1: % $Id: wh.tex 5567 2004-09-09 14:26:54Z ckarney $
2:
3: \newif\ifpreprint\preprinttrue
4: \ifpreprint
5: \documentclass[prl,twocolumn,showkeys,amsfonts,letterpaper,floatfix]{revtex4}
6: \else
7: \documentclass[11pt]{article}\renewcommand{\baselinestretch}{1.65}
8: \fi
9:
10: % dvips omits the newline before %%BeginDocument for included
11: % figures. Fix: use sed to put it back again.
12:
13: % pdflatex wh &&
14: % latex wh &&
15: % dvips -Pwww -f -z wh.dvi |
16: % sed -e 's/%%BeginDocument/~%%BeginDocument/' |
17: % tr '~' '\n' > wh.ps
18:
19: \def\DATE{July 14, 2004; revised September 9, 2004}
20:
21: \usepackage{times}
22: \ifpreprint\else\usepackage{overcite}\usepackage{amssymb}\fi
23: \usepackage{graphicx}\paperwidth=8.5truein\paperheight=11truein
24: \usepackage[bookmarks=false]{hyperref}
25: \hypersetup{pdftitle={Method for Computing Protein Binding Affinity},
26: pdfauthor={C. Karney, J. Ferrara, S. Brunner}}
27:
28: \hyphenation{tryp-sin ca-non-i-cal
29: mol-e-cule mol-e-cules
30: benz-ami-dine benz-ami-dines}
31:
32: % Correct spacing in eqnarray's. Normal spacing around relations is a
33: % ``thick space'' = 5/18 em. The array environment wants half of this,
34: % i.e., 5/36 em.
35: \ifpreprint\else\setlength{\arraycolsep}{0.1389em}
36: \fi
37:
38: \ifpreprint\else
39: % Use more of the page
40: \addtolength\textwidth{0.5in}\addtolength\oddsidemargin{-0.25in}
41: \addtolength\textheight{1.5in}\addtolength\topmargin{-0.75in}
42: \fi
43:
44: % Set width of figures
45: \newlength{\figwidth}
46: \ifpreprint
47: \setlength{\figwidth}{0.48\textwidth}
48: \else
49: \setlength{\figwidth}{0.95\textwidth}
50: \fi
51:
52: \ifpreprint\bibliographystyle{unsrtnat}\urlstyle{rm}\else
53: \bibliographystyle{jacs}\fi
54:
55: \begin{document}
56:
57: \title{Method for Computing Protein Binding Affinity}
58:
59: \ifpreprint
60: \author{Charles F. F. Karney}\email{ckarney@sarnoff.com}
61: \author{Jason E. Ferrara}
62: \affiliation{\href{http://www.sarnoff.com}{Sarnoff Corporation},
63: Princeton, NJ 08543-5300}
64: \author{Stephan Brunner}\altaffiliation
65: [Current address: ]{CRPP, Ecole Polytechnique F\'ed\'erale de Lausanne,
66: CH-1015 Lausanne, Switzerland.}
67: \affiliation{\href{http://www.locuspharma.com}
68: {Locus Pharmaceuticals, Inc.},
69: Blue Bell, PA 19422-2700}
70: \else
71: \author{Charles F. F. Karney\thanks{E-mail: ckarney@sarnoff.com}\enskip
72: and Jason E. Ferrara\\[0.2ex]
73: {\it Sarnoff Corporation, Princeton, NJ 08543-5300, USA}\\[2ex]
74: Stephan Brunner\thanks{Present address: CRPP, Ecole Polytechnique
75: F\'ed\'erale de Lausanne, CH-1015 Lausanne, Switzerland.}\\[0.2ex]
76: {\it Locus Pharmaceuticals, Inc., Blue Bell, PA 19422-2700, USA}}
77: \fi
78:
79: \date{\DATE}
80:
81: \ifpreprint\else \maketitle \fi
82:
83: \begin{abstract}
84: A Monte Carlo method is given to compute the binding affinity of a
85: ligand to a protein. The method involves extending configuration space
86: by a discrete variable indicating whether the ligand is bound to the
87: protein and a special Monte Carlo move which allows transitions between
88: the unbound and bound states. Provided that an accurate protein
89: structure is given, that the protein-ligand binding site is known, and
90: that an accurate chemical force field together with a continuum
91: solvation model is used, this method provides a quantitative estimate of
92: the free energy of binding.
93: \ifpreprint
94: \keywords{free energy; binding affinity; Monte Carlo methods;
95: equilibrium constants; proteins}
96: \else
97: \par\vspace{1.5ex}\noindent
98: Key words: free energy; binding affinity; Monte Carlo methods;
99: equilibrium constants; proteins
100: \fi
101: \end{abstract}
102:
103: \ifpreprint \maketitle \fi
104:
105: \section*{Introduction}
106:
107: Many drugs work by binding to a target protein in an organism and
108: affecting the action of this protein. The binding of the drug molecule,
109: the ligand $\mathrm L$, to the protein $\mathrm P$ under physiological
110: conditions is usually reversible (characterized by weak chemical
111: interactions rather than covalent bonds),
112: \[
113: \mathrm L + \mathrm P \rightleftharpoons \mathrm{LP},
114: \]
115: and, in equilibrium and in the dilute
116: limit, the concentration of the ligand-protein complex $[\mathrm{LP}]$
117: is given by the dissociation constant
118: \[
119: K_d = \frac{[\mathrm L][\mathrm P]}{[\mathrm{LP}]}.
120: \]
121: It is convenient to define the binding affinity as
122: \[\mathrm pK_d =
123: -\log_{10}\biggl(\frac{K_d/N_\mathrm A}{1\,\mathrm{kmol\,m^{-3}}}\biggr),
124: \]
125: where $N_\mathrm A$ is the Avogadro constant. A high value for
126: $\mathrm pK_d$ is crucial to obtaining a good drug molecule and,
127: consequently, the ability to compute $\mathrm pK_d$ accurately would
128: greatly accelerate drug discovery by allowing many molecules to be
129: screened \emph{in silico} before any time-consuming syntheses and assays
130: are done. The dissociation constant can also be expressed in terms of
131: the binding free energy $\Delta F$,
132: \begin{equation}
133: \label{kdfree}
134: K_d = \exp(\beta \Delta F)/V_0,
135: \end{equation}
136: where $V_0$ is the system volume, $\beta = 1/(kT)$, $T$ is the
137: temperature, and $k$ is the Boltzmann constant. Similarly $\mathrm
138: pK_d$ is given by
139: \[
140: \mathrm pK_d = -\frac{\beta \Delta F}{\ln10} +
141: \log_{10}(V_0 N_\mathrm A\times 1\, \mathrm{kmol\, m^{-3}}).
142: \]
143: The quantity $\Delta F$ is the free energy difference of the ligand and
144: the protein forming a bound complex $\mathrm{LP}$, the ``bound system,''
145: compared to the ligand and the protein isolated from one another
146: $\mathrm L + \mathrm P$, the ``unbound system.''
147:
148: In order to compute $\mathrm pK_d$, we require (1)~a sufficiently
149: accurate model of the protein and the ligand and their interaction and
150: (2)~a good way to compute the resulting value of $\mathrm pK_d$. In
151: this paper, we assume the first requirement is fulfilled and instead
152: focus on meeting the second.
153:
154: The standard methods of computing free energies
155: \nocite{mezei86}\cite{mezei86,kollman93}
156: are not capable of computing $\Delta F$ directly because the unbound and
157: bound systems are too dissimilar, which hinders transitions between
158: these systems. Instead, typically, two close ligands $\mathrm L_a$ and
159: $\mathrm L_b$, are compared separately unbound and bound to the protein,
160: thereby obtaining the difference in the free energies $\Delta F_a -
161: \Delta F_b$.
162:
163: We present here a practical method for directly computing $\Delta F$ and
164: hence $\mathrm pK_d$. The method consists of: (1)~formulating the
165: problem in an extended phase space which allows the unbound and bound
166: systems to be treated as a single system and $K_d$ to be expressed as
167: the ratio of two canonical averages; (2)~introducing a new Monte Carlo
168: move, the ``wormhole move,'' to
169: make transitions between the unbound and bound states in this extended
170: system; and (3)~a method to determine the ``portals'' needed for
171: the wormhole move.
172:
173: \section*{Formulation}
174:
175: \ifpreprint
176: Consider a system of volume $V_0$ consisting of a ligand molecule
177: $\mathrm L$ and a protein molecule $\mathrm P$ dissolved in $N_\mathrm
178: S$ molecules $\mathrm S$. The overall state of the system is given by
179: $[\Gamma, \Gamma_\mathrm S]$ where $\Gamma$ represents the phase space
180: configuration of $\mathrm L$ and $\mathrm P$ and $\Gamma_\mathrm S$
181: represents the configurations of all the solvent molecules $\mathrm S$.
182: The energy of the system is given by $E(\Gamma, \Gamma_\mathrm S)$
183: and, in equilibrium, the system obeys the Boltzmann distribution
184: \cite{landau69}
185: \[
186: f(\Gamma, \Gamma_S) = \frac
187: {\exp[-\beta E(\Gamma, \Gamma_S)]}
188: {\int \exp[-\beta E(\Gamma, \Gamma_S)]\, d\Gamma\,d\Gamma_\mathrm S}
189: .
190: \]
191: It is frequently useful to average over the configurations of the
192: solvent molecules by integrating the Boltzmann distribution over
193: $\Gamma_\mathrm S$ to give a reduced Boltzmann distribution
194: \begin{eqnarray*}
195: f(\Gamma) &=&
196: \int f(\Gamma, \Gamma_\mathrm S) \,d\Gamma_\mathrm S\\
197: &=& \frac
198: {\exp[-\beta E(\Gamma)]}
199: {\int \exp[-\beta E(\Gamma)]\, d\Gamma}
200: ,
201: \end{eqnarray*}
202: where
203: \[
204: E(\Gamma) = -\frac1\beta \ln\biggl(
205: \int\exp[-\beta E(\Gamma, \Gamma_\mathrm S)]\,d\Gamma_\mathrm S
206: \biggr)
207: \]
208: is the energy of the system with ligand and protein configurations
209: specified by $\Gamma$ and with the equilibrium effects of the solvent
210: implicitly included as a solvation free energy.
211: In this paper, we will assume that $E(\Gamma)$ is
212: given.
213:
214: Typical molecular interactions have a short range. In view of this, let
215: us define $\Sigma_0$ to represent all accessible $\Gamma$ space
216: (i.e., $\mathrm L$ and $\mathrm P$ somewhere in the system volume
217: $V_0$), and $\Sigma_1$ to represent that portion of $\Sigma_0$
218: where there is an appreciable interaction between $\mathrm L$ and
219: $\mathrm P$ which therefore form the complex $\mathrm{LP}$. In the
220: phase-space volume $\Sigma_1$ we write the energy as $E_1(\Gamma)$
221: which is just an alternate notation for the full energy $E(\Gamma)$,
222: while in the volume $\Sigma_0 - \Sigma_1$ we may write the
223: energy as $E_0(\Gamma)$ which we define as the ``unbound'' energy, i.e.,
224: $E(\Gamma)$ excluding the interaction between $\mathrm L$ and $\mathrm
225: P$. The dissociation constant can then be written as
226: \begin{eqnarray*}
227: K_d &=& \frac1{V_0}\frac
228: {\Bigl(
229: \int_{\Sigma_0 - \Sigma_1} e^{-\beta E(\Gamma)} \,d\Gamma
230: \Bigr)^2}
231: {
232: \int_{\Sigma_0} e^{-\beta E(\Gamma)} \,d\Gamma
233: \int_{\Sigma_1} e^{-\beta E(\Gamma)} \,d\Gamma}\\
234: &=& \frac1{V_0}\frac
235: {\Bigl(
236: \int_{\Sigma_0 - \Sigma_1} e^{-\beta E_0(\Gamma)} \,d\Gamma
237: \Bigr)^2}
238: {\int_{\Sigma_1} e^{-\beta E_1(\Gamma)} \,d\Gamma}\times\\
239: &&\qquad
240: \frac1
241: {\int_{\Sigma_0 - \Sigma_1} e^{-\beta E_0(\Gamma)} \,d\Gamma
242: + \int_{\Sigma_1} e^{-\beta E_1(\Gamma)} \,d\Gamma}
243: .
244: \end{eqnarray*}
245: In the dilute limit $V_0 \rightarrow \infty$, this can be simplified
246: by ignoring the second term in the denominator of the last factor and by
247: extending the limits of the integrals over $\Sigma_0 - \Sigma_1$
248: to include $\Sigma_1$. In extending the definition of $E_0(\Gamma)$
249: to $\Gamma\in\Sigma_1$, we include the intramolecular energy and the
250: solvation free energy but continue to omit the intermolecular
251: (ligand-protein) energy. This yields
252: \nocite{bennett76}\cite{mezei86,bennett76,luo02}
253: \begin{equation}\label{kd}
254: K_d = \frac1{V_0}\frac
255: {\int_{\Sigma_0} \exp[-\beta E_0(\Gamma)] \,d\Gamma}
256: {\int_{\Sigma_1} \exp[-\beta E_1(\Gamma)] \,d\Gamma}
257: .
258: \end{equation}
259: \else
260: Consider a system of volume $V_0$ consisting of a ligand molecule
261: $\mathrm L$ and a protein molecule $\mathrm P$ in a solvent. The state
262: of the system is given by $\Gamma = [\Gamma_\mathrm L, \Gamma_\mathrm
263: P]$ where $\Gamma_\mathrm M$ represents the phase space configuration
264: (position, orientation, and conformation) of molecule $\mathrm M$.
265: In equilibrium, the system obeys the Boltzmann distribution
266: \cite{landau69}
267: \[
268: f(\Gamma) = \frac{\exp[-\beta E(\Gamma)]}
269: {\int \exp[-\beta E(\Gamma)] \,d\Gamma},
270: \]
271: where $E(\Gamma)$ is the energy of the system with the equilibrium
272: effects of the solvent
273: implicitly included as a solvation free energy.
274: In this paper, we will assume that $E(\Gamma)$ is
275: given.
276:
277: Typical molecular interactions have a short range. In view of this, let
278: us define $\Sigma_0$ to represent all accessible $\Gamma$ space
279: (i.e., $\mathrm L$ and $\mathrm P$ somewhere in the system volume
280: $V_0$), and $\Sigma_1$ to represent that portion of $\Sigma_0$
281: where there is an appreciable interaction between $\mathrm L$ and
282: $\mathrm P$ which therefore form the complex $\mathrm{LP}$. In the
283: dilute limit, $V_0 \rightarrow \infty$,
284: the dissociation constant can then be written as
285: \nocite{bennett76}\cite{mezei86,bennett76,luo02}
286: \begin{equation}\label{kd}
287: K_d = \frac1{V_0}\frac
288: {\int_{\Sigma_0} \exp[-\beta E_0(\Gamma)] \,d\Gamma}
289: {\int_{\Sigma_1} \exp[-\beta E_1(\Gamma)] \,d\Gamma}
290: ,
291: \end{equation}
292: where $E_1(\Gamma) = E(\Gamma)$ is the full energy and $E_0(\Gamma)$ is
293: the unbound energy given by ignoring the interaction between the protein
294: and the ligand.
295: \fi
296: The Helmholtz free energy of the unbound and bound
297: systems is \cite{landau69}
298: \[
299: F_\lambda = -\frac1\beta\ln\biggl(
300: \int_{\Sigma_\lambda} \exp[-\beta E_\lambda(\Gamma)] \,d\Gamma
301: \biggr),
302: \]
303: for $\lambda = 0$ and $1$, and Eq.~(\ref{kdfree}) is obtained from
304: Eq.~(\ref{kd}) with $\Delta F = F_1 - F_0$. The definition of $K_d$,
305: Eq.~(\ref{kd}), is strictly independent of $V_0$ because of
306: translational symmetry (ignoring boundary effects). It is also
307: independent of the precise definition of $\Sigma_1$, provided that
308: $\Sigma_1$ includes the protein-ligand binding site and does not
309: include a ``macroscopic'' volume beyond this.
310:
311: In this formulation, we have assumed that the system volume $V_0$ is
312: fixed. However, in most physiological systems, the pressure is held
313: constant and the binding affinity is then related to the differences in
314: the Gibbs free energy which introduces a correction term which is the
315: product of the pressure and the change in the volume caused by the
316: formation of the $\mathrm{LP}$ complex \cite{gilson97}. We expect this
317: correction to be small for typical ligand-protein interactions in
318: solution.
319:
320: We would like to cast Eq.~(\ref{kd}) as the ratio of canonical averages
321: which can be computed using the canonical-ensemble Monte Carlo method
322: \cite{metropolis53}. To achieve this, we combine the unbound and bound
323: systems by extending phase space to include a discrete index $\lambda\in
324: \{0,1\}$ and consider a system in this extended space with energy
325: $E^*_\lambda(\Gamma)$ for which the canonical average is defined by
326: \[
327: \langle X \rangle = \frac
328: {\sum_\lambda \int d\Gamma\, \exp[-\beta E^*_\lambda(\Gamma)]
329: X_\lambda(\Gamma)}
330: {\sum_\lambda \int d\Gamma\, \exp[-\beta E^*_\lambda(\Gamma)]}
331: .
332: \]
333: We take $E^*_\lambda(\Gamma)$ to be infinite for $\Gamma \notin
334: \Sigma_\lambda$ and finite otherwise. Now Eq.~(\ref{kd}) can be
335: rewritten as
336: \begin{equation}\label{kdcanon}
337: K_d = \frac1{V_0} \frac
338: {\bigl< \delta_{\lambda0} \,
339: \exp\bigl(-\beta [E_0(\Gamma) - E^*_0(\Gamma)]\bigr) \bigr>}
340: {\bigl< \delta_{\lambda1} \,
341: \exp\bigl(-\beta [E_1(\Gamma) - E^*_1(\Gamma)]\bigr) \bigr>}
342: ,
343: \end{equation}
344: where $\delta_{\lambda\mu}$ is the Kronecker delta. If
345: $E^*_\lambda(\Gamma) \approx E_\lambda(\Gamma)$, the terms being
346: averaged are $O(1)$. Because the definition of $K_d$ is independent of
347: $V_0$, we can pick $V_0 \sim 1/K_d$ so that approximately the same
348: number of samples contribute to each of the canonical averages.
349: We show later, Eq.~\ref{err}, that this choice minimizes the error in
350: the estimate of $K_d$.
351:
352: We can evaluate $E^*_\lambda(\Gamma)$ with short energy cutoffs allowing
353: it to be computed more quickly than $E_\lambda(\Gamma)$ and the terms
354: contributing to the averages in Eq.~(\ref{kdcanon}) can be accumulated
355: every hundred steps, for example. Since there is typically a high
356: correlation between successive steps in a Monte Carlo simulation, this
357: method allows the averages to be computed to a given degree of accuracy
358: in less time than if we had used $E^*_\lambda(\Gamma) = E_\lambda(\Gamma)$.
359:
360: The extension of phase space has been used in other free energy
361: calculations, to combine, for example, systems at several different
362: temperatures \cite{lyubartsev92} or to treat the ``reaction coordinate''
363: controlling the transition between two chemical species as a dynamic
364: variable \nocite{tidor93}\cite{tidor93,kong96}.
365: In our case, the use of the wormhole
366: Monte Carlo (described in the next section) allows us to include just the two
367: systems of interest without the need to compute the properties of
368: (possibly unphysical) intermediate systems.
369:
370: \section*{Wormhole Monte Carlo}
371:
372: We can compute the canonical averages in Eq.~(\ref{kdcanon}) using
373: the Monte Carlo method \cite{metropolis53} to make steps from $[\Gamma,
374: \lambda]$ to $[\Gamma', \lambda']$ with probability
375: \[
376: \min\bigl[1,
377: \exp\bigl(-\beta [E^*_{\lambda'}(\Gamma')- E^*_\lambda(\Gamma)]\bigr)
378: \bigr].
379: \]
380: However, the estimate of the ratio in Eq.~(\ref{kdcanon}) will be very
381: poor, because transitions between $\lambda = 0$ and $1$ will be
382: extremely rare---typically, $E^*_0$ is shallow and wide, while $E^*_1$
383: is deep and narrow; see Fig.~\ref{energyfig}(a).
384: \begin{figure}
385: \begin{center}%
386: \includegraphics[width=0.8\figwidth]{energy}\\[0.5ex]
387: \includegraphics[width=0.8\figwidth]{wormhole}\\[-1.5ex]
388: \end{center}
389: \caption{\label{energyfig}
390: (a) Schematic representation of $E^*_0(\Gamma)$ (shallow and wide)
391: and $E^*_1(\Gamma)$ (deep and narrow). Conventional Monte Carlo moves
392: between $\lambda = 0$ and $1$ (shown as dashed lines) are nearly always
393: rejected because they lead to large increases in energy.
394: (b) Schematic representation of typical portals for the wormhole moves
395: for the case illustrated in (a), The large ratio of the volume of the
396: unbound ($\lambda = 0$) portal compared to the bound ($\lambda = 1$)
397: portals compensates for the higher energy of the unbound configurations.
398: This results in accepted wormhole moves (dashed lines) between all the
399: portals.
400: }
401: \end{figure}
402: One possible way of remedying this is to treat $\lambda$ as a continuous
403: variable \cite{tidor93,kong96}, providing a suitably interpolated
404: definition of $E_\lambda(\Gamma)$, and allowing Monte Carlo steps with
405: small changes in $\lambda$. In practice, this approach only ``works''
406: if the two endpoints are sufficiently similar, thus limiting this method
407: to the comparison of chemically close molecules.
408:
409: Here we propose an alternative way of carrying out a Monte Carlo
410: simulation of the $E^*_\lambda(\Gamma)$ system.
411: We restrict the standard
412: moves to changes in $\Gamma$ only, and allow changes in $\lambda$ via
413: ``wormhole moves'' \cite{whnote} which connect otherwise disconnected
414: regions of configuration space.
415: This obviates the
416: need to treat (possibly unphysical) intermediate values of $\lambda$,
417: permitting us to compute the free energy differences needed to
418: determine the absolute binding affinity.
419:
420: Assume we have some system defined on a phase space $\Upsilon$ whose
421: equilibrium distribution is proportional to $g(\Upsilon)$. The
422: canonical average of a quantity $X(\Upsilon)$ is defined by
423: \[
424: \langle X\rangle = \frac
425: {\int d\Upsilon\,g(\Upsilon) X(\Upsilon)}
426: {\int d\Upsilon\,g(\Upsilon)}.
427: \]
428: In our application, we make the identification $\Upsilon =
429: [\Gamma,\lambda]$, $\int d\Upsilon = \sum_\lambda \int d\Gamma$, and
430: $g(\Upsilon)=\exp[-\beta E_\lambda(\Gamma)]$.
431:
432: Let us define a set of ``portal functions,'' $w$, $w'$, $w''$, \ldots,
433: on $\Upsilon$, with properties
434: \[
435: 0\le w(\Upsilon) \le 1/v < \infty,
436: \]
437: \[
438: \int d\Upsilon\, w(\Upsilon) = 1,
439: \]
440: where $v$ is a representative phase-space volume of the portal
441: function. A wormhole move consists of the following steps: select a
442: pair of portals $(w, w')$ with probability $p_{ww'}$; reject the move
443: with probability $1 - v w(\Upsilon)$, where $\Upsilon$ is the current
444: state; otherwise, with probability $v w(\Upsilon)$, pick a
445: configuration $\Upsilon'$ with probability $w'(\Upsilon')$; and accept
446: the move to $\Upsilon'$ with probability $P_{ww'}(\Upsilon,\Upsilon')$.
447: If the move is rejected, $\Upsilon$ is retained as the new state.
448: We determine
449: $P_{ww'}(\Upsilon,\Upsilon')$ by demanding that the rate of making
450: transitions from $\Upsilon$ to $\Upsilon'$ via portals $(w,w')$ is
451: balanced by the reverse rate from $\Upsilon'$ to $\Upsilon$ via
452: portals $(w',w)$, i.e.,
453: \[
454: R_{ww'}(\Upsilon, \Upsilon') = R_{w'w}(\Upsilon', \Upsilon)
455: \]
456: (the condition of detailed balance), where the rates are given by
457: \begin{eqnarray*}
458: R_{ww'}(\Upsilon, \Upsilon') &=&
459: p_{ww'} \frac{g(\Upsilon)}
460: {\int d\Upsilon\,g(\Upsilon)}\times\\
461: &&\qquad
462: v w(\Upsilon) w'(\Upsilon') P_{ww'}(\Upsilon,\Upsilon').
463: \end{eqnarray*}
464: A possible solution for the acceptance probability is
465: \[
466: P_{ww'}(\Upsilon,\Upsilon') =
467: \min\biggl(1, \frac{p_{w'w}}{p_{ww'}} \frac{g(\Upsilon')}{g(\Upsilon)}
468: \frac{v'}{v}\biggr),
469: \]
470: where we have made use of the identity $\alpha \min(1, \alpha^{-1}) =
471: \min(1,\alpha)$, which is valid for $\alpha > 0$.
472:
473: In order to apply this move, let us use a specific rather simple form
474: for the portal functions, $w(\Upsilon)$, namely
475: \[
476: w(\Upsilon) = \left\{
477: \begin{array}{l@{\hspace{1em}}l}
478: 1/v, & \mbox{for $\Upsilon \in w$},\\
479: 0, & \mbox{otherwise},
480: \end{array}\right.
481: \]
482: where we now denote a portal by $w$ which defines an arbitrary subset of
483: $\Upsilon$ space of volume $v$. In principle, there is no restriction on the
484: choice of the portals; however, practical considerations, discussed
485: below, dictate how they are chosen. Furthermore, for simplicity, we
486: will assume that the wormhole probabilities are all equal,
487: $p_{ww'} = \mathrm{const}$.
488:
489: Let us now describe the wormhole move in $[\Gamma,\lambda]$ space.
490: Starting with a configuration $[\Gamma, \lambda]$, first pick a
491: random portal $w$. If $[\Gamma, \lambda] \notin w$, then reject the
492: move. Otherwise, pick a random configuration $[\Gamma', \lambda']$
493: uniformly in a randomly chosen portal $w'$, and accept the move to
494: $[\Gamma',
495: \lambda']$ with probability
496: \begin{equation}\label{whacc}
497: \min\biggl(1,
498: \frac{\exp[-\beta E^*_{\lambda'}(\Gamma')]}
499: {\exp[-\beta E^*_\lambda(\Gamma)]}
500: \frac{v'}v
501: \biggr).
502: \end{equation}
503: This differs from the standard Boltzmann acceptance probabilty by
504: the ratio of volumes, $v'/v$, which can be large enough to compensate for
505: the difference in the mean energies in the portals. Indeed, if the
506: mean energy of configurations in $w$ scales as
507: $\beta^{-1} \ln v + \mathrm{const.}$, the acceptance probability,
508: Eq.~(\ref{whacc}), is $O(1)$, thereby allowing moves between
509: shallow wide wells and deep narrow ones; see Fig.~\ref{energyfig}(b).
510: In practice, each portal will occupy one of the energy wells of either
511: $E^*_0(\Gamma)$ or $E^*_1(\Gamma)$. This implies that the $\lambda = 0$
512: portals will permit unrestricted translation and rotation of the
513: ligand and the protein subject to the $V_0$ constraint, and the $\lambda
514: = 1$ portals will allow unrestricted motion of one of the molecules.
515:
516: The wormhole move embodies two concepts which have been used separately
517: in other works: stretching or shrinking $\Gamma$ space when making the
518: move compensates for possibly large energy differences between wells
519: \nocite{miller00}\cite{miller00,zhu02};
520: and jumping between disconnected regions of phase
521: space enables the Markov chain to explore regions of phase space
522: separated by large energy barriers
523: \nocite{voter85}\cite{voter85,senderowitz95}. In the
524: following sections, we will show how these elements may be combined to
525: permit the computation of protein binding affinities with realistic
526: force fields.
527:
528: \section*{Finding the portals}
529:
530: In order for the wormhole method to be practical, we need a reliable way
531: of choosing the portals. We describe this process first in the
532: general case. Let us write $\Gamma = [\Gamma_\mathrm L, \Gamma_\mathrm
533: P]$, where $\Gamma_\mathrm M = [X_\mathrm M, \Xi_\mathrm M]$ represents
534: the state of molecule $\mathrm M$, $X_\mathrm M$ represents its position
535: and orientation, and $\Xi_\mathrm M$ represents its conformation.
536:
537: We assume that $\Xi_\mathrm M$ is expressed in such a way that any
538: constraints on the positions of the atoms in $\mathrm M$ (e.g., bond
539: lengths and bond angles) is implicitly accounted for, so that the
540: dimensionality of $\Xi_\mathrm M$ reflects the number of degrees of
541: conformational freedom, $n_\mathrm M$, for this molecule.
542:
543: Because of the translational and rotational symmetry of the system,
544: certain components of $\Gamma$ are ignorable. We can therefore write
545: $E^*_\lambda(\Gamma) = E^*_\lambda(\Xi_\lambda)$, with
546: \begin{eqnarray*}
547: \Xi_0 &=& [\Xi_\mathrm L,\Xi_\mathrm P],\\
548: \Xi_1 &=& [Y, \Xi_\mathrm L,\Xi_\mathrm P],
549: \end{eqnarray*}
550: where $Y = X_\mathrm L - X_\mathrm P$ denotes the position and
551: orientation of the ligand relative to the protein. The dimensionality
552: of $\Xi_\lambda$ is $n_\lambda$ with $n_0 = n_\mathrm L + n_\mathrm P$
553: and $n_1 = n_0 + 6$.
554:
555: The strategy for determining the portals is to carry out conventional
556: canonical Monte Carlo simulations with $E^*_\lambda(\Xi_\lambda)$
557: separately for $\lambda = 0$ and $1$. For each $\lambda$, we obtain a
558: canonical set of configurations $\{\Xi\}$ (suppressing the $\lambda$
559: subscripts for brevity), to which we fit $n$-dimensional ellipsoids, as
560: follows. First we try to fit a single ellipsoid to $\{\Xi\}$. The
561: center of the ellipsoid is given by the mean configuration
562: $\langle\Xi\rangle$. For each configuration in $\{\Xi\}$, we determine
563: the deviation from the mean, $\delta \Xi = \Xi - \langle\Xi\rangle$, and
564: compute a covariance matrix for the configurations which can be
565: diagonalized as
566: \[
567: \langle \delta\Xi \, \delta\Xi\rangle =
568: \mathsf P \Lambda \mathsf P^\mathrm T,
569: \]
570: where $\mathsf P$ is the matrix of (column) eigenvectors and $\Lambda$
571: is the diagonal matrix of eigenvalues. Because of the properties of the
572: covariance, $\mathsf P$ is real and orthogonal, $\mathsf P^{-1} =
573: \mathsf P^\mathrm T$, and the eigenvalues are real and non-negative. If there
574: are no hidden constraints on the motion, we additionally can assume that
575: the eigenvalues are strictly positive. We find it convenient to define
576: \[
577: \mathsf B = \mathsf P \Lambda^{1/2}, \qquad
578: \mathsf B^{-1} = \Lambda^{-1/2} \mathsf P^\mathrm T,
579: \]
580: so that we can write
581: \[
582: \langle \delta\Xi \, \delta\Xi\rangle = \mathsf B \mathsf B^\mathrm T.
583: \]
584: We take the semi-axes of the ellipsoid to be the columns of $\sqrt{n}
585: \mathsf B$. The multiplier here, $\sqrt{n}$, is chosen to ensure that
586: $O(1)$ of the configurations in $\{\Xi\}$ lie within the
587: ellipsoid. This choice is motivated by considering a symmetric
588: $n$-dimensional Gaussian
589: \[
590: f(\mathbf r) = \frac{\exp(-\frac12 r^2)}{(2\pi)^{n/2}},
591: \]
592: for which we have
593: \[
594: \langle r^2 \rangle = \int r^2 f(\mathbf r)\, d^n\mathbf r = n.
595: \]
596:
597: Ellipsoids are a natural choice to use to fit the set of configurations
598: for the following reasons: (1)~The iso-density contours of the
599: distribution in a harmonic well are ellipsoids. (2)~It is easy to
600: sample points randomly from an ellipsoid. (3)~Conversely, it is easy to
601: test that a point lies inside an ellipsoid. (4)~The volume of an
602: $n$-dimensional ellipsoid is given by
603: \[
604: v_n = \frac{\pi^{n/2}}{(n/2)!} \prod_{i=1}^n a_i,
605: \]
606: where $a_i$ is the length of $i$th semi-axis. Note well the degenerate
607: case of this result, $v_0 = 1$, which is used if the protein and ligand
608: are both rigid ($n_0 = 0$). The sampling and testing of points, (2) and
609: (3), can either be accomplished by transforming with the matrix $\mathsf
610: B$. Alternatively, we can use the simpler Cholesky decomposition of the
611: covariance matrix
612: \[
613: \langle \delta\Xi \, \delta\Xi\rangle =
614: \mathsf C \mathsf C^\mathrm T,
615: \]
616: where $\mathsf C$ is a lower triangular matrix. Both $\mathsf C$ and
617: $\mathsf C^{-1}$ may be computed by direct (non-iterative) methods
618: \cite{golub96}.
619:
620: Having defined an ellipsoid in this way, we test its suitability as a
621: portal by demanding that $O(1)$ of the configurations sampled
622: uniformly from it have energies close to its mean energy $\langle
623: E^*(\Xi)\rangle$. If this test fails, we split $\{\Xi\}$ into two sets
624: according to the sign of $\delta \Xi$ projected along the largest
625: semi-axis of the ellipsoid and construct new trial portals from each
626: of these sets.
627:
628: The ellipsoids that result from this process constitute our portals.
629: When computing the volumes of the portals for use in
630: Eq.~(\ref{whacc}), we need to account for the freedom to place
631: $2-\lambda$ molecules at arbitrary positions and orientations in the
632: volume $V_0$. In practice, this means we multiply the volume of the
633: $\lambda=0$ portals by $\sigma V_0$ where $V_0$ is the translational
634: volume and $\sigma$ is the orientational volume (given below).
635:
636: In order to complete the specification of the portals, we need to
637: describe how $\langle\Xi\rangle$ and $\delta\Xi$ are formed, since, as a
638: consequence of working in the subspace where the molecular constraints
639: are implicitly satisfied, $\Xi$ is not simply a point in $\mathbb R^n$.
640: We wish to represent each ellipsoid on a locally Cartesian space
641: $\mathbb R^n$ which lets us use familiar formulas for defining the
642: ellipsoid. We also demand that the mapping to $\mathbb R^n$ have constant
643: Jacobian in order to maintain detailed balance when sampling from the
644: ellipsoids.
645:
646: We illustrate this by considering the case of the protein and the ligand
647: being made up of $n_\mathrm M+1$ rigid fragments simply connected by
648: $n_\mathrm M$ bonds each of which allow rotation only (i.e., the bond
649: lengths and bond angles are fixed). The relative position and
650: orientation of $\mathrm L$ with respect to $\mathrm P$, $Y$, can be
651: defined in terms of one of the rigid fragments of each molecule.
652: Finally, we use unit quaternions
653: \cite{hamilton47} to represent orientation. Since the quaternions
654: $q$ and $-q$
655: represent the same rotation, orientations are given by an \emph{axis}
656: \cite{mardia99} of the sphere $S^3$.
657:
658: The coordinates making up $\Xi$ are then: (a)~the position component of
659: $Y$, a point in $\mathbb R^3$; (b)~the orientation component of $Y$, an
660: axis of the sphere $S^3$; and (c)~the dihedral angles of the rotatable
661: bonds of the ligand and protein, points on the circle $S^1$. The
662: definition of
663: $\langle\Xi\rangle$ and $\delta\Xi$ is straightforward for (a), since we
664: use the normal arithmetic definitions. For (c), we define the mean for
665: each angle \cite{mardia99} by the direction of the mean of the points on
666: $S^1$ embedded in $\mathbb R^2$. We form the deviation in this case by
667: subtraction modulo $2\pi$ so that the result lies in $[-\pi,\pi]$.
668:
669: To find the mean of the orientations (b), we similarly embed $S^3$ in
670: $\mathbb R^4$. We define the mean orientation \cite{mardia99} as the
671: axis in $\mathbb R^4$ about which the moment of inertia of the sample
672: axes is minimum.
673: To find the deviation of the orientation, we compute
674: the differential rotation $\delta q = q \langle q\rangle^*$ which takes
675: the mean orientation to the sample orientation and we project this into a
676: ``turn'' vector $\mathbf u$ in the unit ball in $\mathbb R^3$ so as to
677: preserve the metric. This is achieved by a generalization of the
678: Lambert azimuthal equal-area projection as described in the Appendix.
679: In this representation, the volume of orientational
680: space (which contributes to the multiplier for the unbound volumes) is
681: $\sigma = \frac43\pi$.
682:
683: Occasionally, the point sampled from $w'$ corresponds to one of the
684: coordinates ``wrapping around,'' i.e., the change in the dihedral lies
685: outside $[-\pi,\pi]$ or the turn $\mathbf u$ lies outside the unit ball.
686: In order to preserve detailed balance, we reject the resulting move.
687:
688: There are four ways in which we can improve the quality of the
689: portals obtained by this method so as to make successful wormhole
690: moves more likely. (1)~The Monte Carlo runs used to obtain the
691: samples from which the portals are defined should begin with an
692: ``annealing'' phase where the temperature is started at some high value
693: and slowly reduced to $T$ at which point we start gathering samples.
694: This allows the Monte Carlo sampling to explore phase space more
695: thoroughly.
696: (2)~Other methods of finding conformational energy minima \cite{head97}
697: can be applied to provide additional starting points for the Monte Carlo
698: runs.
699: (3)~The samples can be supplemented with those
700: obtained by applying those symmetry operations which leave the molecules
701: invariant. In the case of non-chiral ligands, we would also apply a
702: reflection of the ligand in the unbound case since this will leave the
703: energy unchanged. In this way, the portal moves allow all the
704: symmetric variants of the molecules to be explored so that symmetry is
705: included in a systematic way in the computation of the binding affinity.
706: (4)~When forming $\langle \delta\Xi_0 \, \delta\Xi_0\rangle$, we
707: should set the intermolecular cross terms to zero because the
708: conformations of the two molecules are independent when $\lambda = 0$.
709:
710: This method of finding portals depends on the samples ``spanning'' a
711: volume of phase space. This requires that the dimensionality of phase
712: space be sufficiently small and this, in turn, implies the use of an
713: implicit solvation model. In addition the portals can be more
714: reliably found if the ``hard'' degrees of freedom in the molecules are
715: replaced by constraints (e.g., by fixing the bond lengths and bond
716: angles as described above).
717:
718: The method of successively subdividing the samples may lead
719: to suboptimal portals in some cases, for example, by dividing a
720: contiguous set of samples. We have recently experimented with fitting a
721: mixture of Gaussians to the samples using the expectation-maximization
722: (EM) algorithm \nocite{dempster77}\cite{dempster77,verbeek03}.
723: Since this optimizes the
724: fit to all the samples, it usually results in fewer, better, portals.
725: In this case, we are naturally lead to use Gaussians for the portal
726: functions rather than the more restrictive ellipsoids. By enforcing the
727: symmetries of the ligand when making the fits, ligand symmetry can also
728: be included in a rigorous way. These improvements will be described in a
729: subsequent publication.
730:
731: \section*{The free energy calculation}
732:
733: Prior to the calculation of the free energy, we estimate a suitable
734: value of $V_0$, which enters into the definition of the volumes of the
735: $\lambda=0$ portals, by assigning an estimated statistical weight of
736: $v\exp(-\beta \langle E^*\rangle)$ to each portal and estimating
737: \[
738: V_0 \sim \frac
739: {\sum_{\lambda=1} v \exp(-\beta \langle E^*\rangle)}
740: {\sum_{\lambda=0} (v/V_0) \exp(-\beta \langle E^*\rangle)}
741: ,
742: \]
743: where the sums in numerator (denominator) are over the bound (unbound)
744: portals and $v/V_0$ for the unbound portals is the volume of the
745: ellipsoid (multiplied by $\sigma$) and thus does \emph{not} depend on
746: $V_0$.
747: If this estimate for $V_0$ results in the sampling being
748: too heavily weighted toward $\lambda =
749: 0$ or $1$, $V_0$ may be changed and
750: the current values of the sample sums for $K_d$ can be
751: adjusted to account for this change.
752:
753: The choice of the starting configuration for the free energy calculation
754: may introduce some bias in the results. We can remove much of this bias
755: by using the portals to select the starting configuration:
756: select a portal $w$ with probability
757: proportional to its estimated statistical weight; select a
758: configuration, $[\Gamma, \lambda]$, uniformly from this portal; and
759: accept the configuration with probability
760: \[
761: \min\bigl[1, \exp\bigl(-\beta[E^*_\lambda(\Gamma) -
762: \langle E^*\rangle]\bigr)\bigl],
763: \]
764: where $\langle E^*\rangle$ is the canonical average energy over the
765: portal. This procedure is repeated until a configuration is accepted.
766: The bias can be further reduced by running the Monte Carlo calculation
767: for several correlation times, defined by Eq.~(\ref{diffusion}), prior to
768: gathering the data for Eq.~(\ref{kdcanon}).
769:
770: During the course of the free energy calculation, normal and wormhole
771: Monte Carlo moves are mixed. With the normal moves, we only attempt to
772: change $\Gamma$ and, for this reason, it is possible to have the Monte
773: Carlo step size depend on $\lambda$ (with, usually, the step size being
774: larger with $\lambda = 0$).
775: In practice, most ($\sim 90\%$) of the attempted moves are wormhole moves
776: because frequently we have
777: $[\Gamma,\lambda] \notin w$ and the attempted wormhole move is inexpensively
778: rejected.
779:
780: The method is robust in the sense that it does not depend on the
781: particular form of the energy function. In addition, a failure to make
782: transitions between $\lambda = 0$ and $1$ can be detected. This may be
783: because the test $[\Gamma, \lambda] \in w$ never succeeds (i.e., the
784: configuration has been trapped in a new well), or because the acceptance
785: criterion is never met, which indicates that there is a deep well within
786: the well of one of the portals. In both cases, the problem can be
787: corrected by adding a new portal based on recent configurations.
788:
789: \section*{Example}
790:
791: The efficacy of the wormhole method depends on how frequently a
792: configuration lies within one of the portals and how often the jump to
793: the new portal is accepted.
794: \begin{figure}
795: \begin{center}
796: \vspace{1.0ex}
797: \includegraphics[height=9ex]{benzamidine}
798: \vspace{-1.5ex}
799: \end{center}
800: \caption{\label{benzamidine}
801: The structure of $p$-amino-benzamidine.
802: }
803: \end{figure}
804: In order to assess these questions, we
805: have computed the binding affinity of $p$-amino-benzamidine, whose
806: structure is shown in Fig.~\ref{benzamidine},
807: to the digestive enzyme trypsin, at $T = 290\,\mathrm K$.
808: We emphasize that the primary goal of this exercise is to assess how
809: well wormhole moves allow the free energy calculation to converge. For
810: this purpose, we are not so interested in comparing the resulting
811: computed binding affinity to the experimental data since this will
812: depend in large measure on the accuracy of the force field and of the
813: protein structure. Nevertheless, since the convergence will depend on
814: the complexity of the energy ``landscape,'' we use this example to
815: epitomize the binding of a small ligand to a protein.
816:
817: The coordinates of the atoms in
818: trypsin are taken from a trypsin-benzamidine complex, 1BTY
819: \cite{katz95}. At physiological $\mathrm{pH}$, the amidine group is
820: protonated (net charge of $+1$) and, in the complex, it is attracted to
821: a negatively charged aspartate residue in trypsin inhibiting its
822: enzymatic action. We employ the Amber 7 force field
823: \nocite{amber7}\nocite{cornell95}\cite{amber7,cornell95,bayley93}
824: and the GB/SA solvation model
825: \nocite{still90}\nocite{hawkins95}\cite{still90,hawkins95,tsui00}.
826: The protein is taken to be rigid,
827: $n_\mathrm P = 0$, and two bonds of the ligand are allowed to rotate,
828: namely, those connecting the amidine and the amino groups to the
829: benzene ring, $n_\mathrm L = 2$. The published force field
830: \cite{amber7} does not provide a satisfactory torsion for the bond
831: between the benzene ring and the amidine group and this term was
832: determined using Gaussian 98 \cite{gaussian98} as $[-14.2
833: \cos(2 \phi) + 3.3 \cos(4 \phi) + 0.5 \cos(6 \phi)]\,\mathrm{kJ/mol}$,
834: where $\phi$ is the dihedral angle.
835:
836: Five unbound and five bound canonical simulations of 1000 steps each
837: were carried out to find the portals. The resulting configurations
838: were supplemented by those obtained by applying the symmetry operations
839: which leave the ligand invariant. This gave 16 unbound and 8 bound
840: portals. The configurations of the bound portals are all the same
841: ``pose'' of the ligand on the protein and correspond to the symmetries
842: of $p$-amino-benzamidine given by rotating the amino, benzene, and
843: amidine groups by $180^\circ$.
844: The unbound ligand also exhibits 8 symmetries given by rotating the
845: amino and amidine groups by $180^\circ$ relative to the central benzene
846: ring and by including the mirror images. However since the bond
847: parameters allow partially free rotation of the amidine group, each
848: symmetric set of configurations is represented by two portals.
849:
850: We estimated a suitable value for $V_0$ of $0.39\times10^{-18}\,
851: \mathrm{m^3}$ based on the volumes and the mean energies of the computed
852: portals. During the binding affinity calculation, wormhole moves were
853: attempted on 90\% of the steps (the other 10\% were standard Monte Carlo
854: moves). Of these attempted wormhole moves, 97\% failed (inexpensively)
855: because the configuration was not in the chosen $w$. Of the remaining
856: 3\%, about 60\% lead to a successful move, of which about 40\% involved
857: switching from $\lambda =0$ to $1$ and \emph{vice versa}. The major
858: computational cost in the free energy calculation is the evaluation of
859: the bound energies. In this example, the wormhole moves required less
860: than 3 bound energy evaluations, on average, to effect the transition
861: from a bound configuration to an unbound one and back. This enables an
862: accurate estimate to be made of the ratio of the averages in
863: Eq.~(\ref{kdcanon}) which after $5\times10^6$ steps yields $\mathrm pK_d
864: = 7.99 \pm 0.01$. The error estimate is derived in the next section
865: and represents a 2\% relative error in $K_d$.
866:
867: This example illustrates that the method is effective at allowing
868: sufficient transitions between the bound and unbound states to enable
869: the binding affinity of protein-ligand systems to be accurately
870: computed. We note that our computed binding affinity differs from the
871: experimental results of $5.1$ to $5.2$
872: \nocite{maresguia65}\cite{maresguia65,schwarzl02}.
873: This discrepancy may be accounted for by modest, $\sim20\%$, errors in
874: the force field.
875:
876: \section*{Error analysis}
877:
878: In order to assess the errors in the computation of $\mathrm pK_d$ in
879: more detail, we carried out 10 independent runs similar to the one
880: described in previous section. Each of these used the same portals
881: and the same value of $V_0$. We computed cumulative estimates for
882: $\mathrm pK_d$ based on the first $s$ steps of the Markov chains. When
883: forming the averages in Eq.~(\ref{kdcanon}) we sample every 100th step.
884: (As we shall see, there is a high degree of correlation within 100
885: steps; thus there would be little improvement in the estimate of
886: $\mathrm pK_d$ by sampling more frequently.)
887: \begin{figure}
888: \begin{center}
889: \vspace{-0.5ex}
890: \includegraphics[width=\figwidth]{cumulative}
891: \vspace{-1.5ex}
892: \end{center}
893: \caption{\label{cumulative}
894: Cumulative estimates $\mathrm pK_d(s)$ obtained by sampling every 100th
895: step from the first $s$ steps of 5 independent Monte Carlo runs.
896: The dashed lines shows
897: convergence as $1/\sqrt s$ to the mean value of $7.99$.}
898: \end{figure}
899: The results for 5 of these runs are shown in Fig.~\ref{cumulative}. The
900: convergence toward the mean is as $1/\sqrt s$; after $5\times10^6$
901: steps, the error in $\mathrm pK_d$ has been reduced to about $\pm 0.01$.
902:
903: For the purposes of further analysis, let us assume that the computation
904: is carried out with $E_\lambda^*(\Gamma) = E_\lambda(\Gamma)$ so that
905: all samples have the same statistical weight. The probability that the
906: system is in the bound (resp.\ unbound) state is $p = \langle
907: \delta_{\lambda1} \rangle$ (resp.\ $q = 1-p = \langle \delta_{\lambda0}
908: \rangle$) and Eq.~(\ref{kdcanon}) becomes
909: \begin{equation}\label{kdcanona}
910: K_d = \frac1{V_0}\frac qp.
911: \end{equation}
912: The determination of $K_d$ is then
913: equivalent to estimating $q/p$ by taking the ratio of the
914: number of unbound and bound steps in the Monte Carlo simulation. How
915: well this estimate converges depends, naturally, on the ``correlation
916: time'' of the system. If wormhole moves rarely cause the system to
917: switch between bound and unbound states, the correlation time is large
918: and the convergence will be slow.
919:
920: In order to make these ideas quantitative, we define the
921: $\lambda$-correlation function,
922: \[
923: C_t =
924: \langle (\lambda_s - p)
925: (\lambda_{s+t} - p)
926: \rangle_s,
927: \]
928: where $\lambda_s$ is the value of $\lambda$ at simulation step $s$ and
929: $\langle\ldots\rangle_s$ denotes an average over steps.
930: \begin{figure}
931: \begin{center}
932: \vspace{-0.5ex}
933: \includegraphics[width=\figwidth]{correlation}
934: \vspace{-1.5ex}
935: \end{center}
936: \caption{\label{correlation}
937: The $\lambda$-correlation function $C_t$. The curves show $C_t$ for
938: $V_0/(0.39\times10^{-18} \,\mathrm{m^3}) =$ (a)~$\frac1{10}$,
939: (b)~$\frac13$, (c)~$1$, (d)~$3$, and (e)~$10$.
940: The ratio $q/p$ varies correspondingly (from
941: approximately $\frac18$ to $\frac{100}8$). The correlation times are
942: (a)~$329$, (b)~$346$, (c)~$352$, (d)~$382$, and (e)~$415$.
943: In determining the
944: correlation times, we limit the sum in Eq.~(\ref{diffusion}) to $0< t
945: \le 2500$ in order to avoid including the small but noisy terms for
946: large~$t$.}
947: \end{figure}
948: Figure \ref{correlation} shows $C_t$ for several different values of
949: $V_0$. (From Eq.~(\ref{kdcanona}), we have $q/p\propto V_0$.) In a
950: Markov chain of length $s$, the expected number of bound states is $ps$,
951: while, for $s\rightarrow\infty$, the variance in the number of bound
952: states is $2Ds$, where
953: \begin{equation}\label{diffusion}
954: D = \frac12 C_0 + \sum_{t>0} C_t = \frac12 C_0 \tau.
955: \end{equation}
956: This provides us with the definition of the correlation time, $\tau$.
957: From Fig.~\ref{correlation}, we see that $C_t$ decays approximately
958: exponentially so that the sum converges.
959:
960: We can compare the Monte Carlo simulation to a simpler Markov process of
961: independent trials, e.g., tossing a coin where the probability of heads
962: is $p$. In this case, we have $C_{t>0} = 0$ and $D=\frac12 C_0 =
963: \frac12 pq$. In the limit $s\rightarrow\infty$, the relative errors in
964: estimating $p$ for the Monte Carlo simulation will be the same as that
965: for $s/\tau$ tosses of the coin. We can illustrate this by making 5
966: independent simulations of a coin tossing experiment to match the data
967: in Fig.~\ref{cumulative}, for which $p = 0.443$ and $\tau = 352$.
968: \begin{figure}
969: \begin{center}
970: \vspace{-0.5ex}
971: \includegraphics[width=\figwidth]{cointoss}
972: \vspace{-1.5ex}
973: \end{center}
974: \caption{\label{cointoss}
975: Estimation of the bias of a coin with $p = 0.443$. Five independent
976: runs are shown. In each run, the coin was tossed $5\times 10^6/\tau =
977: 14200$ times, where $\tau$ is the correlation time for
978: Fig.~\ref{correlation}(c), and the cumulative value of $h_n/t_n$ is
979: recorded. The expected value, $p/q$, is shown as a dashed line. The
980: axes have been adjusted to allow the figure to be directly compared with
981: Fig.~\ref{cumulative}.}
982: \end{figure}
983: The results are given in Fig.~\ref{cointoss}, where we plot $h_n/t_n$
984: against $n$ where $h_n$ (resp.\ $t_n$) is the number of heads (resp.\
985: tails) after $n$ throws. As expected, Figs.~\ref{cumulative} and
986: \ref{cointoss} both exhibit the same convergence behavior.
987:
988: Since the distribution of outcomes $h_n$ in the coin tossing experiment
989: is the binomial distribution, the mean and variance for $l_n =
990: \ln(h_n/t_n)$ can be calculated, yielding
991: \[
992: \langle l_n \rangle
993: = \ln\biggl(\frac pq \biggr) +
994: \frac1{2n}\biggl(\frac pq - \frac qp\biggr) +
995: O(n^{-2}),
996: \]
997: \[
998: \langle (l_n - \langle l_n \rangle)^2 \rangle
999: = \frac1{npq} + O(n^{-2}).
1000: \]
1001: The standard error in the estimate of $\mathrm pK_d$ from a Monte Carlo
1002: run with $s$ steps is found by substituting $n=s/\tau$ and scaling by
1003: $\ln 10$ (since $\mathrm pK_d$ is defined in terms of the common
1004: logarithm) to give
1005: \begin{equation}\label{err}
1006: \Delta \mathrm pK_d \approx \frac1{\ln 10} \sqrt{\frac{\tau}{spq}}.
1007: \end{equation}
1008: In the case of the simulations shown in Fig.~\ref{cumulative}, we find
1009: $\Delta \mathrm pK_d \approx 0.01$ consistent with the data in the
1010: figure. From Fig.~\ref{correlation}, we see that $\tau$ is weakly
1011: dependent on $p$. Thus for a given $s$, we minimize the error by
1012: choosing $p = q = \frac12$. Alternatively, we may wish to minimize the
1013: error for a given amount of computational effort. If bound steps are
1014: $f$ times more expensive to carry out than unbound steps, the error is
1015: minimized by taking $q/p = \sqrt f$, i.e., $p = 1/(1+\sqrt f)$.
1016:
1017: \section*{Discussion}
1018:
1019: The important aspects of this method that enable us to compute the
1020: binding affinity of a ligand to a protein are: (1)~formulation in the
1021: extended $[\Gamma,\lambda]$ space, which allows the unbound and bound
1022: systems to be treated as a single canonical system and $K_d$ to be
1023: expressed as the ratio of canonical averages, Eq.~(\ref{kdcanon});
1024: (2)~the wormhole move which allows transitions between the bound and
1025: unbound systems; and (3)~the use of an implicit solvation model which
1026: reduces the number of degrees of freedom and so allows portals to be
1027: identified.
1028:
1029: A method similar to ours is ``simulated mutational equilibration,''
1030: \cite{senderowitz97} which employs an extended phase space, uses an
1031: implicit solvation model, and allows jumps between different systems.
1032: However, in place of the wormhole move, this work employed a more
1033: restrictive ``jumping between wells'' move, which limited its
1034: applicability to computing the difference in the binding affinities for
1035: two enantiomers.
1036: In contrast, our wormhole method can be used to compute directly the
1037: binding affinity of a ligand with several rotatable bonds to a protein
1038: target. This allows us to study a wide range of interesting drug-like
1039: ligands.
1040:
1041: In the special case of binding rigid molecules (for which we have
1042: $E^*_0(\Gamma) = 0$), wormhole Monte Carlo is
1043: isomorphic to a grand canonical Monte Carlo
1044: simulation \cite{adams75}. Consider a
1045: grand canonical system with a single protein molecule and with ligand
1046: molecules at a fixed chemical potential $\mu$; we make the additional
1047: restriction that the ligand-ligand interaction energy is infinite, so that the
1048: system can only accommodate 0 or 1 ligand molecule. For the wormhole
1049: simulation, we specify the bound portal to include the full simulation
1050: volume with arbitrary orientation; the unbound port is degenerate and
1051: corresponds merely to specifying $\lambda = 0$. The wormhole
1052: moves from $\lambda = 0$ to $1$ and \emph{vice versa}
1053: correspond to particle insertion and deletion in
1054: the grand canonical simulation; and we can verify the acceptance
1055: probabilities are the same.
1056:
1057: The application described here can be extended by allowing a greater
1058: degree of flexibility for the protein and the ligand. This permits the
1059: treatment of side-chain rotation and ligand-induced loop movement on the
1060: part of the protein and the treatment of flexible rings for the ligand.
1061: Many standard Monte Carlo techniques can be used with this method, if
1062: appropriate---preferential sampling, early rejection, force bias, etc.
1063: Wormhole moves could also be used to treat other discrete, or nearly
1064: discrete, transitions. Examples are: the treatment of molecules, such
1065: as cyclohexane, which can assume distinct conformations; discrete sets
1066: of side chain rotations in the protein; protonation and tautomerization
1067: states for either molecule. In each of these cases, the acceptance
1068: probability for the transitions should account for the free energy
1069: difference between the discrete states.
1070:
1071: \section*{Acknowledgment}
1072:
1073: This work was supported, in part, by the U.S. Army Medical Research and
1074: Materiel Command under Contract No.\ DAMD17-03-C-0082.
1075:
1076: \appendix
1077: \ifpreprint
1078: \subsection*{Appendix: Generalized Lambert projection}
1079: \else
1080: \section*{Appendix: Generalized Lambert projection}
1081: \fi
1082:
1083: The Lambert azimuthal equal-area projection projects a point on $S^2$ to
1084: a point on a disk in $\mathbb R^2$. Here, we will generalize this to
1085: arbitrary dimensions, i.e., we will find a projection from $S^n$ to a
1086: ball in $\mathbb R^n$. (Thus the circle $S^1$ is projected into a
1087: line segment; the surface of a sphere $S^2$ is project into a disk;
1088: $S^3$ is projected into a sphere; etc.) The projection is azimuthal, so
1089: that directions from the pole are preserved in the projected space.
1090:
1091: The sequence $S^n$ can be defined recursively by
1092: \begin{eqnarray*}
1093: S^0 &=& [\pm1], \\
1094: S^n &=& [\cos\theta, \sin\theta \, S^{n-1}],
1095: \end{eqnarray*}
1096: where $\theta$ is the colatitude and $0 \le\theta\le\pi$.
1097: The area of $S^n$ lying between $\theta$ and $\theta + d\theta$ is
1098: therefore
1099: \[
1100: dA = a_{n-1} \sin^{n-1}\theta \,d\theta,
1101: \]
1102: where $a_n$ is the area of $S^n$. We project that portion of $S^n$
1103: with colatitude in $[0,\theta]$ to a ball in $\mathbb R^n$ of radius
1104: $t$. Equating the measures (area on $S^n$ and volume in $\mathbb R^n$),
1105: we obtain
1106: \[
1107: v_n t^n = a_{n-1} \int_0^\theta \sin^{n-1}\theta'\,d\theta'
1108: \]
1109: where $v_n$ is the volume of a unit ball in $\mathbb R^n$. Using the
1110: relation,
1111: \[
1112: a_{n-1} = \left.\frac{d(v_n t^n)}{dt}\right|_{t=1} = n v_n,
1113: \]
1114: we obtain
1115: \begin{equation}\label{genlambert}
1116: t = \biggl(n\int_0^\theta \sin^{n-1}\theta'\,d\theta'\biggr)^{1/n}.
1117: \end{equation}
1118: The general ``equal-area'' projection is given by the
1119: mapping
1120: \[
1121: [\cos\theta, \sin\theta \, S^{n-1}] \rightarrow \mathbf t,
1122: \]
1123: where $\mathbf t \in \mathbb R^n$, $\mathbf{\hat t}$ is given by the
1124: point on $S^{n-1}$, and $t$ is given by Eq.~(\ref{genlambert}). Some
1125: special cases of Eq.~(\ref{genlambert}) are
1126: \[
1127: t = \left\{
1128: \begin{array}{l@{\hspace{1em}}l}
1129: \theta, & \mbox{for $n=1$,}\\
1130: 2\sin\frac12\theta, & \mbox{for $n=2$,}\\
1131: \bigl[\frac34(2\theta - \sin2\theta)\bigr]^{1/3}, & \mbox{for $n=3$,}\\
1132: 2\sin\frac12\theta\bigl[\frac13(1+2\cos^2 \frac12\theta)\bigr]^{1/4},
1133: & \mbox{for $n=4$.}\\
1134: \end{array}
1135: \right.
1136: \]
1137: The case $n=1$ corresponds to unwrapping a circle onto a line; and $n=2$
1138: gives the Lambert azimuthal equal-area projection. We are interested in
1139: the case $n=3$ as a way of mapping orientations from unit quaternion
1140: space to $\mathbb R^3$. A quaternion $q = [\cos\theta,
1141: \sin\theta \,\mathbf{\hat u}]$ represents a rotation of $\psi = 2\theta$
1142: about $\mathbf{\hat u}$. Noting that $q$ and $-q$ represent the same
1143: rotation, we may make the restriction $0 \le \theta \le \frac12\pi$ .
1144: We choose, therefore, to map this hemisphere of $S^3$ onto $\mathbf u$
1145: in the unit ball, by making the substitutions $\theta = \frac12\psi$ and
1146: $t = \bigl(\frac34\pi\bigr)^{1/3}u$ to give
1147: \[
1148: u = \biggl(\frac{\psi - \sin\psi}\pi\biggr)^{1/3}.
1149: \]
1150: (The other hemisphere, corresponding to $\frac12\pi<\theta\le\pi$, maps
1151: onto the shell $1 < u \le 2^{1/3}$.) We call this $\mathbf u$
1152: representation of orientations ``turn space.''
1153:
1154: \bibliography{free,wh}
1155:
1156: \end{document}
1157: