Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

The energy landscape of adenylate kinase during catalysis

Abstract

Kinases perform phosphoryl-transfer reactions in milliseconds; without enzymes, these reactions would take about 8,000 years under physiological conditions. Despite extensive studies, a comprehensive understanding of kinase energy landscapes, including both chemical and conformational steps, is lacking. Here we scrutinize the microscopic steps in the catalytic cycle of adenylate kinase, through a combination of NMR measurements during catalysis, pre-steady-state kinetics, molecular-dynamics simulations and crystallography of active complexes. We find that the Mg2+ cofactor activates two distinct molecular events: phosphoryl transfer (>105-fold) and lid opening (103-fold). In contrast, mutation of an essential active site arginine decelerates phosphoryl transfer 103-fold without substantially affecting lid opening. Our results highlight the importance of the entire energy landscape in catalysis and suggest that adenylate kinases have evolved to activate key processes simultaneously by precise placement of a single, charged and very abundant cofactor in a preorganized active site.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Adk free-energy landscape of catalysis and exploration of the P-transfer step by X-ray crystallography.
Figure 2: Role of active site dynamics in efficient P-transfer versus unproductive hydrolysis.
Figure 3: Catalytic effect of the Mg2+ cofactor.
Figure 4: EAdk structure and dynamics during catalysis with and without Mg2+, studied by NMR.
Figure 5: The nature of the divalent cation drastically affects P-transfer but not Adk conformational dynamics.

Similar content being viewed by others

Accession codes

Primary accessions

Biological Magnetic Resonance Data Bank

Protein Data Bank

Referenced accessions

Protein Data Bank

References

  1. Westheimer, F.H. Why nature chose phosphates. Science 235, 1173–1178 (1987).

    Article  CAS  PubMed  Google Scholar 

  2. Bowler, M.W., Cliff, M.J., Waltho, J.P. & Blackburn, G.M. Why did nature select phosphate for its dominant roles in biology? New J. Chem. 34, 784–794 (2010).

    Article  CAS  Google Scholar 

  3. Schroeder, G.K., Lad, C., Wyman, P., Williams, N.H. & Wolfenden, R. The time required for water attack at the phosphorus atom of simple phosphodiesters and of DNA. Proc. Natl. Acad. Sci. USA 103, 4052–4055 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Kamerlin, S.C., Sharma, P.K., Prasad, R.B. & Warshel, A. Why nature really chose phosphate. Q. Rev. Biophys. 46, 1–132 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Lassila, J.K., Zalatan, J.G. & Herschlag, D. Biological phosphoryl-transfer reactions: understanding mechanism and catalysis. Annu. Rev. Biochem. 80, 669–702 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Stockbridge, R.B. & Wolfenden, R. The intrinsic reactivity of ATP and the catalytic proficiencies of kinases acting on glucose, N-acetylgalactosamine, and homoserine: a thermodynamic analysis. J. Biol. Chem. 284, 22747–22757 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Jencks, W.P. Catalysis in chemistry and enzymology (Dover, New York, 1987).

  8. Knowles, J.R. Enzyme-catalyzed phosphoryl transfer reactions. Annu. Rev. Biochem. 49, 877–919 (1980).

    Article  CAS  PubMed  Google Scholar 

  9. Cleland, W.W. & Hengge, A.C. Enzymatic mechanisms of phosphate and sulfate transfer. Chem. Rev. 106, 3252–3278 (2006).

    Article  CAS  PubMed  Google Scholar 

  10. Thatcher, G.R.J. & Kluger, R. Mechanism and catalysis of nucleophilic-substitution in phosphate-esters. Adv. Phys. Org. Chem. 25, 99–265 (1989).

    CAS  Google Scholar 

  11. Hunter, T. Tyrosine phosphorylation: thirty years and counting. Curr. Opin. Cell Biol. 21, 140–146 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Jura, N. et al. Catalytic control in the EGF receptor and its connection to general kinase regulatory mechanisms. Mol. Cell 42, 9–22 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Adams, J.A. Kinetic and catalytic mechanisms of protein kinases. Chem. Rev. 101, 2271–2290 (2001).

    Article  CAS  PubMed  Google Scholar 

  14. Endicott, J.A., Noble, M.E. & Johnson, L.N. The structural basis for control of eukaryotic protein kinases. Annu. Rev. Biochem. 81, 587–613 (2012).

    Article  CAS  PubMed  Google Scholar 

  15. Henzler-Wildman, K., Lei, M., Thai, V., Karplus, M. & Kern, D. A hierarchy of timescales in protein dynamics linked to enzyme catalysis. Nature 450, 913–916 (2007).

    Article  CAS  PubMed  Google Scholar 

  16. Masterson, L.R. et al. Dynamics connect substrate recognition to catalysis in protein kinase A. Nat. Chem. Biol. 6, 821–828 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Mukherjee, K., Sharma, M., Jahn, R., Wahl, M.C. & Sudhof, T.C. Evolution of CASK into a Mg2+-sensitive kinase. Sci. Signal. 3, ra33 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Bao, Z.Q., Jacobsen, D.M. & Young, M.A. Briefly bound to activate: transient binding of a second catalytic magnesium activates the structure and dynamics of CDK2 kinase for catalysis. Structure 19, 675–690 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Tan, Y.W., Hanson, J.A. & Yang, H. Direct Mg2+ binding activates adenylate kinase from Escherichia coli. J. Biol. Chem. 284, 3306–3313 (2009).

    Article  CAS  PubMed  Google Scholar 

  20. Bhabha, G. et al. Divergent evolution of protein conformational dynamics in dihydrofolate reductase. Nat. Struct. Mol. Biol. 20, 1243–1249 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Bhabha, G. et al. A dynamic knockout reveals that conformational fluctuations influence the chemical step of enzyme catalysis. Science 332, 234–238 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Kamerlin, S.C.L. & Warshel, A. At the dawn of the 21st century: is dynamics the missing link for understanding enzyme catalysis? Proteins 78, 1339–1375 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Adamczyk, A.J., Cao, J., Kamerlin, S.C. & Warshel, A. Catalysis by dihydrofolate reductase and other enzymes arises from electrostatic preorganization, not conformational motions. Proc. Natl. Acad. Sci. USA 108, 14115–14120 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Nagel, Z.D. & Klinman, J.P. A 21st century revisionist's view at a turning point in enzymology. Nat. Chem. Biol. 5, 543–550 (2009).

    Article  CAS  PubMed  Google Scholar 

  25. Klinman, J.P. & Kohen, A. Hydrogen tunneling links protein dynamics to enzyme catalysis. Annu. Rev. Biochem. 82, 471–496 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Müller, C.W., Schlauderer, G.J., Reinstein, J. & Schulz, G.E. Adenylate kinase motions during catalysis: an energetic counterweight balancing substrate binding. Structure 4, 147–156 (1996).

    Article  PubMed  Google Scholar 

  27. Müller, C.W. & Schulz, G.E. Structure of the complex between adenylate kinase from Escherichia coli and the inhibitor Ap5a refined at 1.9 Å resolution: a model for a catalytic transition-state. J. Mol. Biol. 224, 159–177 (1992).

    Article  PubMed  Google Scholar 

  28. Berry, M.B., Bae, E.Y., Bilderback, T.R., Glaser, M. & Phillips, G.N. Crystal structure of ADP/AMP complex of Escherichia coli adenylate kinase. Proteins 62, 555–556 (2006).

    Article  CAS  PubMed  Google Scholar 

  29. Berry, M.B. et al. The closed conformation of a highly flexible protein: the structure of Escherichia coli adenylate kinase with bound AMP and AMPPNP. Proteins 19, 183–198 (1994).

    Article  CAS  PubMed  Google Scholar 

  30. Cowan, J.A. Metal activation of enzymes in nucleic acid biochemistry. Chem. Rev. 98, 1067–1088 (1998).

    Article  CAS  PubMed  Google Scholar 

  31. Kladova, A.V. et al. Cobalt-, zinc- and iron-bound forms of adenylate kinase (AK) from the sulfate-reducing bacterium Desulfovibrio gigas: purification, crystallization and preliminary X-ray diffraction analysis. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 65, 926–929 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Wittinghofer, A. Signaling mechanistics: aluminum fluoride for molecule of the year. Curr. Biol. 7, R682–R685 (1997).

    Article  CAS  PubMed  Google Scholar 

  33. Coleman, D.E. et al. Structures of active conformations of Gi alpha 1 and the mechanism of GTP hydrolysis. Science 265, 1405–1412 (1994).

    Article  CAS  PubMed  Google Scholar 

  34. Sondek, J., Lambright, D.G., Noel, J.P., Hamm, H.E. & Sigler, P.B. GTPase mechanism of Gproteins from the 1.7-Å crystal structure of transducin α-GDP AIF4 . Nature 372, 276–279 (1994).

    Article  CAS  PubMed  Google Scholar 

  35. Cliff, M.J. et al. Transition state analogue structures of human phosphoglycerate kinase establish the importance of charge balance in catalysis. J. Am. Chem. Soc. 132, 6507–6516 (2010).

    Article  CAS  PubMed  Google Scholar 

  36. Baxter, N.J. et al. Atomic details of near-transition state conformers for enzyme phosphoryl transfer revealed by rather than by phosphoranes. Proc. Natl. Acad. Sci. USA 107, 4555 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Baxter, N.J. et al. Anionic charge is prioritized over geometry in aluminum and magnesium fluoride transition state analogs of phosphoryl transfer enzymes. J. Am. Chem. Soc. 130, 3952–3958 (2008).

    Article  CAS  PubMed  Google Scholar 

  38. Warshel, A. et al. Electrostatic basis for enzyme catalysis. Chem. Rev. 106, 3210–3235 (2006).

    Article  CAS  PubMed  Google Scholar 

  39. Armstrong, R.N., Kondo, H. & Kaiser, E.T. Cyclic AMP-dependent ATPase activity of bovine heart protein kinase. Proc. Natl. Acad. Sci. USA 76, 722–725 (1979).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Prowse, C.N. & Lew, J. Mechanism of activation of ERK2 by dual phosphorylation. J. Biol. Chem. 276, 99–103 (2001).

    Article  CAS  PubMed  Google Scholar 

  41. Wang, Y., Gan, L., Wang, E. & Wang, J. Exploring the dynamic functional landscape of adenylate kinase modulated by substrates. J. Chem. Theory Comput. 9, 84–95 (2013).

    Article  CAS  PubMed  Google Scholar 

  42. Gur, M., Madura, J.D. & Bahar, I. Global transitions of proteins explored by a multiscale hybrid methodology: application to adenylate kinase. Biophys. J. 105, 1643–1652 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Henzler-Wildman, K.A. et al. Intrinsic motions along an enzymatic reaction trajectory. Nature 450, 838–844 (2007).

    Article  CAS  PubMed  Google Scholar 

  44. Wolf-Watz, M. et al. Linkage between dynamics and catalysis in a thermophilic-mesophilic enzyme pair. Nat. Struct. Mol. Biol. 11, 945–949 (2004).

    Article  CAS  PubMed  Google Scholar 

  45. Palmer, A.G. NMR characterization of the dynamics of biomacromolecules. Chem. Rev. 104, 3623–3640 (2004).

    Article  CAS  PubMed  Google Scholar 

  46. Yan, H.G. & Tsai, M.D. Nucleoside monophosphate kinases: structure, mechanism, and substrate specificity. Adv. Enzymol. Relat. Areas Mol. Biol. 73, 103–134 (1999).

    CAS  PubMed  Google Scholar 

  47. Page, M.I. & Jencks, W.P. Entropic contributions to rate accelerations in enzymic and intramolecular reactions and the chelate effect. Proc. Natl. Acad. Sci. USA 68, 1678–1683 (1971).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Privett, H.K. et al. Iterative approach to computational enzyme design. Proc. Natl. Acad. Sci. USA 109, 3790–3795 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Röthlisberger, D. et al. Kemp elimination catalysts by computational enzyme design. Nature 453, 190–195 (2008).

    Article  CAS  PubMed  Google Scholar 

  50. Kiss, G., Celebi-Olcum, N., Moretti, R., Baker, D. & Houk, K.N. Computational enzyme design. Angew. Chem. Int. Edn Engl. 52, 5700–5725 (2013).

    Article  CAS  Google Scholar 

  51. Blomberg, R. et al. Precision is essential for efficient catalysis in an evolved Kemp eliminase. Nature 503, 418–421 (2013).

    Article  CAS  PubMed  Google Scholar 

  52. Schwartz, S.D. & Schramm, V.L. Enzymatic transition states and dynamic motion in barrier crossing. Nat. Chem. Biol. 5, 551–558 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Benkovic, S.J. & Hammes-Schiffer, S. A perspective on enzyme catalysis. Science 301, 1196–1202 (2003).

    Article  CAS  PubMed  Google Scholar 

  54. Klinman, J.P. An integrated model for enzyme catalysis emerges from studies of hydrogen tunneling. Chem. Phys. Lett. 471, 179–193 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Boehr, D.D., Dyson, H.J. & Wright, P.E. An NMR perspective on enzyme dynamics. Chem. Rev. 106, 3055–3079 (2006).

    Article  CAS  PubMed  Google Scholar 

  56. Rozovsky, S. & McDermott, A.E. The time scale of the catalytic loop motion in triosephosphate isomerase. J. Mol. Biol. 310, 259–270 (2001).

    Article  CAS  PubMed  Google Scholar 

  57. Fierke, C.A., Johnson, K.A. & Benkovic, S.J. Construction and evaluation of the kinetic scheme associated with dihydrofolate reductase from Escherichia coli. Biochemistry 26, 4085–4092 (1987).

    Article  CAS  PubMed  Google Scholar 

  58. Battye, T.G., Kontogiannis, L., Johnson, O., Powell, H.R. & Leslie, A.G. iMOSFLM: a new graphical interface for diffraction-image processing with MOSFLM. Acta Crystallogr. D Biol. Crystallogr. 67, 271–281 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Evans, P. Scaling and assessment of data quality. Acta Crystallogr. D Biol. Crystallogr. 62, 72–82 (2006).

    Article  CAS  PubMed  Google Scholar 

  60. McCoy, A.J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Murshudov, G.N. et al. REFMAC5 for the refinement of macromolecular crystal structures. Acta Crystallogr. D Biol. Crystallogr. 67, 355–367 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Winn, M.D. et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D Biol. Crystallogr. 67, 235–242 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 (2004).

    Article  CAS  PubMed  Google Scholar 

  64. Emsley, P., Lohkamp, B., Scott, W.G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Adams, P.D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66, 213–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Afonine, P.V. et al. phenix.model_vs_data: a high-level tool for the calculation of crystallographic model and data statistics. J. Appl. Crystallogr. 43, 669–676 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Theobald, D.L. & Steindel, P.A. Optimal simultaneous superpositioning of multiple structures with missing data. Bioinformatics 28, 1972–1979 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. MacKerell, A.D. et al. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 102, 3586–3616 (1998).

    Article  CAS  PubMed  Google Scholar 

  69. MacKerell, A.D., Banavali, N. & Foloppe, N. Development and current status of the CHARMM force field for nucleic acids. Biopolymers 56, 257–265 (2000).

    Article  CAS  PubMed  Google Scholar 

  70. Brooks, B.R. et al. Charmm: a program for macromolecular energy, minimization, and dynamics calculations. J. Comput. Chem. 4, 187–217 (1983).

    Article  CAS  Google Scholar 

  71. Pavelites, J.J., Gao, J.L., Bash, P.A. & Mackerell, A.D. A molecular mechanics force field for NAD+, NADH, and the pyrophosphate groups of nucleotides. J. Comput. Chem. 18, 221–239 (1997).

    Article  CAS  Google Scholar 

  72. Price, D.J. & Brooks, C.L. A modified TIP3P water potential for simulation with Ewald summation. J. Chem. Phys. 121, 10096–10103 (2004).

    Article  CAS  PubMed  Google Scholar 

  73. Gresh, N. et al. Analysis of the interactions taking place in the recognition site of a bimetallic Mg(II)-Zn(II) enzyme, isopentenyl diphosphate isomerase. a parallel quantum-chemical and polarizable molecular mechanics study. J. Phys. Chem. B 114, 4884–4895 (2010).

    Article  CAS  PubMed  Google Scholar 

  74. Pontikis, G., Borden, J., Martinek, V. & Florian, J. Linear energy relationships for the octahedral preference of Mg, Ca and transition metal ions. J. Phys. Chem. A 113, 3588–3593 (2009).

    Article  CAS  PubMed  Google Scholar 

  75. Yang, S.Y. et al. Whether proton transition to the triphosphate tail of ATP occurs at protein kinase environment: a Car-Parrinello ab initio molecular dynamics study. Int. J. Quantum Chem. 108, 1239–1245 (2008).

    Article  CAS  Google Scholar 

  76. Peters, M.B. et al. Structural survey of zinc containing proteins and the development of the zinc amber force field (ZAFF). J. Chem. Theory Comput. 6, 2935–2947 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Ditzler, M.A., Otyepka, M., Sponer, J. & Walter, N.G. Molecular dynamics and quantum mechanics of RNA: conformational and chemical change we can believe in. Acc. Chem. Res. 43, 40–47 (2010).

    Article  CAS  PubMed  Google Scholar 

  78. Banás, P., Jurecka, P., Walter, N.G., Sponer, J. & Otyepka, M. Theoretical studies of RNA catalysis: hybrid QM/MM methods and their comparison with MD and QM. Methods 49, 202–216 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Venkatramani, R. & Radhakrishnan, R. Computational study of the force dependence of phosphoryl transfer during DNA synthesis by a high fidelity polymerase. Phys. Rev. Lett. 100, 088102 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Banás, P., Walter, N.G., Sponer, J. & Otyepka, M. Protonation states of the key active site residues and structural dynamics of the glmS Riboswitch as revealed by molecular dynamics. J. Phys. Chem. B 114, 8701–8712 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Khavrutskii, I.V., Grant, B., Taylor, S.S. & McCammon, J.A. A transition path ensemble study reveals a linchpin role for Mg2+ during rate-limiting ADP release from protein kinase A. Biochemistry 48, 11532–11545 (2009).

    Article  CAS  PubMed  Google Scholar 

  82. Phillips, J.C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem. 26, 1781–1802 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an N•log(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  CAS  Google Scholar 

  84. Ryckaert, J.P., Ciccotti, G. & Berendsen, H.J.C. Numerical-integration of Cartesian equations of motion of a system with constraints: molecular-dynamics of N-alkanes. J. Comput. Phys. 23, 327–341 (1977).

    Article  CAS  Google Scholar 

  85. Brünger, A.T. X-PLOR, Version 3.1: a system for X-ray crystallography and NMR (Yale University Press, New Haven, 1992).

  86. Feller, S.E., Zhang, Y.H., Pastor, R.W. & Brooks, B.R. Constant-pressure molecular-dynamics simulation: the Langevin piston method. J. Chem. Phys. 103, 4613–4621 (1995).

    Article  CAS  Google Scholar 

  87. Martyna, G.J., Tobias, D.J. & Klein, M.L. Constant-pressure molecular-dynamics algorithms. J. Chem. Phys. 101, 4177–4189 (1994).

    Article  CAS  Google Scholar 

  88. Hess, B., Kutzner, C., van der Spoel, D. & Lindahl, E. GROMACS 4: algorithms for highly efficient, load-balanced, and scalable molecular simulation. J. Chem. Theory Comput. 4, 435–447 (2008).

    Article  CAS  PubMed  Google Scholar 

  89. Hess, B., Bekker, H., Berendsen, H.J.C. & Fraaije, J.G.E.M. LINCS: a linear constraint solver for molecular simulations. J. Comput. Chem. 18, 1463–1472 (1997).

    Article  CAS  Google Scholar 

  90. Miyamoto, S. & Kollman, P.A. Settle: an analytical version of the shake and rattle algorithm for rigid water models. J. Comput. Chem. 13, 952–962 (1992).

    Article  CAS  Google Scholar 

  91. Nose, S. A molecular-dynamics method for simulations in the canonical ensemble. Mol. Phys. 52, 255–268 (1984).

    Article  CAS  Google Scholar 

  92. Hoover, W.G. Canonical dynamics: equilibrium phase-space distributions. Phys. Rev. A 31, 1695–1697 (1985).

    Article  CAS  Google Scholar 

  93. Parrinello, M. & Rahman, A. Polymorphic transitions in single-crystals: a new molecular-dynamics method. J. Appl. Phys. 52, 7182–7190 (1981).

    Article  CAS  Google Scholar 

  94. Nose, S. & Klein, M.L. Constant pressure molecular-dynamics for molecular-systems. Mol. Phys. 50, 1055–1076 (1983).

    Article  CAS  Google Scholar 

  95. Liang, J., Edelsbrunner, H. & Woodward, C. Anatomy of protein pockets and cavities: measurement of binding site geometry and implications for ligand design. Protein Sci. 7, 1884–1897 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Le Guilloux, V., Schmidtke, P. & Tuffery, P. Fpocket: an open source platform for ligand pocket detection. BMC Bioinformatics 10, 138 (2009).

    Article  Google Scholar 

  97. Patel, A.J., Varilly, P., Chandler, D. & Garde, S. Quantifying density fluctuations in volumes of all shapes and sizes using indirect umbrella sampling. J. Stat. Phys. 145, 265–275 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  98. Baker, N.A., Sept, D., Joseph, S., Holst, M.J. & McCammon, J.A. Electrostatics of nanosystems: application to microtubules and the ribosome. Proc. Natl. Acad. Sci. USA 98, 10037–10041 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. 14, 33–38 (1996).

    Article  CAS  PubMed  Google Scholar 

  100. Loria, J.P., Rance, M. & Palmer, A.G.A. TROSY CPMG sequence for characterizing chemical exchange in large proteins. J. Biomol. NMR 15, 151–155 (1999).

    Article  CAS  PubMed  Google Scholar 

  101. Delaglio, F. et al. NMRPipe: a multidimensional spectral processing system based on UNIX pipes. J. Biomol. NMR 6, 277–293 (1995).

    Article  CAS  PubMed  Google Scholar 

  102. Johnson, B.A. & Blevins, R.A. NMR View: a computer program for the visualization and analysis of NMR data. J. Biomol. NMR 4, 603–614 (1994).

    Article  CAS  PubMed  Google Scholar 

  103. Vranken, W.F. et al. The CCPN data model for NMR spectroscopy: development of a software pipeline. Proteinss 59, 687–696 (2005).

    Article  CAS  Google Scholar 

  104. Davis, D.G., Perlman, M.E. & London, R.E. Direct measurements of the dissociation-rate constant for inhibitor-enzyme complexes via the T1 rho and T2 (CPMG) methods. J. Magn. Reson. B. 104, 266–275 (1994).

    Article  CAS  PubMed  Google Scholar 

  105. Carver, J. & Richards, R. A general two-site solution for the chemical exchange produced dependence of T2 upon the Carr-Purcell pulse separation. J. Magn. Res. 6, 89–105 (1972).

    CAS  Google Scholar 

  106. Jen, J. Chemical exchange and NMR T2 relaxation: the multisite case. J. Magn. Res. 30, 111–128 (1978).

    CAS  Google Scholar 

  107. Millet, O., Loria, J.P., Kroenke, C.D., Pons, M. & Palmer, A.G. III. The static magnetic field dependence of chemical exchange linebroadening defines the NMR chemical shift time scale. J. Am. Chem. Soc. 122, 2867–2877 (2000).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We dedicate this manuscript to Tom Alber, a truly amazing and inspiring scientist, and a close friend who will live in our hearts forever; his creativity, joy and generosity have deeply influenced several generations of scientists. We are grateful to the staff at the Advanced Light Source–Berkeley Center for Structural Biology and the Advanced Photon Source (APS) for support, Y. Xiong, T. Lang, P. Afonine, R. Read and mentors from the Collaborative Computational Project No. 4 School at APS (2011) for advice in the refinement of X-ray data, members of the C. Miller laboratory for handling crystals, the staff at the National Energy Research Scientific Computing Center, and K.A. Johnson for assistance with the KinTek Explorer software and fitting kinetic data. We thank P. Varilly (University of Cambridge) for kindly providing scripts. This work was supported by the Howard Hughes Medical Institute (HHMI), the Office of Basic Energy Sciences, Catalysis Science Program, US Department of Energy (award DE-FG02-05ER15699), the US National Institutes of Health (RO1-GM100966) and the Teragrid (XSEDE) allocation TG-MCB090166 (D.K.). R.O. is supported as an HHMI Fellow of the Damon Runyon Cancer Research Foundation (DRG-2114-12).

Author information

Authors and Affiliations

Authors

Contributions

S.J.K., R.V.A., Y.-J.C., D.V.P., F.P., M.F.H. and D.K. designed experiments; S.J.K., R.V.A., Y.-J.C., F.P., R.O., D.V.P., L.A.P. and P.N.M. performed experiments; S.J.K., R.V.A., Y.-J.C., F.P., R.O., D.V.P., S.K., L.A.P., P.N.M., V.T., T.A. and D.K. analyzed data; S.J.K., R.V.A., F.P., R.O. and D.K. wrote the manuscript.

Corresponding author

Correspondence to Dorothee Kern.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Magnesium coordination and water in the active site.

a,b. Stereo view of Fig. 1e showing detailed active-site interactions for the Mg2+–ADP–ADP (blue) and Mg2+–ADP–AMP–AlF4 structures (green). c. Superimposition of X-ray structures (PDB 3SR0 and 2RGX), and a representative snapshot from the MD simulation. The positions of the divalent cation, water molecules, and the nucleotide oxygen atoms in the simulation are in full agreement with the corresponding atoms in the crystal structures. The blue wireframe surface represents the isosurface of value 0.99 for the fractional occupancy of the water oxygen atoms during a typical MD simulation. d. Octahedral coordination of the Mg2+ ion in the active site monitored during a 200 ns MD simulation of AAdk bound to Mg2+–ADP–ADP. The coordination partners are depicted on the structure and the corresponding time-dependent Mg2+–O distances during the MD simulation are shown.

1. Henzler-Wildman, K.A. et al. Intrinsic motions along an enzymatic reaction trajectory. Nature 450, 838-44 (2007)

Supplementary Figure 2 Mg2+ prevents water molecules trapped in the active site from exchanging with bulk water.

(a­–e) Behavior of all water molecules in the active-site cavity for 100 ns MD simulations of AAdk bound to ADP–ADP with and without Mg2+. a. The fraction of the water molecules present in the active site at the beginning of the simulation, that are still inside the cavity at time t is plotted for four simulations in the presence (dark blue to green) and absence (red to yellow) of Mg2+. The cavity considered for the analysis is displayed in the structures in panels be. The water oxygen atoms initially present in the cavity are displayed as large red spheres and the four water molecules coordinating Mg2+ are shown in red and white sticks. In the simulations with Mg2+ (b and c), all water molecules stay in the cavity (see also Movie 1). In the simulations without Mg2+ (d and e), the water molecules in the pocket progressively diffuse out, exchanging with bulk water and are substituted by other water molecules (depicted as cyan spheres). The average number of water molecules in the cavity is constant in all simulations.

Supplementary Figure 3 Comparison of AAdk and EAdk sequences and pre-steady-state kinetics with and without Mg2+, and determination of the on-enzyme equilibrium for EAdk.

a. Sequence alignment of adenylate kinase from Aquifex aeolicus (AAdk) and Escherichia coli (EAdk). Identical residues are colored in red and similar residues in purple. To avoid confusion, the residue numbering used in the present manuscript refers always to the AAdk sequence. For example, when we indicate R150 (marked with * in the sequence), we are in fact referring to R150 for AAdk and R156 for EAdk. b. Quench-flow experiments performed in the forward direction with 100 μM AAdk, 4 mM ATP and AMP, and 8 mM Mg2+. c. Quench-flow experiments performed in the forward direction with 25 μM AAdk, 4 mM ATP and AMP, and 50 mM EDTA to remove Mg2+ contamination. Data shows that the rate-limiting step of AAdk turnover, like EAdk (Fig. 3), is lid-opening in the presence of Mg2+ (b) and phosphoryl transfer in the absence of Mg2+ (c). This highlights the mechanistic similarity between the two proteins, which is further bolstered by the nearly identical active site in crystal structures of AAdk (PDB 2RGX) and EAdk (PDB 1AKE) in complex with the bi-substrate inhibitor Ap5A. d. EDTA titration of EAdk + 4 mM ADP complex monitored by a change in kcat for the reverse reaction. Due to the fact that 50 mM EDTA was needed to remove trace elements of divalent cations, all “no divalent cation” kinetics experiments were performed with 50 mM EDTA to ensure that the measured rates were not influenced by minor divalent cation contamination. e. Quench-flow experiments performed at different EAdk concentrations in the forward (25 and 100 μM) and reverse (485 μM) directions with 4 mM ATP and AMP or 4 mM ADP. 50 mM EDTA was present in all experiments to remove residual Mg2+. Data show the same behavior as at low EAdk concentration (Fig. 3), but fitting was less reliable because the back reaction had to be accounted for, due to the increased amount of generated product. f,g. Determination of the on-enzyme equilibrium in the presence of Mg2+ (f, 600 μM of MgADP) and absence of Mg2+ at (g, 600 μM ADP + 50 mM EDTA). The [ADP]/2*[ATP] ratio was measured as a function of increasing concentration of EAdk. At low enzyme concentration most nucleotides are free, and the ADP to ATP ratio mainly reflects equilibrium in solution. When the plateau is reached, all nucleotides are bound and the corresponding [ADP]/2*[ATP] ratio reflects the on-enzyme equilibrium. In the presence of Mg2+, fitting it to a generalized hyperbola (y = a - b/(1 + c*x)^(1/d)) yielded an equilibrium value of 11.6 ± 0.7, which is in good agreement with the number obtained from kinetic experiments (see Fig. 3a). In the absence of Mg2+ the equilibrium was shifted even further, yielding a value of 29 ± 1.

1. Henzler-Wildman, K.A. et al. Intrinsic motions along an enzymatic reaction trajectory. Nature 450, 838-44 (2007).

2. Muller, C.W. & Schulz, G.E. Structure of the Complex between Adenylate Kinase from Escherichia-Coli and the Inhibitor Ap5a Refined at 1.9 a Resolution - a Model for a Catalytic Transition-State. Journal of Molecular Biology 224, 159-177 (1992)

Figa

Supplementary Figure 4 Effect of the nature of the divalent cation on structure and kinetics.

a. The overlay of [1H-15N]-TROSY-HSQC spectra of EAdk with saturating concentrations of Mg2+ (blue) or Ca2+ (red) and saturating concentrations of nucleotides shows only minor chemical shifts changes indicating that the bound structures are very similar. The NMR experiments were collected with 2 mM EAdk, 20 mM ADP, and either 20 mM Mg2+ or 20 mM Ca2+. b,c. Pre-steady-state kinetics of EAdk measured by quench-flow at 25 °C in the forward (green) and reverse direction (blue) with different divalent ions to obtain the rate constants in Table 2. b. 25 μM EAdk, 4 mM ATP and AMP, and 8 mM CaCl2 in the forward reaction, and 100 μM EAdk and 4 mM ADP and CaCl2 in the reverse direction (blue). c. 100 μM EAdk, 4 mM ATP and AMP, and 8 mM CoCl2 in the forward direction and 100 μM EAdk and 4 mM ADP and CoCl2 in the reverse direction. For Ca2+ and Co2+, P-transfer is rate limiting in the reverse reaction, whereas lid-opening is rate limiting in the forward reaction (burst represents P-transfer).

Supplementary Figure 5 Pre-steady-state kinetics measurements of EAdk R150K at 25 °C.

a. Quench-flow experiments performed with 100 μM EAdk–R150K and 4 mM MgAMP + 4 mM MgATP in the forward direction and 100 μM EAdk–R150K and 4 mM MgADP in the reverse direction. b. 100 μM EAdK–R150K, 4 mM ATP + 4 mM AMP, and 50 mM EDTA in the forward direction and 100 μM EAdK-R150K, 4 mM ADP, and 50 mM EDTA in the reverse direction. For R150K, P-transfer is rate limiting in both directions with and without Mg2+.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–5, Supplementary Tables 1 and 2, and Supplementary Notes 1–4 (PDF 5442 kb)

Water dynamics in the presence of Mg2+

Even in the presence of Mg2+, the water molecules trapped inside the active site are highly dynamic and repeatedly swap their positions, although they cannot escape the active site pocket and exchange with the bulk. This can be easily visualized if the oxygen atom of each water molecule is labeled with a different color so that one can follow their individual dance inside the active site pocket during a 100ns MD trajectory. (MP4 5458 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kerns, S., Agafonov, R., Cho, YJ. et al. The energy landscape of adenylate kinase during catalysis. Nat Struct Mol Biol 22, 124–131 (2015). https://doi.org/10.1038/nsmb.2941

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.2941

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing