next up previous
Next: A Langevin Model Up: Conformational Dynamics Previous: Conformational Dynamics

Theory of Conformational Substates

Generally, a coarse-grained effective description of a (microcanonical) system with a Hamiltonian , where denote the 3N Cartesian coordinates of the N particles within the system and their momenta, can be achieved by explicitly considering the dynamics of only a few degrees of freedom , commonly referred to as `conformational coordinates'[65]. If the remaining degrees of freedom are regarded as a heat bath, the resulting system of reduced dimension belongs to a canonical ensemble.

The free energy landscape of that sub-system is a potential of mean force[66], , which determines, together with the heat bath, the dynamics of the conformational coordinates . By means of the canonical (projected) phase space density :

 

can be derived from the phase space density generated by the dynamics of the entire system:

 

As an illustration, Figure gif shows a contour-plot of the free energy landscape of our protein model, which has been determined from according to (gif). Here, the two distances and between atoms #12 and #36 as well as #12 and #87, respectively, have been chosen as conformational coordinates (cf. Fig. gif). The canonical ensemble, from which we derived , has been generated from the complete set of 232 1-ns-trajectories by recording both distances every 8 fs in the course of the simulations. Discrete values for have then been determined on a grid of resolution by computing a two-dimensional histogram using square bins of side length. This choice represents a compromise between the conflicting aims of high resolution and low statistical fluctuations of frequency counts per bin. The contour plot in Fig. gif represents a smoothened version of the two-dimensional histogram.

  
Figure: Contour-plot of the free energy landscape derived from a projection of the phase space density onto two conformational coordinates, namely the distances between atoms #12 and #36, and atoms #12 and #87, respectively (cf. Fig. gif). The inset shows the values of the energy landscape in units of along a hypothetical reaction coordinate (bold line) connecting the points A-B-C-D-B-E.

The inset of Fig. gif illustrates the shape of the free energy landscape W by plotting its value in units of (at T=300K) along an arbitrarily chosen `reaction coordinate' (bold line) passing through the three minima. W has been interpolated using a cubic spline function.

The energy landscape in Fig. gif exhibits three distinct minima, denoted as `B', `C', and `D', respectively. As suggested on the basis of experimental data by Frauenfelder[4] such regions of low free energy in conformational space, separated by free energy barriers, generally define distinct conformational substates of a protein. Accordingly, we define our model to be in substate `B', `C', or `D', respectively, if its conformation lies in the corresponding region of the free energy landscape. Like the Brownian motion of a particle coupled to a heat bath, the dynamics of the system within the energy landscape shown in Fig. gif is diffusive. Occasionally, the fluctuating forces generated by the heat bath drive the system across one of the energy barriers and induce a conformational transition, which reveals itself as a rapid change in the sterical structure of the model. Such sudden structural transitions are, e.g., apparent in the lower parts of Fig. gif.

Note, that the above definition of conformational substates differs from the approach commonly employed for theoretical explorations of substate hierarchies[67,68]. In these studies, the distribution of thermally accessible local minima of potential energy within configurational space is studied, and it is assumed, that these local minima or clusters thereof can provide information on the distribution of conformational substates. At low temperatures, where entropic contributions are small and safely can be neglected, the potential energy landscape definitely can serve as a tool for the analysis of conformational substates. However, at room temperature the suggested relation between accessible minima of potential energy and conformational substates, defined as minima of free energy, is questionable. In contrast, our approach allows the study of conformational substates at physiological temperatures, as it refers to the free energy landscape within conformational space. Therefore, if applied to realistic protein models, our approach enables comparisons of theory and experiment. Admittedly, much larger computational effort is involved in such analysis, because a sufficiently dense sampling of phase space by extended simulations is required for the determination of . At present, that computational effort restricts our method to studies of simplified protein models.

As can be seen in Fig. gif, the conformational substates of our protein model are stable on the time scale of several ten picoseconds; conformational transitions occur on scales above a few hundred picoseconds. In order to study, whether memory effects are present in the dynamics on these two time scales, we will now consider two simple stochastic models.

The first model will refer to and will be derived from the protein dynamics within one particular conformational substate. We choose a one-dimensional Langevin model, which describes the Brownian motion of a particle in a potential of mean force, . We will derive this effective potential from the simulation employing the same procedure that was used to compute the free energy landscape shown in Fig. gif. The second model serves to describe the conformational dynamics of transitions between the three conformational substates apparent in Fig. gif on the longer time scale above several hundred picoseconds. We will choose a Markov model which neglects the protein dynamics within each substate.



next up previous
Next: A Langevin Model Up: Conformational Dynamics Previous: Conformational Dynamics



Helmut Grubmueller
Mon Nov 6 16:25:56 MET 1995