66 ± 8
+UAB111/GRIP-1 70.4 ± 0.2 991 ± 29 928 ± 8 20 ± 2 380 ± 35 −314 ± 30 66 ± 9
a

Equilibrium unfolding parameters obtained by extrapolating to v−1 is zero, in the absence of DMSO.

b

DMSO control was apo-RXRα LBD containing 1% w/v DMSO.

Data for the DMSO control and all others were obtained at a v of 4 °C/min and in the presence of 1% DMSO.

Rexinoid concentrations that caused maximum increases in Tm determined by DSF (30 μM for UAB30, and 10 μM for either UAB110 or UAB111) were used in DSC to minimize unwanted effects of the rexinoids at high concentrations (Figure S11). Similar to apo-RXRα LBD homodimer, only one DSC unfolding transition was observed in the presence of rexinoids (Figure 8), and the Tm of each holo-RXRα LBD increased due to the shift of unfolding equilibrium toward the native state caused by rexinoid binding to the native apo-dimer. To verify that the unfolding parameters obtained at a v of 4 °C/min for the holo-RXRα LBD homodimers were good approximations of the equilibrium parameters, the scan-rate dependence of the DSC endotherms of RXRα LBD:UAB30 (holo-RXRα LBD homodimer bound with UAB30) was examined (Figure S12 and Table S1). Similar to apo-RXRα LBD, a linear relationship existed between each unfolding parameter and v−1 (Figure S12B). The extrapolated equilibrium unfolding parameters for RXRα LBD:UAB30 were Tmeq = 63.5 ± 0.1 °C, ΔHveq = 761 ± 38 kJ/mol, and ΔHceq = 798 ± 8 kJ/mol (Table S1). The unfolding parameters obtained at a v of 4 °C/min were very similar to these extrapolated values. The ratio of ΔHveq to ΔHceq was close to 1, as found for the apo-homodimer.

Figure 8.

Figure 8.

RXRα LBD is strongly stabilized by rexinoid binding and coactivator peptide GRIP-1. DSC molar heat capacity profiles of 1.5 μM RXRα LBD at pH 7.0, in the presence of no rexinoid or GRIP-1 (Apo), 0.4 mM GRIP-1 (+GRIP), 30 μM UAB30 (UAB30), 30 μM UAB30 and 0.4 mM GRIP-1 (UAB30 + GRIP), 10 μM UAB110 (UAB110), 10 μM UAB110 and 0.4 mM GRIP-1 (UAB110 + GRIP), 10 μM UAB111 (UAB111), or 10 μM UAB111 and 0.4 mM GRIP-1 (UAB111 + GRIP). The scan rate, v, was 4.0 °C/min. All samples contained 1% DMSO.

An unfolding Tm of 66.6 ± 0.5 °C was obtained for RXRα LBD:UAB110 at a v of 4 °C/min, which was 3 °C higher than the Tm of RXRα LBD:UAB30 (Figure 8 and Table 2). The ΔHc was 849 ± 13 kJ/mol, which was approximately 260 kJ/mol higher than the apo-homodimer in 1% DMSO. RXRα LBD:UAB111 displayed similar unfolding parameters as UAB110:RXRα LBD, with a Tm of 66.9 ± 0.3 °C and a ΔHc of 840 ± 21 kJ/mol. Both RXRα LBD:UAB110 and RXRα LBD:UAB111 also exhibited a ΔHvHc ratio close to unity, consistent with the holo-RXRα LBD homodimers being the cooperative unfolding unit.

To determine the ΔCpu values of the holo-RXRα LBD homodimers, Tm and ΔHc were measured at pH 8.8 and 9.5. In a similar manner to apo-RXRα LBD homodimers, DSF was first used to evaluate whether pH modulates the degrees of stabilization caused by the presence of rexinoids (Figure S13 and Table S2). The shifts in Tm caused by rexinoid were within experimental errors at all three pH values. Using DSC, the measured Tm values and enthalpies were lower as pH decreased, but the relative positions and magnitudes of all DSC endotherm remained the same (Figure S14). The ΔHc value of each holo-homodimer was fitted linearly with respect to Tm to determine ΔCpu values (Table 2).

Using ΔCpu, Tm, and ΔHc from DSC measurements at a v of 4 °C/min, ΔH, ΔS, and ΔG of unfolding at 37 °C for each holo-RXRα LBD homodimer were calculated (Table 2). Compared to the apo-RXRα LBD homodimer DMSO control, the holo-RXRα LBD homodimers had significantly higher ΔG values consistent with tight binding and strong stabilization of the native state by the rexinoids. The degrees that UAB110 and UAB111 stabilized the native apo-RXRα LBD homodimer (ΔΔG = 19 and 20 kJ/mol, respectively) were notably higher than UAB30 (ΔΔG = 12 kJ/mol), which was consistent with the improved binding affinities of UAB110 and UAB111 over that of UAB30.

The difference in ΔH between the DMSO control and holo-RXRα LBD homodimers (ΔΔH) at 37 °C was essentially zero for each of the three rexinoids. A near-zero value for ΔΔH suggested that the binding of rexinoid to apo-RXRα LBD at physiological temperatures was accompanied by a small binding enthalpy. In order to gather information on the thermodynamics of rexinoid binding to apo-RXRα LBD, ITC measurements were performed on UAB30 between 10 and 30 °C. No ITC isotherms were obtained due to low solubility of the rexinoid in water. When reverse titrations were performed at these temperatures, the measured heat changes were within the noise levels (Figure S15), which suggested a low-enthalpy binding reaction at these temperatures, consistent with the DSC results.

The ligand binding pocket (LBP) of RXRα LBD containing UAB30 is lined with 16 hydrophobic residues contributed by 4 protein helices that surround the UAB rexinoids.8,9,13,14 The main forces that stabilize the interactions between the rexinoids and LBD residues are the ionic interaction between the carboxylate group and Arg316, and numerous hydrophobic contacts throughout the binding pocket. Burying nonpolar atoms upon ligand binding leads to a negative heat capacity change (ΔCpa).9,37 Exposing these buried nonpolar surface areas upon protein unfolding and ligand dissociation increases the unfolding ΔCpu. The ΔCpu of each of the three holo-RXRα LBD homodimers was slightly higher than that of the apo-RXRα LBD homodimer in 1% DMSO (+1 to 2 kJ/(mol K)). The − TΔS values of the holo-RXRα LBD homodimers were each less negative than that of the apo-RXRα LBD homodimer in DMSO (+17 to 31 kJ/mol), which contributed favorably to the ΔG values. Therefore, rexinoid binding to apo-RXRα LBD appeared to be entropically driven at physiological temperatures. Crystallographic data indicate that, upon binding to UAB110 or UAB111, the size of the LBP expanded nearly 20% (compared to LBP in the presence of UAB30) to accommodate the two larger rexinoids (Figure 1E).14 The contact surface areas of these two rexinoids were about 100 Å2 larger than those observed for UAB30. Burying more hydrophobic surfaces likely resulted in more favorable entropy changes during the binding process and, hence, the more favorable ΔG of folding for holo-RXRα LBD homodimers bound to UAB110 or UAB111. In addition, the increase in stability of the holo-RXRα LBD homodimers could also arise from reduced dynamics due to rexinoid binding. HDX MS studies have revealed that the dynamics of helices 3 and 11 are significantly decreased when UAB30 is bound to the homodimer8 and reduced even further in the presence of UAB110 and UAB111 whose binding affinities were higher than that of UAB30 (unpublished data). In summary, the LBP was expanded but more thermodynamically stable when bound with UAB110 or UAB111. How these conformational changes affect coactivator binding was examined next.

III. Thermal Unfolding of Holo-RXRα LBD Homodimers Bound with Rexinoids and a Coactivator Peptide.

It was previously shown by using ITC measurements that the 13-mer coactivator peptide, GRIP-1 (Figure 1C), binds to holo-RXRα LBD complexes with a micromolar Kd at 25 °C in an exothermic reaction with binding enthalpies in the range from −36 to −42 kJ/mol per monomer.8,14 Because the binding affinity is expected to be weaker at Tm, DSF was first used to determine the GRIP-1 concentration required to saturate the peptide binding site at Tm (Figure S16). The Tm of UAB30:RXRα-LBD:GRIP-1 unfolding incrementally increased when GRIP-1 concentration was increased from 12.5 to 800 μM using a dimer protein concentration of 2.5 μM and 15 μM UAB30. A concentration of 400 μM GRIP-1 was chosen for the collection of thermodynamic data on RXRα LBD homodimer bound with rexinoid and GRIP-1.

DSF was used to examine the dependence of Tm on v (Figure 7). In the absence of rexinoids, GRIP-1 increased the Tm of apo-RXRα LBD by ~1 °C, consistent with its low affinity to the apo-homodimer at 25 °C.9 A much larger increase in Tm was observed in the presence of each rexinoid. All Tm values appeared linearly dependent on v−1, and the slopes of the linear fits were similar to those obtained in the absence of GRIP-1. This again suggested that the rates of the irreversible step of the Lumry–Eyring unfolding mechanism were similar in the presence or absence of rexinoids or GRIP-1. The RXRα LBD:UAB111:GRIP-1 complex displayed the highest extrapolated Tm of 68.7 ± 0.1 °C, nearly 11 °C higher than that of the apo-homodimer. The RXRα LBD:UAB110:GRIP-1 complex had an extrapolated Tm of 67.9 ± 0.2 °C, and the RXRα LBD:UAB30:GRIP-1 complex had an extrapolated Tm of 66.4 ± 0.1 °C. Using DSC (Figure 8 and Table 2), similar increases in Tm were observed for the ternary complexes.

The ΔCpu values of holo-RXRα LBD homodimers in complex with GRIP-1 were determined from the linear dependence of ΔHc on Tm and listed in Table 2. The unfolding parameters of apo-RXRα LBD homodimer in the presence of 400 μM GRIP-1 were similar to the apo-homodimer with only slightly increased Tm and ΔHc, and almost identical ΔCpu. This indicated that significant dissociation of GRIP-1 likely occurred before the apo-homodimer unfolded because of low binding affinity at Tm. For two of the three RXRα LBD:rexinoid:GRIP-1 complexes, ΔCpu increased with respect to their corresponding holo-homodimer without GRIP-1, while there was little observed change in ΔCpu for RXRα LBD:UAB30:GRIP-1. The largest increase in ΔCpu was 3 kJ/(mol K) for RXRα LBD:UAB111:GRIP-1, which was within the experimental errors of the ΔCpu measurement. Therefore, an average ΔCpu value of 18 ± 2 kJ/(mol K) was calculated from the experimentally determined values for the three ternary complexes.

Using the above average ΔCpu value and the Tm and ΔHc values from DSC measurements at 4 °C/min, ΔH, ΔS, and ΔG of unfolding at 37 °C for the holo-RXRα LBD homodimers bound with GRIP-1 peptide were calculated (Table 2). Binding of GRIP-1 increased ΔG compared to measurements obtained in the absence of the coactivator peptide: 13 ± 4 kJ/mol for RXRα LBD:UAB30, 14 ± 3 kJ/mol for RXRα LBD:UAB110, and 13 ± 3 kJ/mol for RXRα LBD:UAB111. Using eq 5 in Appendix I (Supporting Information) based on the shifts in Tm, the estimated Kd of GRIP-1 binding at Tm was 22 μM for RXRα LBD:UAB30, 26 μM for RXRα LBD:UAB110, and 24 μM for RXRα LBD:UAB111.

In our previous analysis using ITC,8 the binding enthalpy of GRIP-1 to RXRα LBD:UAB30 was determined to be −44.6 ±0.1 kJ/mol per monomer at 30 °C, with a binding heat capacity change (ΔCpa) of −1.5 ± 0.1 kJ/(mol K). Using these values, binding enthalpy at 67.5 °C (Tm of the RXRα LBD:UAB30:-GRIP-1 ternary complex) was calculated to be −100 kJ/mol per monomer or −200 kJ/mol per dimer, which was similar to the measured value of −129 ± 35 kJ/mol. Based on the van’t Hoff equation (eq 6 in Appendix I of Supporting Information), and a Kd of 26 μM at Tm, the Kd at 25 °C was calculated to be 0.74 μM, in close agreement with the Kd determined by ITC at this temperature.8

To obtain an accurate measurement of the ΔCpa values for GRIP-1 binding to RXRα LBD:UAB110 or RXRα LBD:UAB111, ITC was performed between 10 and 30 °C (Table S3). As observed for the RXRα LBD:UAB30 complex, the binding stoichiometry was nearly 1:1 (coactivator peptide/RXRα LBD monomer unit). The free energy change of GRIP-1 binding to RXRα LBD:rexinoid complexes was driven strongly by a large negative enthalpy change. The ΔCpa was determined to be 1.19 ± 0.01 kJ/(mol K) for GRIP-1 binding to RXRα LBD:UAB110, and 1.11 ± 0.07 kJ/(mol K) for GRIP-1 binding to RXRα LBD:UAB111 (per monomer unit). The Kd at 25 °C was calculated to be 1.5 μM for RXRα LBD:UAB110 and 1.8 μM for RXRα LBD:UAB110, also in close agreement with the Kd determined by ITC at this temperature.14

In summary, we report here the thermal unfolding of three species of RXRα LBD homodimers: (1) the apo-homodimer alone, N2, which reversibly unfolded to a dimeric intermediate, I2, that irreversibly denatured to D with a rate constant, k; (2) the holo-homodimer bound with a rexinoid to each monomer, N2R2; and (3) the ternary complex of the homodimer bound with a rexinoid and a coactivator peptide GRIP-1 to each monomer, N2R2P2.

N2I2DΔG(37°C)=33kJ/mol;k(I2D)=2.1min1 (2)
N2R2N2+2RΔΔG(37°C)=1220kJ/mol(rexinoid-dependent) (3)
N2R2P2N2R2+2PΔΔG(37°C)=14kJ/mol(rexinoid-independent) (4)

From the analyses of DSC endotherms of N2 at different scan rates, we estimated that N2 was 33 kJ/mol more stable than I2 at 37 °C. Relative to this equilibrium, the analyses of DSC endotherms of N2R2 bound with three different UAB rexinoids indicated that N2 was stabilized by an additional 12–20 kJ/mol, depending on the structure of the rexinoid. This stabilization was consistent with our reported 9–35 nM binding affinity of UAB rexinoids determined by fluorescence quenching assays.14 The ΔΔG of stabilization caused by each rexinoid positively correlated with their binding affinities, with UAB111 being the tightest binder. The DSC analyses of N2R2P2 revealed that GRIP-1 binding further stabilized the complex by an additional 14 kJ/mol at 37 °C. Unlike rexinoid binding, this additional stabilization was nearly identical for the homodimers bound with three different rexinoids, which was in close agreement with the Kd values of GRIP-1 release obtained by ITC.14 Even though GRIP-1 binding stabilized N2R2 the same, N2R2P2 bound with UAB110 or UAB111 was 8 kJ/mol more stable than this complex with UAB30.

The equilibrium ΔΔG of binding was determined in this study even with the apo-homodimer unfolding irreversibly. Rexinoid binding to each monomer of N2 shifted the population of N2 to N2R2, causing the equilibrium N2 ↔ I2 to require higher temperature to reach 50% unfolding. GRIP-1 binding further shifted the population to N2R2P2, thus requiring even higher temperature to produce unfolding. As long as the rexinoid/coactivator binding processes are in equilibrium, the thermodynamic parameters of binding determined from ΔΔG (Tm) and extrapolated to 37 °C using measured ΔCp will be reasonable. These measurements could be compromised if the ligand-bound species, N2R2 and/or N2R2P2, irreversibly unfold to new denatured species with different rates. This does not appear to be the case for this study since, as shown in Figure 7, the Tm vs v−1 plots display linear relationships with nearly identical slopes, suggesting that the rate constant and the activation energy for the irreversible step are similar for each native species.

CONCLUSIONS

In this study we demonstrated that apo-RXRα LBD homodimer is favored in the folded state by at least 33 kJ/mol in free energy change at 37 °C, driven by a favorable enthalpy change. On a per residue basis, the enthalpy value is almost 50% less than other small globular proteins, most likely due to two factors: (1) thermal unfolding which produced an incompletely unfolded state and (2) the less compact nature of the apo-homodimer that contains a large unfilled hydrophobic rexinoid binding pocket and a dynamic c-terminal end. Upon rexinoid binding, the hydrophobic pocket is filled, stabilized by numerous hydrophobic and hydrophilic interactions between the rexinoids and binding pocket residues. These interactions also help bridge the helices interacting with the rexinoids to those at the terminal end where the coactivator peptide binds. We clearly demonstrate here for the first time that rexinoid binding to apo-RXRα LBD is entropically driven at physiological temperatures with negligible enthalpy change. The rexinoids enhanced the stability of the homodimer complex by 12–20 kJ/mol, depending on their structure. The ΔΔG of stabilization is 8 kJ/mol less for UAB30 than the larger and more hydrophobic UAB110 or UAB111. The binding of coactivator peptide results in an additional stabilization to the homodimer complex of approximately 14 kJ/mol. The incremental increase in free energy is the same for each of the three rexinoids studied. Our data on GRIP-1 binding to apo-RXRα LBD also indicate that a bound rexinoid is required for coactivator peptide binding. Structurally, the homodimer complexes with GRIP-1 are essentially identical. It seems as if GRIP-1 does not sense the subtle differences in the LBP conformation caused by different rexinoids as long as the LBP is occupied by an agonist of sufficient size and hydrophobicity to promote GRIP-1 binding to its site on the surface of the domain. These results suggest that, for future studies on RXR heterodimers, we would want to investigate RXR heterodimer interactions in addition to differential coactivator affinity. Alternatively, we may find that different RXR coactivator LXXLL motifs have unique binding characteristics with unique RXR ligands.

Finally, even though the GRIP-1 binding thermodynamics are the same for each of the three holo-homodimers, the complexes of UAB110 and UAB111 are each more stable than the UAB30 complex (8 kJ/mol) not due to enhanced coactivator interactions but due to hydrophobic interactions of the rexinoid in its interior binding pocket. The enhanced stability of UAB110 and UAB111 correlates well with in vitro transcriptional assays14 that clearly demonstrate that UAB111 is one of the most potent RXRα agonists discovered to date (40-fold more potent than UAB30). As reported here, despite the essentially identical X-ray crystal structures of the homodimer complexes, the two homodimer complexes with UAB110 or UAB111 are clearly more stable than the complex with UAB30. This demonstrates the value in thermodynamic measurements in predicting agonist potency. The gathering of this information was made more complicated by an irreversible step in the thermal unfolding of apo-RXRα LBD dimers. The comprehensive analyses presented here not only yield specific thermodynamic information on the rexinoids studied but also provide a general method of how to deal with this complication. These studies lay the groundwork for future thermodynamic studies on RXR agonists and antagonists, as well as RXR heterodimers and other coactivators.

Supplementary Material

PMID33792309 supplementary

ACKNOWLEDGMENTS

The authors would like to thank Dr. Christie Brouillete for her helpful comments on the manuscript. Access to the VP-Capillary DSC, Auto-iTC200, and Prometheus NT.48 instruments was provided by the Biocalorimetry Lab supported by the NIH Shared Instrumentation Grant 1S10RR026478 and Shared Facility Program of the UAB Comprehensive Cancer Center, Grant 316851.

Funding

NIH Grant P01 CA210946

ABBREVIATIONS

RXRα LBD

retinoid X receptor-alpha ligand binding domain

LBP

ligand binding pocket

9cRA

9-cis retinoic acid

GRIP-1

glucocorticoid receptor interacting protein-1

RXRα LBD:UAB30

RXRα LBD bound to UAB30

RXRα LBD:UAB30:GRIP-1

RXRα LBD bound to UAB30 and GRIP-1

RXRα LBD:UAB110

RXRα LBD bound to UAB110

RXRα LBD:UAB110:GRIP-1

RXRα LBD bound to UAB110 and GRIP-1

RXRα LBD:UAB111

RXRα LBD bound to UAB111

RXRα LBD:UAB111:GRIP-1

RXRα LBD bound to UAB111 and GRIP-1

SRC

steroid receptor coactivator

DSC

differential scanning calorimetry

v

scan or heating rate

T m

thermal unfolding temperature

ΔHv

van’t Hoff enthalpy of unfolding

ΔHc

calorimetric enthalpy

ΔG

Gibbs free energy of unfolding

ΔS

entropy of unfolding

ΔCpu

unfolding heat capacity change

K u

unfolding equilibrium constant

DSF

differential scanning fluorimetry

F350/F330

ratio of fluorescence intensity at 350 nm divided by fluorescence intensity at 330 nm

λ ex

excitation wavelength for fluorescence measurements

CD

circular dichroism

ITC

isothermal titration calorimetry

ΔHa

binding enthalpy

ΔCpa

binding heat capacity change

K a

ligand binding constant

K d

ligand dissociation constant

HDX-MS

hydrogen–deuterium exchange mass spectrometry

N

native folded protein

U

reversibly unfolded protein

D

irreversibly denatured protein

I

partially unfolded intermediate

k

rate constant of irreversible step

R

rexinoid

P

coactivator peptide GRIP-1

N2

native folded apo-RXRα LBD homodimer protein

I2

partially unfolded apo-RXRα LBD dimeric intermediate

N2R2

native folded holo-RXRα LBD homodimer bound with rexinoids

N2R2P2

native folded holo-RXRα LBD homodimer bound with rexinoids and GRIP-1

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biochem.0c00865.

Size exclusion chromatography and SDS-PAGE, typical DSC endotherm and the cubic baseline, two-state fits of the DSC endotherms and extrapolation to infinite scan rate, partial reversibility in DSC, scan-rate dependence of DSF unfolding transitions of apo-RXRα LBD, simulated DSC unfolding transitions of a hypothetical dimeric protein using two-state models with or without dimer dissociation, pH dependence of DSF unfolding transitions, DSC of apo-RXRα LBD at pH 7.0 in different buffers, effect of rate constant of the irreversible step on DSC Tm and ΔH, determination of saturating rexinoid concentrations at Tm, nonspecific protein destabilization by UAB110 and UAB111 at high concentrations, equilibrium unfolding parameters of RXRα LBD:UAB30 obtained by extrapolation to infinite scan rate, DSC parameters of RXRα LBD:UAB30 at different scan rates and extrapolated to infinite scan rate, DSF Tm of RXRα LBD bound with UAB30 at different pH values, DSF Tm-shifts of holo-RXRα LBD that do not vary with pH, DSC of apo-RXRα LBD and holo-RXRα LBD with and without GRIP-1 at pH 8.8 and pH 9.5, ITC of apo-RXRα LBD and UAB30 at different temperatures, DSF Tm of UAB30:RXRα LBD as a function of GRIP-1 concentration, summary of ITC measurements of GRIP-1 to RXRα LBD:rexinoid complexes (Table S3), and equations for the calculation of thermodynamic parameters (Appendix I) (PDF)

Accession Codes

The UniProt accession code for the retinoid X receptor-alpha is P19793.

Complete contact information is available at: https://pubs.acs.org/10.1021/acs.biochem.0c00865

The authors declare no competing financial interest.

Contributor Information

Zhengrong Yang, Department of Biochemistry & Molecular Genetics, School of Medicine, University of Alabama at Birmingham, Birmingham, Alabama 35294, United States;.

Donald D. Muccio, Department of Biochemistry & Molecular Genetics, School of Medicine, University of Alabama at Birmingham, Birmingham, Alabama 35294, United States

Nathalia Melo, Department of Biochemistry & Molecular Genetics, School of Medicine, University of Alabama at Birmingham, Birmingham, Alabama 35294, United States.

Venkatram R. Atigadda, Department of Dermatology, School of Medicine, University of Alabama at Birmingham, Birmingham, Alabama 35294, United States

Matthew B. Renfrow, Department of Biochemistry & Molecular Genetics, School of Medicine, University of Alabama at Birmingham, Birmingham, Alabama 35294, United States;

REFERENCES

  • (1).Dawson MI, and Xia Z (2012) The retinoid X receptors and their ligands. Biochim. Biophys. Acta, Mol. Cell Biol. Lipids 1821, 21–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (2).Mangelsdorf DJ, Thummel C, Beato M, Herrlich P, Schutz G, Umesono K, Blumberg B, Kastner P, Mark M, Chambon P, and Evans RM (1995) The nuclear receptor superfamily: the second decade. Cell 83, 835–839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Zhao Q, Chasse SA, Devarakonda S, Sierk ML, Ahvazi B, and Rastinejad F (2000) Structural basis of RXR-DNA interactions. J. Mol. Biol 296, 509–520. [DOI] [PubMed] [Google Scholar]
  • (4).Bourguet W, Ruff M, Chambon P, Gronemeyer H, and Moras D (1995) Crystal structure of the ligand-binding domain of the human nuclear receptor RXR-alpha. Nature 375, 377–382. [DOI] [PubMed] [Google Scholar]
  • (5).Egea PF, Mitschler A, Rochel N, Ruff M, Chambon P, and Moras D (2000) Crystal structure of the human RXRalpha ligand-binding domain bound to its natural ligand: 9-cis retinoic acid. EMBOJ. 19, 2592–2601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).Egea PF, Mitschler A, and Moras D (2002) Molecular recognition of agonist ligands by RXRs. Mol. Endocrinol 16, 987–997. [DOI] [PubMed] [Google Scholar]
  • (7).Eberhardt J, McEwen AG, Bourguet W, Moras D, and Dejaegere A (2019) A revisited version of the apo structure of the ligand-binding domain of the human nuclear receptor retinoic X receptor alpha. Acta Crystallogr., Sect. F: Struct. Biol. Commun 75, 98–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Boerma LJ, Xia G, Qui C, Cox BD, Chalmers MJ, Smith CD, Lobo-Ruppert S, Griffin PR, Muccio DD, and Renfrow MB (2014) Defining the communication between agonist and coactivator binding in the retinoid X receptor alpha ligand binding domain. J. Biol. Chem 289, 814–826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).Xia G, Boerma LJ, Cox BD, Qiu C, Kang S, Smith CD, Renfrow MB, and Muccio DD (2011) Structure, energetics, and dynamics of binding coactivator peptide to the human retinoid X receptor alpha ligand binding domain complex with 9-cis-retinoic acid. Biochemistry 50, 93–105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (10).Nahoum V, Perez E, Germain P, Rodriguez-Barrios F, Manzo F, Kammerer S, Lemaire G, Hirsch O, Royer CA, Gronemeyer H, de Lera AR, and Bourguet W (2007) Modulators of the structural dynamics of the retinoid X receptor to reveal receptor function. Proc. Natl. Acad. Sci. U. S. A 104, 17323–17328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (11).Zhang H, Li L, Chen L, Hu L, Jiang H, and Shen X (2011) Structure basis of bigelovin as a selective RXR agonist with a distinct binding mode. J. Mol. Biol 407, 13–20. [DOI] [PubMed] [Google Scholar]
  • (12).Lippert WP, Burschka C, Gotz K, Kaupp M, Ivanova D, Gaudon C, Sato Y, Antony P, Rochel N, Moras D, Gronemeyer H, and Tacke R (2009) Silicon analogues of the RXR-selective retinoid agonist SR11237 (BMS649): chemistry and biology. ChemMedChem 4, 1143–1152. [DOI] [PubMed] [Google Scholar]
  • (13).Atigadda VR, Xia G, Desphande A, Boerma LJ, Lobo-Ruppert S, Grubbs CJ, Smith CD, Brouillette WJ, and Muccio DD (2014) Methyl substitution of a rexinoid agonist improves potency and reveals site of lipid toxicity. J. Med. Chem 57, 5370–5380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Atigadda VR, Xia G, Deshpande A, Wu L, Kedishvili N, Smith CD, Krontiras H, Bland KI, Grubbs CJ, Brouillette WJ, and Muccio DD (2015) Conformationally Defined Rexinoids and Their Efficacy in the Prevention of Mammary Cancers. J. Med. Chem 58, 7763–7774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Heery DM, Kalkhoven E, Hoare S, and Parker MG (1997) A signature motif in transcriptional co-activators mediates binding to nuclear receptors. Nature 387 (6634), 733–736. [DOI] [PubMed] [Google Scholar]
  • (16).Yan X, Broderick D, Leid ME, Schimerlik MI, and Deinzer ML (2004) Dynamics and ligand-induced solvent accessibility changes in human retinoid X receptor homodimer determined by hydrogen deuterium exchange and mass spectrometry. Biochemistry 43, 909–917. [DOI] [PubMed] [Google Scholar]
  • (17).Chalmers MJ, Busby SA, Pascal BD, He Y, Hendrickson CL, Marshall AG, and Griffin PR (2006) Probing protein ligand interactions by automated hydrogen/deuterium exchange mass spectrometry. Anal. Chem 78, 1005–1014. [DOI] [PubMed] [Google Scholar]
  • (18).Lu J, Cistola DP, and Li E (2006) Analysis of ligand binding and protein dynamics of human retinoid X receptor alpha ligand-binding domain by nuclear magnetic resonance. Biochemistry 45, 1629–1639. [DOI] [PubMed] [Google Scholar]
  • (19).Chandra V, Huang P, Hamuro Y, Raghuram S, Wang Y, Burris TP, and Rastinejad F (2008) Structure of the intact PPAR-gamma-RXR- nuclear receptor complex on DNA. Nature 456, 350–356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Chandra V, Wu D, Li S, Potluri N, Kim Y, and Rastinejad F (2017) The quaternary architecture of RARbeta-RXRalpha heterodimer facilitates domain-domain signal transmission. Nat. Commun 8, 868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (21).Zhang J, Chalmers MJ, Stayrook KR, Burris LL, Garcia-Ordonez RD, Pascal BD, Burris TP, Dodge JA, and Griffin PR (2010) Hydrogen/deuterium exchange reveals distinct agonist/partial agonist receptor dynamics within vitamin D receptor/retinoid X receptor heterodimer. Structure 18, 1332–1341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Lou X, Toresson G, Benod C, Suh JH, Philips KJ, Webb P, and Gustafsson JA (2014) Structure of the retinoid X receptor alpha-liver X receptor beta (RXRalpha-LXRbeta) heterodimer on DNA. Nat. Struct. Mol. Biol 21, 277–281. [DOI] [PubMed] [Google Scholar]
  • (23).Kojetin DJ, Matta-Camacho E, Hughes TS, Srinivasan S, Nwachukwu JC, Cavett V, Nowak J, Chalmers MJ, Marciano DP, Kamenecka TM, Shulman AL, Rance M, Griffin PR, Bruning JB, and Nettles KW (2015) Structural mechanism for signal transduction in RXR nuclear receptor heterodimers. Nat. Commun 6, 8013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (24).Grubbs CJ, Lubet RA, Atigadda VR, Christov K, Deshpande AM, Tirmal V, Xia G, Bland KI, Eto I, Brouillette WJ, and Muccio DD (2006) Efficacy of new retinoids in the prevention of mammary cancers and correlations with short-term biomarkers. Carcinogenesis 27, 1232–1239. [DOI] [PubMed] [Google Scholar]
  • (25).Muccio DD, Atigadda VR, Brouillette WJ, Bland KI, Krontiras H, and Grubbs CJ (2017) Translation of a Tissue-Selective Rexinoid, UAB30, to the Clinic for Breast Cancer Prevention. Curr. Top. Med. Chem 17, 676–695. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (26).Garner EF, Stafman LL, Williams AP, Aye JM, Goolsby C, Atigadda VR, Moore BP, Nan L, Stewart JE, Hjelmeland AB, Friedman GK, and Beierle EA (2018) UAB30, a novel RXR agonist, decreases tumorigenesis and leptomeningeal disease in group 3 medulloblastoma patient-derived xenografts. J. Neuro-Oncol 140, 209–224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).Chou CF, Hsieh YH, Grubbs CJ, Atigadda VR, Mobley JA, Dummer R, Muccio DD, Eto I, Elmets CA, Garvet WT, and Chang PL (2018) The retinoid X receptor agonist, 9-cis UAB30, inhibits cutaneous T-cell lymphoma proliferation through the SKP2-p27kip1 axis. J. Dermatol. Sci 90, 343–356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (28).Gniadecki R, Assaf C, Bagot M, Dummer R, Duvic M, Knobler R, Ranki A, Schwandt P, and Whittaker S (2007) The optimal use of bexarotene in cutaneous T-cell lymphoma. Br. J. Dermatol 157, 433–440. [DOI] [PubMed] [Google Scholar]
  • (29).Muccio DD, Brouillette WJ, Breitman TR, Taimi M, Emanuel PD, Zhang X, Chen G, Sani BP, Venepally P, Reddy L, Alam M, Simpson-Herren L, and Hill DL (1998) Conformationally defined retinoic acid analogues. 4. Potential new agents for acute promyelocytic and juvenile myelomonocytic leukemias. J. Med. Chem 41, 1679–1687. [DOI] [PubMed] [Google Scholar]
  • (30).Atigadda VR, Vines KK, Grubbs CJ, Hill DL, Beenken SL, Bland KI, Brouillette WJ, and Muccio DD (2003) Conformationally defined retinoic acid analogues. 5. Large-scale synthesis and mammary cancer chemopreventive activity for (2E,4E,6Z,8E)-8-(3′,4′-dihydro-1′(2′H)-naphthalen-1′-ylidene)-3,7-dimethyl-2,4,6- octatrienoic acid (9cUAB30). J. Med. Chem 46, 3766–3769. [DOI] [PubMed] [Google Scholar]
  • (31).Kolesar JM, Hoel R, Pomplun M, Havighurst T, Stublaski J, Wollmer B, Krontiras H, Brouillette WJ, Muccio DD, Kim K, Grubbs CJ, and Bailey HE (2010) A pilot, first-in-human, pharmacokinetic study of 9cUAB30 in healthy volunteers. Cancer Prev. Res 3, 1565–1570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (32).Kolesar JM, Andrews S, Green H, Havighurst TC, Wollmer BW, DeShong K, Laux DE, Krontiras H, Muccio DD, Kim K, Grubbs CJ, House MG, Parnes HL, Heckman-Stoddard BM, and Bailey HH (2019) A Randomized, Placebo-Controlled, Double-Blind, Dose Escalation, Single Dose, and Steady-State Pharmacokinetic Study of 9cUAB30 in Healthy Volunteers. Cancer Prev. Res 12, 903–912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (33).Desphande A, Xia G, Boerma LJ, Vines KK, Atigadda VR, Lobo-Ruppert S, Grubbs CJ, Moeinpour FL, Smith CD, Christov K, Brouillette WJ, and Muccio DD (2014) Methyl-substituted conformationally constrained rexinoid agonists for the retinoid X receptors demonstrate improved efficacy for cancer therapy and prevention. Bioorg. Med. Chem 22, 178–185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (34).Doyle ML (1997) Characterization of binding interactions by isothermal titration calorimetry. Curr. Opin. Biotechnol 8, 31–35. [DOI] [PubMed] [Google Scholar]
  • (35).Ward WH, and Holdgate GA (2001) Isothermal titration calorimetry in drug discovery. Prog. Med. Chem 38, 309–376. [DOI] [PubMed] [Google Scholar]
  • (36).Brandts JF, and Lin LN (1990) Study of strong to ultratight protein interactions using differential scanning calorimetry. Biochemistry 29, 6927–6940. [DOI] [PubMed] [Google Scholar]
  • (37).Waldron TT, and Murphy KP (2003) Stabilization of proteins by ligand binding: application to drug screening and determination of unfolding energetics. Biochemistry 42, 5058–5064. [DOI] [PubMed] [Google Scholar]
  • (38).Schimerlik MI, Peterson VJ, Hobbs PD, Dawson MI, and Leid M (1999) Kinetic and thermodynamic analysis of 9-cis-retinoic acid binding to retinoid X receptor alpha. Biochemistry 38, 6732–6740. [DOI] [PubMed] [Google Scholar]
  • (39).Yang Z, and Brouillette CG (2016) A Guide to Differential Scanning Calorimetry of Membrane and Soluble Proteins in Detergents. Methods Enzymol. 567, 319–358. [DOI] [PubMed] [Google Scholar]
  • (40).Protasevich IR, Yang Z, Wang C, Atwell S, Zhao X, Emtage S, Wetmore D, Hunt JF, and Brouillette CG (2010) Thermal unfolding studies show the disease causing F508del mutation in CFTR thermodynamically destabilizes nucleotide-binding domain1. Pro. Sci 19, 1917–1931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (41).Harder ME, Malencik DA, Yan X, Broderick D, Leid M, Anderson SR, Deinzer ML, and Schimerlik MI (2009) Equilibrium unfolding of the retinoid X receptor ligand binding domain and characterization of an unfolding intermediate. Biophys. Chem 141, 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (42).Privalov PL (1979) Stability of proteins: small globular proteins. Adv. Protein Chem 33, 167–241. [DOI] [PubMed] [Google Scholar]
  • (43).Garidel P, Hegyi M, Bassarab S, and Weichel M (2008) A rapid, sensitive and economical assessment of monoclonal antibody conformational stability by intrinsic tryptophan fluorescence spectroscopy. Biotechnol. J 3, 1201–1211. [DOI] [PubMed] [Google Scholar]
  • (44).Zelent B, Sharp KA, and Vanderkooi JM (2010) Differential scanning calorimetry and fluorescence study of lactoper-oxidase as a function of guanidinium-HCl, urea, and pH. Biochim. Biophys. Acta, Proteins Proteomics 1804, 1508–1515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (45).Yin SW, Tang CH, Yang XQ, and Wen QB (2011) Conformational study of red kidney bean (Phaseolus vulgaris L.) protein isolate (KPI) by tryptophan fluorescence and differential scanning calorimetry. J. Agric. Food Chem 59, 241–248. [DOI] [PubMed] [Google Scholar]
  • (46).Thiagarajan G, Semple A, James JK, Cheung JK, and Shameem M (2016) A comparison of biophysical characterization techniques in predicting monoclonal antibody stability. MAbs 8, 1088–1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (47).Freire E, van Osdol WW, Mayorga OL, and Sanchez-Ruiz JM (1990) Calorimetrically determined dynamics of complex unfolding transitions in proteins. Annu. Rev. Biophys. Biophys. Chem 19, 159–188. [DOI] [PubMed] [Google Scholar]
  • (48).Senisterra GA, Markin E, Yamazaki K, Hui R, Vedadi M, and Awrey DE (2006) Screening for ligands using a generic and high-throughput light-scattering-based assay. J. Biomol. Screening 11, 940–948. [DOI] [PubMed] [Google Scholar]
  • (49).Lumry R, and Eyring H (1954) Conformation Changes of Proteins. J. Phys. Chem 58, 110–120. [Google Scholar]
  • (50).Sanchez-Ruiz JM (1992) Theoretical analysis of Lumry-Eyring models in differential scanning calorimetry. Biophys. J 61, 921–935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (51).Lepock JR, Ritchie KP, Kolios MC, Rodahl AM, Heinz KA, and Kruuv J (1992) Influence of transition rates and scan rate on kinetic simulations of differential scanning calorimetry profiles of reversible and irreversible protein denaturation. Biochemistry 31, 12706–12712. [DOI] [PubMed] [Google Scholar]
  • (52).Tello-Solis SR, and Hernandez-Arana A (1995) Effect of irreversibility on the thermodynamic characterization of the thermal denaturation of Aspergillus saitoi acid proteinase. Biochem. J 311, 969–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (53).Vogl T, Jatzke C, and Hinz H (1997) Thermodynamic stability of annexin v e17g: equilibrium parameters from an irreversible unfolding reaction. Biochemistry 36, 1657–1668. [DOI] [PubMed] [Google Scholar]
  • (54).Yang ZW, Tendian SW, Carson WM, Brouillette WJ, Delucas LJ, and Brouillette CG (2004) Dimethyl sulfoxide at 2.5% (v/v) alters the structural cooperativity and unfolding mechanism of dimeric bacterial NAD+ synthetase. Protein Sci. 13, 830–841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (55).Murphy KP, and Freire E (1992) Thermodynamics of structural stability and cooperative folding behavior in proteins. Adv. Protein Chem 43, 313–361. [DOI] [PubMed] [Google Scholar]
  • (56).Nevzorov I, Redwood C, and Levitsky D (2008) Stability of two beta-tropomyosin isoforms: effects of mutation Arg91Gly. J. Muscle Res. Cell Motil 29, 173–176. [DOI] [PubMed] [Google Scholar]
  • (57).Yang Z, Zhou Q, Mok L, Singh A, Swartz DJ, Urbatsch IL, and Brouillette CG (2017) Interactions and cooperativity between P-glycoprotein structural domains determined by thermal unfolding provides insights into its solution structure and function. Biochim. Biophys. Acta, Biomembr 1859, 48–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Robbins RJ, Fleming GR, Beddard GS, Robinson GW, Thistlethwaite PJ, and Woolfe GJ (1980) Photophysics of aqueous tryptophan: pH and temperature effects. J. Am. Chem. Soc 102, 6271–6279. [Google Scholar]
  • (59).Watlaufer DB, Malik SK, Stoller L, and Coffin RL (1964) Nonpolar Group Participation in the Denaturation of Proteins by Urea and Guanidinium Salts. Model Compound Studies. J. Am. Chem. Soc 86, 508–514. [Google Scholar]
  • (60).Nozaki Y, and Tanford C (1970) The solubility of amino acids, diglycine, and triglycine in aqueous guanidine hydrochloride solutions. J. Biol. Chem 245, 1648–1652. [PubMed] [Google Scholar]
  • (61).Venkatesu P, Lee MJ, and Lin HM (2007) Thermodynamic characterization of the osmolyte effect on protein stability and the effect of GdnHCl on the protein denatured state. J. Phys. Chem. B 111, 9045–9056. [DOI] [PubMed] [Google Scholar]
  • (62).Heery DM, Pierrat B, Gronemeyer H, Chambon P, and Losson R (1994) Homo- and heterodimers of the retinoid X receptor (RXR) activated transcription in yeast. Nucleic Acids Res. 22, 726–731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (63).Privalov PL (1982) Stability of proteins. Proteins which do not present a single cooperative system. Adv. Protein Chem 35, 1–104. [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

PMID33792309 supplementary

RESOURCES