Minimization - General Methodology


Many issues are involved in designing an appropriate simulation strategy for a given molecular system, some of which have to do only with the minimization algorithms themselves. One of the most important steps in any simulation is properly preparing the system to be simulated. Calculations on the fastest computer running the most efficient minimization or dynamics algorithm may be worthless if the hydrogen is put on the wrong nitrogen of an active-site histidine or a structural water is omitted from the active site. Unfortunately, it is impossible to provide a single recipe for a successful model-too much depends on the objectives and expectations of each calculation. Are energies to be compared quantitatively? Are conformational changes that are far removed in conformational space from the area being focused on expected or interesting? What is the hypothesis being tested? The effects of tethering, fixing, energy cutoffs, etc. on the results can be answered only by controlled preliminary experiments.

For simulation strategies that involve minimization, several considerations must be addressed, including:

Contents


Significance of Minimum-Energy Structure

In dealing with macromolecular optimization calculations, it is important to keep in mind the theoretical significance of the minimum-energy structure and its calculated energy. For all forcefields used in calculations of this type, the energy zero is arbitrary, and therefore, the total potential energy of different molecules cannot be compared directly. However, it is meaningful to make comparisons of energies calculated for different configurations of chemically identical systems. In principle, the calculated energy of a fully minimized structure is the classical enthalpy at absolute zero, ignoring quantum effects (in particular the zero-point vibrational motion). For a molecule that is sufficiently small that its normal modes can be calculated, quantum corrections for zero-point energy and the free energy at higher temperatures can be taken into account (Hagler et al. 1979c).

The minimized energies calculated for enzyme-substrate complexes can be used to estimate relative binding enthalpies, but there are two caveats. First, for a meaningful comparison of the relative binding of two different substrates, a complete thermodynamic cycle must be considered (Kirkwood 1935, Quirke and Jacucci 1982, Tembe and McCammon 1984, Mezei and Beveridge 1986).

In practical terms, this means that an enthalpy calculation for the various substrates in water must be made. Where the relative binding of two different enzymes to the same substrate is calculated, the energy of each enzyme with solvent in the binding site must be calculated.

A second consideration for using minimization results to estimate relative binding strengths is that the entropy is neglected in such calculations. Direct calculation of entropy differences is a computationally intensive process, and only recently has it been taken into account correctly by calculations of relative free energies (Hagler et al. 1979c, Hwang and Warshel 1987, Warshel et al. 1986, Singh et al. 1987, Straatsma et al. 1986).

The extent of the errors introduced by neglecting entropic contributions in the simpler minimization calculations is difficult to estimate, although, as with the zero-point energy, the entropy can be estimated for a system small enough that its normal mode frequencies can be calculated (Hagler et al. 1979c).

The relative importance of these fundamental considerations depends on the objective of the calculation. When studying the relative binding in an enzyme active site of two substrates, one of which is flexible and the other rigid, entropic effects may be crucial for obtaining even qualitative agreement with experimental binding constants. On the other hand, if a putative compound overlaps sterically with many active-site atoms and causes hundreds of kilocalories of strain energy even in a minimized structure, the compound can be rejected confidently. The bottom line is that physical-chemistry common sense cannot be abandoned when you are setting up a calculation and interpreting the results.


When To Use Different Algorithms

The choice of which algorithm to use depends on two factors--the size of the system and the current state of optimization. Until the derivatives are well below 100 kcal mol-1 Å-1, it is likely that the point is sufficiently distant from a minimum that the energy surface is far from quadratic. Algorithms that assume the energy surface to be quadratic (Newton-Raphson, quasi-Newton-Raphson, or conjugate gradients) can be unstable when the molecule is far from the quadratic limit. The Newton-Raphson method is particularly sensitive because it must invert the Hessian matrix. Therefore, as a general rule, steepest descents is often the best minimizer to use for the first 10-100 steps, after which conjugate gradients or a Newton-Raphson minimizer can be used to complete the minimization to convergence. For highly distorted structures, the presence of cross terms and Morse bond potentials in the forcefield can cause convergence problems. These functional forms can produce either small restoring forces or, more seriously, minima at nonphysical points on the potential energy surface. Thus, it is suggested that, in addition to using steepest descents for such distorted structures, you also use a forcefield with a simple, quadratic functional form. In the Discover program with the default CVFF forcefield, this can be accomplished by turning off cross terms and Morse bond potentials.

Several practical aspects of the conjugate gradients method are worth mentioning. First, the conjugate gradients algorithm requires convergence along each line search before continuing in the next direction. The gradient at step i+1 must be perpendicular to hi or the derivation guaranteeing a conjugate set of directions breaks down. Second, to start conjugate gradients, an initial direction h0 must be chosen that is equal to the initial gradient. Finally, additional storage is required for an extra vector of N elements to hold the N components of the old gradient. For energy minimization in Cartesian space this would be the 3N derivatives of the energy with respect to the x, y, and z coordinates of each atom. This makes conjugate gradients the method of choice for systems that are too large for storing and manipulating a second-derivative matrix, as is required by Newton-Raphson style minimizers.

Also, note that the derivation invokes a quadratic approximation. For nonharmonic systems, conjugate gradients can exhaustively minimize along the conjugate directions without converging. This condition generates an error message from the Discover program that the minimizer may have gotten stuck at a saddle point. If this occurs, you can restart the algorithm. Several minimizations may be required. For a detailed discussion of this algorithm, see the excellent text by Press et al. (1986) or the somewhat more formal treatment by Fletcher (1980).


When To Use Constraints/Restraints

Constraints and restraints are often used to control and direct the minimization. For example, if a substrate is being docked onto an enzyme and a specific hydrogen bond between the enzyme and the ligand is thought to be involved in binding, the donor and acceptor atoms can be pulled together to provide a docking coordinate. In this way, the results are not so dependent on the initial starting configuration, which may have been only a crude graphic alignment. In cases like this, the restraint is turned off at some point to make sure that the biased minimum is close to a true minimum.

Another example of the use of restraints is in modeling incomplete systems. Often, it is difficult or impossible to construct a realistic environment around parts of a model system. Only a partial structure of a large protein complex may be available, and some atoms must be restrained to stay near their initial crystal positions because they do not ``feel'' interactions with neighboring (missing) amino acids, membrane, or solvent. If the site of interest (for instance, a binding site for a competitive inhibitor) is well characterized but other parts of the enzyme are unknown or would require too much computation time if they were included, a limited study can still be carried out with the ends tethered to their crystal coordinates. Usually, these restraints are permanent parts of the model. The results of such calculations must be critically evaluated but can be valid if the ligand binding does not depend on interactions with missing pieces of the model or on conformational flexibility in the tethered regions.

As a final example, tethering can be used to gently relax a crystal structure. Often, crystal coordinates, even if highly refined, have several strained interactions due to intrinsically disordered or poorly defined atomic positions, which, upon minimization, give rise to large initial forces. If these forces are not restrained, they can result in artifactual movement away from the original structure. The general approach is to progressively relax parts of the model in stages, starting with the least well determined atoms, until the entire system can minimize freely. The restraints are ultimately removed so that the final minimum represents an unperturbed conformation. It is usually not necessary to minimize to convergence at each stage-the object is to relax the most-strained parts of the system as quickly as possible without introducing artifacts.

A typical approach to relaxing a crystal-derived protein system would be:

  1. Fix the crystal coordinates of the heavy atoms to allow added hydrogens (and perhaps added solvent) to adjust to a static crystallographically defined environment. Only steepest descents is necessary, and this stage is complete once the derivatives are on the order of 10 (subsequent stages allow you to proceed to convergence).
  2. Fix or tether the main chain atoms while the side chains adjust. This may occur in stages, with the surface side chains being relaxed before the interior residues. Again, use steepest descents until the derivatives are less than 10.
  3. Gradually decrease the tethering constant for the backbone atoms until the system can be totally relaxed. Steepest descents can be used initially but should be replaced with a more efficient algorithm as soon as possible.

Convergence Criteria

In the literature a wide variety of criteria have been used to judge minimization convergence in molecular modeling. Mathematically, a minimum is defined as the point at which the derivatives of the function are zero and the second-derivative matrix is positive definite. Nongradient minimizers can use only the increment in the energy and/or coordinates as criteria. In gradient minimizers, derivatives are available analytically and should be used directly to assess convergence.

In a molecular minimization, the atomic derivatives may be summarized as either an average, a root-mean-square (rms) value, or the largest value. The average, of course, must be an average of the absolute values of the derivatives, because the distribution of derivatives is symmetric about zero. A rms derivative is a better measure than the average, because it weights larger derivatives more, and it is therefore less likely that a few large derivatives would escape detection, which can occur with simple averages. Nevertheless, regardless of whether you choose to report convergence in terms of the average or rms values of the derivatives, you should always check that the maximum derivative is not unreasonable. There can be no ambiguity about the quality of the minimum if all derivatives are less than a given value.

The more difficult question is, What value of the average or rms derivative constitutes convergence? The specific value depends on the objective of the minimization. If you simply want to relax overlapping atoms before beginning a molecular dynamics run, minimizing to a maximum derivative of 1.0 kcal mol-1 Å-1 is usually sufficient. However, to perform a normal mode analysis, the maximum derivative must be less than 10-5, or the frequencies may be shifted by several wavenumbers. Figure 4-8 shows the minimization history of the DHFR-trimethoprim protein-substrate system, which required some 14,000 iterations to converge to an average derivative of 0.0002 kcal mol-1 Å-1 (Roberts and Hagler, unpublished data). Often, minimizations of protein systems are stopped when average derivatives between 0.02 kcal mol-1 Å-1 and 0.5 kcal mol-1 Å-1 are achieved. If this simulation had been terminated when the average derivative had reached just 0.02, the structure would have been 30 kcal mol-1 higher in energy and approximately 0.3 Å (rms) away in structure from the final minimum (Roberts and Hagler, unpublished data). This difference represents 25% of the total movement during minimization. Thus, if quantitative measurements and comparisons of macromolecular structures are to be reliable, it would appear from this result that it may be necessary to allow minimizations to converge to average derivatives on the order of 0.002.

It is instructive to note that, although the energy must decrease monotonically during a minimization, the derivative need not. Figure 4-8 shows how the slope of a function can increase in magnitude during a traversal of the surface. This often occurs for functions as complex as molecular energy surfaces.


Main access page Theory/Methodology access.

Minimization access Minimization Algorithms Example Minimization Calculation

Copyright Biosym/MSI