International
Tables for
Crystallography
Volume F
Crystallography of biological macromolecules
Edited by M. G. Rossmann and E. Arnold

International Tables for Crystallography (2006). Vol. F. ch. 18.2, pp. 377-379   | 1 | 2 |

Section 18.2.4. Searching conformational space

A. T. Brunger,a* P. D. Adamsb and L. M. Ricec

a The Howard Hughes Medical Institute, and Departments of Molecular and Cellular Physiology, Neurology and Neurological Sciences, and Stanford Synchrotron Radiation Laboratory, Stanford Universty, 1201 Welch Road, MSLS P210, Stanford, CA 94305-5489, USA,bThe Howard Hughes Medical Institute and Department of Molecular Biophysics and Biochemistry, Yale University, New Haven, CT 06511, USA, and cDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, CT 06511, USA
Correspondence e-mail:  axel.brunger@stanford.edu

18.2.4. Searching conformational space

| top | pdf |

Annealing denotes a physical process wherein a solid is heated until all particles randomly arrange themselves in a liquid phase and is then cooled slowly so that all particles arrange themselves in the lowest energy state. By formally defining the target, E [equation (18.2.3.1)[link]], to be the equivalent of the potential energy of the system, one can simulate such an annealing process (Kirkpatrick et al., 1983[link]). There is no guarantee that simulated annealing will find the global minimum (Laarhoven & Aarts, 1987[link]). However, compared to conjugate-gradient minimization, where search directions must follow the gradient, simulated annealing achieves more optimal solutions by allowing motion against the gradient (Kirkpatrick et al., 1983[link]). The likelihood of uphill motion is determined by a control parameter referred to as temperature. The higher the temperature, the more likely it is that simulated annealing will overcome barriers (Fig. 18.2.4.1)[link]. It should be noted that the simulated-annealing temperature normally has no physical meaning and merely determines the likelihood of overcoming barriers of the target function in equation (18.2.3.1)[link].

[Figure 18.2.4.1]

Figure 18.2.4.1 | top | pdf |

Illustration of simulated annealing for minimization of a one-dimensional function. The kinetic energy of the system (a `ball' rolling on the one-dimensional surface) allows local conformational transitions with barriers smaller than the kinetic energy. If a larger drop in energy is encountered, the excess kinetic energy is dissipated. It is thus unlikely that the system can climb out of the global minimum once it has reached it.

The simulated-annealing algorithm requires a mechanism to create a Boltzmann distribution at a given temperature, T, and an annealing schedule, that is, a sequence of temperatures [T_{1} \geq T_{2} \geq \ldots \geq T_{l}] at which the Boltzmann distribution is computed. Implementations differ in the way they generate a transition, or move, from one set of parameters to another that is consistent with the Boltzmann distribution at a given temperature. The two most widely used methods are Metropolis Monte Carlo (Metropolis et al., 1953[link]) and molecular dynamics (Verlet, 1967[link]) simulations. For X-ray crystallographic refinement, molecular dynamics has proven extremely successful (Brünger et al., 1987[link]) because it limits the search to physically reasonable `moves'.

18.2.4.1. Molecular dynamics

| top | pdf |

A suitably chosen set of atomic parameters can be viewed as generalized coordinates that are propagated in time by the classical equations of motion (Goldstein, 1980[link]). If the generalized coordinates represent the x, y, z positions of the atoms of a molecule, the classical equations of motion reduce to the familiar Newton's second law: [m_{i} {\partial^{2}{\bf r}_{i} \over \partial t^{2}} = -\nabla_{i}E. \eqno(18.2.4.1)] The quantities [m_{i}] and [{\bf r}_{i}] are, respectively, the mass and coordinates of atom i, and E is given by equation (18.2.3.1)[link]. The solution of the partial differential equations (18.2.4.1)[link] can be achieved numerically using finite-difference methods (Verlet, 1967[link]; Abramowitz & Stegun, 1968[link]). This approach is referred to as molecular dynamics.

Initial velocities for the integration of equation (18.2.4.1)[link] are usually assigned randomly from a Maxwell distribution at the appropriate temperature. Assignment of different initial velocities will generally produce a somewhat different structure after simulated annealing. By performing several refinements with different initial velocities, one can therefore improve the chances of success of simulated-annealing refinement. Furthermore, this improved sampling can be used to study discrete disorder and conformational variability, especially when using torsion-angle molecular dynamics (see below).

Although Cartesian (i.e. flexible bond lengths and bond angles) molecular dynamics places restraints on bond lengths and bond angles [through [E_{\rm chem}], equation (18.2.3.1)[link]], one might want to implement these restrictions as constraints, i.e., fixed bond lengths and bond angles (Diamond, 1971[link]). This is supported by the observation that the deviations from ideal bond lengths and bond angles are usually small in macromolecular X-ray crystal structures. Indeed, fixed-length constraints have been applied to crystallographic refinement by least-squares minimization (Diamond, 1971[link]). It is only recently, however, that efficient and robust algorithms have become available for molecular dynamics in torsion-angle space (Bae & Haug, 1987[link], 1988[link]; Jain et al., 1993[link]; Rice & Brünger, 1994[link]). We chose an approach that retains the Cartesian-coordinate formulation of the target function and its derivatives with respect to atomic coordinates, so that the calculation remains relatively straightforward and can be applied to any macromolecule or their complexes (Rice & Brünger, 1994[link]). In this formulation, the expression for the acceleration becomes a function of positions and velocities. Iterative equations of motion for constrained dynamics in this formulation can be derived and solved by finite-difference methods (Abramowitz & Stegun, 1968[link]). This method is numerically very robust and has a significantly increased radius of convergence in crystallographic refinement compared to Cartesian molecular dynamics (Rice & Brünger, 1994[link]).

18.2.4.2. Temperature control

| top | pdf |

Simulated annealing requires the control of the temperature during molecular dynamics. The current temperature of the simulation [(T_{\rm curr})] is computed from the kinetic energy [E_{\rm kin} = \sum\limits_{i}^{n}{\textstyle{1 \over 2}} m_{i} {\displaystyle\left({\partial r_{i} \over \partial t}\right)}^{2} \eqno(18.2.4.2)] of the molecular-dynamics simulation, [T_{\rm curr} = 2E_{\rm kin}/3nk_{B}. \eqno(18.2.4.3)] Here, n is the number of atoms, [m_{i}] is the mass of the atom and [k_{B}] is Boltzmann's constant. One commonly used approach to control the temperature of the simulation consists of coupling the equations of motion to a heat bath through a `friction' term (Berendsen et al., 1984[link]). Another approach is to rescale periodically the velocities in order to match [T_{\rm curr}] with the target temperature.

18.2.4.3. Annealing schedules

| top | pdf |

The simulated-annealing temperature needs to be high enough to allow conformational transitions, but not so high that the model moves too far away from the correct structure. The optimal temperature for a given starting structure is a matter of trial and error. Starting temperatures that work for the average case have been determined for a variety of simulated-annealing protocols (Brünger, 1988[link]; Adams et al., 1997[link]). However, it might be worth trying a different temperature if a particularly difficult refinement problem is encountered. In particular, significantly higher temperatures are attainable using torsion-angle molecular dynamics. Note that each simulated-annealing refinement is subject to `chance' by using a random-number generator to generate the initial velocities. Thus, multiple simulated annealing runs can be carried out in order to increase the success rate of the refinement. The best structure(s) (as determined by the free R value) among a set of refinements using different initial velocities and/or temperatures can be taken for further refinement or structure-factor averaging (see below).

The annealing schedule can, in principle, be any function of the simulation step (or `time' domain). The two most commonly used protocols are linear slow-cooling or constant-temperature followed by quenching. A slight advantage is obtained with slow cooling (Brünger et al., 1990[link]). The duration of the annealing schedule is another parameter. Too short a protocol does not allow sufficient sampling of conformational space. Too long a protocol may waste computer time, since it is more efficient to run multiple trials than one long refinement protocol (unpublished results).

18.2.4.4. An intuitive explanation of simulated annealing

| top | pdf |

The goal of any optimization problem is to find the global minimum of a target function. In the case of crystallographic refinement, one searches for the conformation or conformations of the molecule that best fit the diffraction data and that simultaneously maintain reasonable covalent and non-covalent interactions. Simulated-annealing refinement has a much larger radius of convergence than conjugate-gradient minimization (see below). It must, therefore, be able to find a lower minimum of the target E [equation (18.2.3.1)[link]] than the local minimum found by simply moving along the negative gradient of E.

It is most easy to visualize this property of simulated annealing in the case of a one-dimensional problem, where the goal is to find the global minimum of a function with multiple minima (Fig. 18.2.4.1)[link]. An intuitive way to understand a molecular-dynamics simulation is to envisage a ball rolling on this one-dimensional surface. When the ball is far from the global minimum, it gains a certain momentum which allows it to cross barriers of the target function [equation (18.2.4.3)[link]]. Slow-cooling temperature control ensures that the ball will eventually reach the global minimum rather than just bouncing across the surface. The initial temperature must be large enough to overcome smaller barriers, but low enough to ensure that the system will not escape the global minimum if it manages to arrive there.

While temperature itself is a global parameter of the system, temperature fluctuations arise principally from local conformational transitions, for example, from an amino-acid side chain falling into the correct orientation. These local changes tend to lower the value of the target E, thus increasing the kinetic energy, and hence the temperature, of the system. Once the temperature control has removed this excess kinetic energy through `heat dissipation', the reverse transition is very unlikely, since it would require a localized increase in kinetic energy where the conformational change occurred in the first place (Fig. 18.2.4.1)[link]. Temperature control maintains a sufficient amount of kinetic energy to allow local conformational corrections, but does not supply enough to allow escape from the global minimum. This explains the observation that, on average, the agreement with the diffraction data will improve, rather than worsen, with simulated annealing.

References

First citation Abramowitz, M. & Stegun, I. (1968). Handbook of mathematical functions. Applied mathematics series, Vol. 55, p. 896. New York: Dover Publications.Google Scholar
First citation Adams, P. D., Pannu, N. S., Read, R. J. & Brünger, A. T. (1997). Cross-validated maximum likelihood enhances crystallographic simulated annealing refinement. Proc. Natl Acad. Sci. USA, 94, 5018–5023.Google Scholar
First citation Bae, D.-S. & Haug, E. J. (1987). A recursive formulation for constrained mechanical system dynamics: Part I. Open loop systems. Mech. Struct. Mach. 15, 359–382.Google Scholar
First citation Bae, D.-S. & Haug, E. J. (1988). A recursive formulation for constrained mechanical system dynamics: Part II. Closed loop systems. Mech. Struct. Mach. 15, 481–506.Google Scholar
First citation Berendsen, H. J. C., Postma, J. P. M., van Gunsteren, W. F., DiNola, A. & Haak, J. R. (1984). Molecular dynamics with coupling to an external bath. J. Chem. Phys. 81, 3684–3690.Google Scholar
First citation Brunger, A. T. (1988). Crystallographic refinement by simulated annealing: application to a 2.8 Å resolution structure of aspartate aminotransferase. J. Mol. Biol. 203, 803–816.Google Scholar
First citation Brunger, A. T., Krukowski, A. & Erickson, J. W. (1990). Slow-cooling protocols for crystallographic refinement by simulated annealing. Acta Cryst. A46, 585–593.Google Scholar
First citation Brunger, A. T., Kuriyan, J. & Karplus, M. (1987). Crystallographic R factor refinement by molecular dynamics. Science, 235, 458–460.Google Scholar
First citation Diamond, R. (1971). A real-space refinement procedure for proteins. Acta Cryst. A27, 436–452.Google Scholar
First citation Goldstein, H. (1980). Classical mechanics. 2nd ed. Reading, Massachusetts: Addison-Wesley.Google Scholar
First citation Jain, A., Vaidehi, N. & Rodriguez, G. (1993). A fast recursive algorithm for molecular dynamics simulation. J. Comput. Phys. 106, 258–268.Google Scholar
First citation Kirkpatrick, S., Gelatt, C. D. & Vecchi, M. P. Jr (1983). Optimization by simulated annealing. Science, 220, 671–680.Google Scholar
First citation Laarhoven, P. J. M. & Aarts, E. H. L. (1987). Editors. Simulated annealing: theory and applications. Dordrecht: D. Reidel Publishing Company.Google Scholar
First citation Metropolis, N., Rosenbluth, M., Rosenbluth, A., Teller, A. & Teller, E. (1953). Equation of state calculations by fast computing machines. J. Chem. Phys. 21, 1087–1092.Google Scholar
First citation Rice, L. M. & Brunger, A. T. (1994). Torsion angle dynamics: reduced variable conformational sampling enhances crystallographic structure refinement. Proteins Struct. Funct. Genet. 19, 277–290.Google Scholar
First citation Verlet, L. (1967). Computer experiments on classical fluids. I. Thermodynamical properties of Lennard–Jones molecules. Phys. Rev. 159, 98–105.Google Scholar








































to end of page
to top of page