Summary

Measuring Biomolecular DSC Profiles with Thermolabile Ligands to Rapidly Characterize Folding and Binding Interactions

Published: November 21, 2017
doi:

Summary

We present a protocol for rapid characterization of biomolecular folding and binding interactions with thermolabile ligands using differential scanning calorimetry.

Abstract

Differential scanning calorimetry (DSC) is a powerful technique for quantifying thermodynamic parameters governing biomolecular folding and binding interactions. This information is critical in the design of new pharmaceutical compounds. However, many pharmaceutically relevant ligands are chemically unstable at the high temperatures used in DSC analyses. Thus, measuring binding interactions is challenging because the concentrations of ligands and thermally-converted products are constantly changing within the calorimeter cell. Here, we present a protocol using thermolabile ligands and DSC for rapidly obtaining thermodynamic and kinetic information on the folding, binding, and ligand conversion processes. We have applied our method to the DNA aptamer MN4 that binds to the thermolabile ligand cocaine. Using a new global fitting analysis that accounts for thermolabile ligand conversion, the complete set of folding and binding parameters are obtained from a pair of DSC experiments. In addition, we show that the rate constant for thermolabile ligand conversion may be obtained with only one supplementary DSC dataset. The guidelines for identifying and analyzing data from several more complicated scenarios are presented, including irreversible aggregation of the biomolecule, slow folding, slow binding, and rapid depletion of the thermolabile ligand.

Introduction

Differential scanning calorimetry (DSC) is a powerful method for quantitating biomolecular binding and folding interactions1,2,3. The strengths of DSC include its ability to elucidate binding and folding mechanisms, and to yield the corresponding thermodynamic parameters2,3. Furthermore, DSC can be performed in solution under near-physiological conditions and does not require labeling of the biomolecule or ligand, e.g., with fluorophores, spin-labels or nuclear isotopes4. The instrument scans in temperature, measuring the amount of heat required to denature the biomolecule in the presence and absence of ligand. The resulting thermograms are used to extract the thermodynamic parameters governing the ligand binding and folding processes. The information provided by DSC or other thermodynamic techniques is critical to guiding the design of drugs targeting biomolecules1,5,6,7,8. However, the repeated scanning to high temperatures (~ 60 – 100 °C) can be problematic. For example, many pharmaceutically important compounds undergo rearrangement or decomposition upon sustained exposure to high temperatures9,10,11, i.e., they are thermolabile. Examination of binding interactions by DSC typically requires multiple forward and reverse scans in order to verify the reproducibility of the thermogram for thermodynamic analyses12. Thermal conversion of an initial ligand to a secondary form with altered binding characteristics leads to pronounced differences in the shape and position of successive thermograms, since the concentration of the initial ligand decreases with each scan while the thermal conversion products accumulate. These datasets are not amenable to traditional analyses.

We have recently developed a global fitting method for thermolabile ligand DSC datasets that yields the complete set of thermodynamic parameters governing the biomolecular folding and binding interactions from a single ligand-bound experiment referenced to the requisite thermogram for the free biomolecule4. The analysis reduces the experimental time and sample required by ~ 10-fold compared to standard DSC approaches. We have accounted for ligand thermal conversion by assuming this happens during the high temperature portion of each scan where the thermogram does not depend on ligand concentration. Therefore, the ligand concentration is a constant within the portion of the thermogram that is used to extract thermodynamic parameters. We additionally demonstrated how the rate constant for ligand thermal conversion can be obtained by performing one supplementary experiment with a longer high temperature equilibration period. For systems where ligand thermal conversion is less temperature-dependent (i.e., occurring appreciably at all temperatures), the analysis can be modified to include variable ligand concentrations. Here we demonstrate this procedure for the DNA aptamer MN4 in the presence of the thermolabile ligand cocaine, which rapidly converts to benzoylecgonine at high temperatures (>60 °C). Quinine is used as a negative control for ligand thermolability since it does not undergo conversion at these experimental temperatures and also binds to MN4. We describe the acquisition of thermolabile ligand DSC datasets and their analysis yielding thermodynamic and kinetic parameters of the folding, binding, and ligand conversion processes.

Protocol

1. Sample Preparation Purify the desired biomolecule13. NOTE: This protocol uses purchased cocaine-binding DNA aptamer MN4 after exchanging against 2 M NaCl three times followed by three rounds of deionized water using a centrifugal filter with a 3 kDa molecular weight cut-off membrane. Synthesize and purify, or purchase the desired thermolabile ligand13. NOTE: MN4 binds the thermolabile ligand cocaine. MN4 also binds quinine, which is us…

Representative Results

Representative data for the thermolabile ligand DSC are shown in Figure 1. The position and height of the thermolabile ligand-bound peak successively shifts down towards that of the unbound biomolecule as the thermolabile ligand is depleted with each scan (Figure 1a). The free denaturation profile is used as a reference for the endpoint of thermolabile ligand conversion (Figure 1b). Data for MN4 boun…

Discussion

Modifications and troubleshooting

The details of the global fitting analysis used in Figure 1 and Figure 2 have been described previously4. Here, we outline practical aspects of performing and analyzing DSC binding experiments with thermolabile ligands. Note that a DSC baseline obtained for the thermolabile ligand alone is subtracted from the ligand + biomolecule dataset, effectively cancelling ou…

Disclosures

The authors have nothing to disclose.

Acknowledgements

R. W. H. V was supported by the McGill Natural Sciences and Engineering Research Council of Canada (NSERC) Training Program in Bionanomachines. A. K. M. and P. E. J. were supported by NSERC grants 327028-09 (A. K. M) and 238562 (P. E. J.).

Materials

Sodium chloride Chem Impex #00829
Sodium phosphate monobasic dihydrate Sigma Aldrich 71502
Sodium phosphate dibasic Sigma Aldrich S9763
Deioinized water for molecular biology Millipore H20MB1001
0.2 micron sterile syringe filters VWR CA28145-477
3 kDa centrifugal filters Millipore UFC900324
Dialysis tubing 0.5-1.0 kDa cutoff Spectrum Laboratories 131048
Silicon tubing VWR 89068-474
Plastic DSC flange caps TA Instruments 6111
DNA aptamer MN4 Integrated DNA Technologies https://www.idtdna.com/site/order/menu
Cocaine Sigma Aldrich C008
Quinine Sigma Aldrich 22620
NanoDSC-III microcalorimeter TA Instruments http://www.tainstruments.com/nanodsc/
DSCRun software TA Instruments http://www.tainstruments.com/support/software-downloads-support/instruments-by-software/
NanoAnalyze software TA Instruments http://www.tainstruments.com/support/software-downloads-support/instruments-by-software/
Contrad-70 VWR 89233-152

References

  1. Bruylants, G., Wouters, J., Michaux, C. Differential scanning calorimetry in life science: thermodynamics, stability, molecular recognition and application in drug design. Curr Med Chem. 12 (17), 2011-2020 (2005).
  2. Privalov, P. L., Dragan, A. I. Microcalorimetry of biological macromolecules. Biophys Chem. 126 (1-3), 16-24 (2007).
  3. Brandts, J. F., Lin, L. N. Study of strong to ultratight protein interactions using differential scanning calorimetry. 生物化学. 29 (29), 6927-6940 (1990).
  4. Harkness, R. W., Slavkovic, S., Johnson, P. E., Mittermaier, A. K. Rapid characterization of folding and binding interactions with thermolabile ligands by DSC. Chem Commun. 52 (92), 13471-13474 (2016).
  5. Garbett, N. C., Chaires, J. B. Thermodynamic studies for drug design and screening. Expert Opin Drug Dis. 7 (4), 299-314 (2012).
  6. Holdgate, G. A., Ward, W. H. J. Measurements of binding thermodynamics in drug discovery. Drug Discov Today. 10 (22), 1543-1550 (2005).
  7. Plotnikov, V., et al. An autosampling differential scanning calorimeter instrument for studying molecular interactions. Assay Drug Dev Technol. 1 (1), 83-90 (2002).
  8. Schon, A., Lam, S. Y., Freire, E. Thermodynamics-based drug design: strategies for inhibiting protein-protein interactions. Future Med Chem. 3 (9), 1129-1137 (2011).
  9. Periánez Parraga, L., G-L, A., Gamón Runnenberg, I., Seco Melantuche, R., Delgado Sánchez, O., Puigventós Latorre, F. Thermolabile Drugs. Operating Procedure In the Event of Cold Chain Failure. Farmacia Hospitalaria. 35 (4), 1-28 (2011).
  10. Murray, J. B., Alshora, H. I. Stability of Cocaine in Aqueous-Solution. J Clin Pharmacy. 3 (1), 1-6 (1978).
  11. Waterman, K. C., et al. Hydrolysis in pharmaceutical formulations. Pharm. Dev. Technol. 7 (2), 113-146 (2002).
  12. Mergny, J. L., Lacroix, L. Analysis of thermal melting curves. Oligonucleotides. 13 (6), 515-537 (2003).
  13. Neves, M. A., Reinstein, O., Johnson, P. E. Defining a stem length-dependent binding mechanism for the cocaine-binding aptamer. A combined NMR and calorimetry study. 生物化学. 49 (39), 8478-8487 (2010).
  14. Bonifacio, G. F., Brown, T., Conn, G. L., Lane, A. N. Comparison of the electrophoretic and hydrodynamic properties of DNA and RNA oligonucleotide duplexes. Biophys J. 73 (3), 1532-1538 (1997).
  15. Durchschlag, H., Hinz, H. -. J. Chapter 3. Thermodynamic Data for Biochemistry and Biotechnology. , 45-128 (1986).
  16. Hellman, L. M., Rodgers, D. W., Fried, M. G. Phenomenological partial-specific volumes for G-quadruplex DNAs. Eur Biophys J Biophy. 39 (3), 389-396 (2010).
  17. Farber, P., Darmawan, H., Sprules, T., Mittermaier, A. Analyzing Protein Folding Cooperativity by Differential Scanning Calorimetry and NMR Spectroscopy. J Am Chem Soc. 132 (17), 6214-6222 (2010).
  18. Reinstein, O., et al. Quinine binding by the cocaine-binding aptamer. Thermodynamic and hydrodynamic analysis of high-affinity binding of an off-target ligand. 生物化学. 52 (48), 8652-8662 (2013).
  19. Tellinghuisen, J. Statistical error propagation. J Phys Chem. A. 105 (15), 3917-3921 (2001).
  20. Drobnak, I., Vesnaver, G., Lah, J. Model-based thermodynamic analysis of reversible unfolding processes. J Phys Chem B. 114 (26), 8713-8722 (2010).
check_url/cn/55959?article_type=t

Play Video

Cite This Article
Harkness V, R. W., Johnson, P. E., Mittermaier, A. K. Measuring Biomolecular DSC Profiles with Thermolabile Ligands to Rapidly Characterize Folding and Binding Interactions. J. Vis. Exp. (129), e55959, doi:10.3791/55959 (2017).

View Video