q-bio0410018/vdwl4.tex
1: % Group addresses by affiliation; use superscriptaddress for long
2: % author lists, or if there are many overlapping affiliations.
3: % For Pdhys. Rev. appearance, change preprint to twocolumn.
4: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
5: %  Add 'draft' option to mark overfull boxes with black boxes
6: %  Add 'showpacs' option to make PACS codes appear
7: %  Add 'showkeys' option to make keywords appear
8: %\documentclass[aps,pre,preprint,superscriptaddress,showpacs,floatfix]{revtex4}
9: \documentclass[aps,prl,twocolumn,superscriptaddress,floatfix]{revtex4}
10: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
11: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}
12: \usepackage{latexsym,hyperref,color,amssymb,graphicx}
13: % You should use BibTeX and apsrev.bst for references
14: % Choosing a journal automatically selects the correct APS
15: % BibTeX style file (bst file), so only uncomment the line
16: % below if necessary.
17: %\bibliographystyle{h-physrev}
18: \newcommand{\beq}{\begin{equation}}
19: \newcommand{\eeq}{\end{equation}}
20: \newcommand{\bea}{\begin{eqnarray}}
21: \newcommand{\eea}{\end{eqnarray}}
22: \newcommand{\bean}{\begin{eqnarray*}}
23: \newcommand{\eean}{\end{eqnarray*}}
24: \newcommand{\bei}{\begin{itemize}}
25: \newcommand{\eei}{\end{itemize}}
26: \newcommand{\ben}{\begin{enumeration}}
27: \newcommand{\een}{\end{enumeration}}
28: \newcommand{\nn}{\nonumber}
29: \newcommand{\thalf}{\textstyle{ {1 \over 2} } }
30: \newcommand{\shalf}{\scriptstyle{\frac{1}{2}}}
31: \newcommand{\sshalf}{\scriptscriptstyle{\frac{1}{2}}}
32: \newcommand{\tcol}[2]{\textcolor{#1}{#2}}
33: \newcommand{\tcr}[1]{\textcolor{red}{#1}}
34: \newcommand{\tcb}[1]{\textcolor{blue}{#1}}
35: \newcommand{\tcg}[1]{\textcolor{green}{#1}}
36: \newcommand{\tcy}[1]{\textcolor{yellow}{#1}}
37: \newcommand{\tcc}[1]{\textcolor{cyan}{#1}}
38: \newcommand{\tcm}[1]{\textcolor{magenta}{#1}}
39: \definecolor{darkorange}{rgb}{.6,.2,.0}
40: \newcommand{\tcdo}[1]{\textcolor{darkorange}{#1}}
41: \definecolor{darkgreen}{rgb}{0.0,0.7,0.0}
42: \newcommand{\tcdg}[1]{\textcolor{darkgreen}{#1}}
43: \newcommand{\tcp}[5]{\definecolor{#1}{rgb}{#2,#3,#4}\textcolor{#1}{#5}}
44: \begin{document}
45: 
46: \title{Rydberg-London Potential for Diatomic Molecules
47: and Unbonded Atom Pairs}
48: 
49: \author{Kevin Cahill}
50: \email{cahill@unm.edu}
51: \affiliation{Department of Physics and Astronomy, 
52: University of New Mexico, Albuquerque, NM 87131}
53: \affiliation{Center for Molecular Modeling, 
54: Center for Information Technology, 
55: National Institutes of Health, 
56: Bethesda, Maryland 20892-5624}
57: \author{V.~Adrian Parsegian}
58: \email{aparsegi@helix.nih.gov}
59: \affiliation{National Institute of Child Health
60: and Human Development, National Institutes of Health, 
61: Bethesda, Maryland 20892-0924}
62: 
63: \date{\today}
64: 
65: \begin{abstract}
66: We propose and test a pair potential that
67: is accurate at all relevant distances
68: and simple enough 
69: for use in large-scale computer simulations.  
70: A combination of the Rydberg potential from spectroscopy 
71: and the London inverse-sixth-power energy, 
72: the proposed form fits 
73: spectroscopically determined potentials
74: better than the Morse, Varnshi,
75: and Hulburt-Hirschfelder potentials
76: and much better than the 
77: Lennard-Jones and harmonic potentials.
78: At long distances, it 
79: goes smoothly to the correct London force 
80: appropriate for gases and preserves 
81: van der Waals's ``continuity of the gas and liquid states," 
82: which is routinely violated by coefficients assigned 
83: to the Lennard-Jones 6-12 form.
84: \end{abstract}
85: 
86: %\pacs{82.35.Pq,34.20.Cf,77.84.Jd}
87: 
88: \maketitle
89: There are at least three classes of 
90: interatomic potentials.
91: Commercial codes
92: use the Lennard-Jones 
93: \beq
94: V_{LJ}(r) = |V(r_0)| \, \left[ \left( \frac{r_0}{r} \right)^{12}
95: - 2 \, \left( \frac{r_0}{r} \right)^6 \right]
96: \label{LJ}
97: \eeq
98: and harmonic 
99: \beq
100: V_H(r) = V(r_0) 
101: + \frac{(r - r_0)^2}{2} \, 
102: \frac{d^2 V(r_0)}{dr^2} 
103: \label{Ha}
104: \eeq
105: forms, which
106: are accurate
107: near the minimum at \(r=r_0\)\@.
108: But this first class of potentials may be too simple
109: for complex materials
110: away from from equilibrium.
111: \par
112: The Morse~\cite{Morse1929}
113: \beq
114: V_M(r) = |V(r_0)| \, \left[
115: \left( 1 - e^{-\kappa \, x} \right)^2
116: - 1 \right]
117: \label{Morse}
118: \eeq
119: (\(x = r - r_0\), \( \kappa = \sqrt{k_e/(2|V(r_0)|)} \)),
120: Varnshi~\cite{Varnshi1957}
121: \beq
122: V_V(r) = |V(r_0)| \, \left[
123: \left( 1 - \frac{r_0}{r} \, e^{- \beta \, (r^2 - r_0^2)} 
124: \right)^2
125: - 1 \right],
126: \label{Varnshi}
127: \eeq
128: (\( \beta = (\kappa r_0^2 -1)/(2r_0^2) \)), 
129: and Hulburt-Hirschfelder~\cite{Hulbert1961}
130: \beq
131: V_{HH}(r) = V_M(r) + |V(r_0)| \, 
132: c'\kappa^3x^3\,e^{-2\kappa x}(1+\kappa b'x)
133: \label{HH}
134: \eeq
135: (for \(b'\), \(c'\), see ref.\cite{Vanderslice1961})
136: potentials represent
137: a second class of potentials~\cite{Vanderslice1961}
138: accurate over a wider
139: range of distances.
140: \par
141: Quantum 
142: chemists~\cite{Scoles1984,Aziz1993,Meath2001,Tang2003,Varandas2004}
143: have derived a third class 
144: that reproduce spectroscopic
145: and thermodynamic data with impressive fidelity.
146: But the potentials of this class
147: involve many parameters and
148: may be too cumbersome for use in large-scale simulations.
149: \par
150: We propose and test a form
151: \beq
152: V(r) = a e^{-b \, r}\,( 1 - c \, r ) 
153: - \frac{d}{r^6 + e \, r^{-6}}
154: \label{phenpot}
155: \eeq
156: that is nearly as accurate as the class-3 potentials 
157: but simpler than many class-2 potentials.
158: It is a combination of the Rydberg formula 
159: used in spectroscopy 
160: and the London formula for pairs of atoms.
161: In Eq.(\ref{phenpot}), the terms involving  \(a\), \(b\), and \(c\) 
162: were proposed by Rydberg~\cite{Rydberg1931}
163: to incorporate spectroscopic data,
164: but were largely ignored until 
165: recently~\cite{Ferrante1991,Murrell1984}.
166: The constant \(d = C_6\)
167: is the coefficient of the London tail.
168: The new term \( e \, r^{-6} \) cures
169: the London singularity.
170: As \(r \to 0\),
171: \( V(r) \to a\), finite;
172: as \(r \to \infty\), \(V(r)\) 
173: approaches the London term,
174: \(V(r) \to - d/r^6 = -C_6/r^6\)\@.
175: In a perturbative analysis~\cite{Cahill2003},
176: the \(a, b, c\) terms 
177: arise in first order, and the \(d\) term
178: in second order. 
179: \par
180: \begin{figure}
181: \centering
182: \input RKRH2
183: \caption{The hybrid form \(V\)
184: (with \(a = 53.8\) eV,
185: \(b = 2.99\) \AA\(^{-1}\),
186: \(c = 2.453\) \AA\(^{-1}\),
187: \(d = C_6 = 3.884\) eV\,\AA\(^6\),
188: and \(e = 47.6\) \AA\(^{12}\))
189: (solid, red) 
190: fits the RKR spectral points for 
191: the ground state of
192: molecular hydrogen (pluses, blue)
193: and gives the correct London tail for \(r > 3\) \AA\@.
194: The Lennard-Jones \(V_{LJ}\) (dashes, green) and harmonic
195: \(V_H\) (dots, magenta) forms fit
196: only near the minimum.}
197: \label{rkr}
198: \end{figure}
199: \begin{figure}
200: \centering
201: \input H2REV
202: \caption{The hybrid form \(V\)
203: (Eq.(\ref{phenpot}) as in Fig.~\ref{rkr},
204: solid, red) 
205: fits the RKR spectral points for 
206: the ground state of
207: molecular hydrogen (pluses, blue).
208: The Morse \(V_{M}\) (Eq.(\ref{Morse}), dashes, green),
209: Varnshi \(V_V\) (Eq.(\ref{Varnshi}), dots, magenta), and 
210: Hulburt-Hirschfelder
211: \(V_{HH}\) (Eq.(\ref{HH}), dot-dash, cyan) 
212: are too low for
213: \(1.5 < r < 3\) \AA\@.}
214: \label{h2rev}
215: \end{figure}
216: \begin{figure}
217: \centering
218: \input RKRN2
219: \caption{For the ground state of molecular nitrogen,
220: the proposed potential \(V\) of Eq.(\ref{phenpot}) 
221: (with \(a = 3.25\) keV,
222: \(b = 4.30\) \AA\(^{-1}\),
223: \(c = 1.1906\) \AA\(^{-1}\),
224: \(d = 14.1\) eV\,\AA\(^6\), and
225: \(e = 27.1\) \AA\(^{12}\))
226: (solid, red) 
227: fits the RKR spectral points (pluses, blue)
228: and the FO points (x's, cyan)\@.
229: \(V_{LJ}\) and \(V_H\) as in Fig.~\ref{rkr}\@.}
230: \label{rkrn2}
231: \end{figure}
232: \begin{figure}
233: \centering
234: \input N2REV
235: \caption{For the ground state of molecular nitrogen,
236: the hybrid form \(V\) (Eq.(\ref{phenpot}),
237: as in Fig.~\ref{rkrn2}, solid, red) 
238: fits the RKR spectral points (pluses, blue)
239: and the FO points (x's, black)\@.
240: \(V_{M}\), \(V_V\), \&
241: \(V_{HH}\) as in Fig.~\ref{h2rev}.}
242: \label{n2rev}
243: \end{figure}
244: \begin{figure}
245: \centering
246: \input RKRO2
247: \caption{For the ground state of molecular oxygen,
248: the potential \(V\) of Eq.(\ref{phenpot}) 
249: (with \(a = 3.61\) keV,
250: \(b = 4.48\) \AA\(^{-1}\),
251: \(c = 1.05\) \AA\(^{-1}\),
252: \(d = 16.08\) eV\,\AA\(^6\), and
253: \(e = 58.4\) \AA\(^{12}\))
254: (solid, red) fits
255: the RKR points (pluses, blue)\@.
256: \(V_{LJ}\) and \(V_H\) as in Figs.~\ref{rkr} 
257: \& \ref{rkrn2}\@.}
258: \label{rkro2}
259: \end{figure}
260: \begin{figure}
261: \centering
262: \input O2REV
263: \caption{For the ground state of molecular oxygen,
264: the hybrid form \(V\) (Eq.(\ref{phenpot}),
265: as in Fig.~\ref{rkro2}, solid, red) 
266: fits the RKR spectral points (pluses, blue)\@.
267: \(V_{M}\), \(V_V\), \&
268: \(V_{HH}\) as in Fig.~\ref{h2rev}.}
269: \label{o2rev}
270: \end{figure}
271: \begin{figure}
272: \centering
273: \input AR2374
274: \caption{The potential \(V\) (Eq.(\ref{phenpot}) 
275: for the Ar-Ar ground state  
276: with \(a = 1720\) eV,
277: \(b = 2.6920\) \AA\(^{-1}\),
278: \(c = 0.2631\) \AA\(^{-1}\),
279: \(d = 37.943\) eV\,\AA\(^6\), and
280: \(e = 177588\) \AA\(^{12}\))
281: (solid, red) fits
282: the Ar\(_2\) Aziz potential (pluses, blue)
283: with the correct London tail.
284: When matched at the minimum,
285: the Lennard-Jones form \(V_{LJ}\) 
286: (Eq.(\ref{LJ}) with \(r_0 = 3.757\) \AA\ 
287: and \(V(r_0) = -0.01234\) eV) (dashes, green) 
288: is too low for \(r > 4\) \AA.}
289: \label{ar237}
290: \end{figure}
291: \begin{figure}
292: \centering
293: \input AR24
294: \caption{Positive potential \(V\) (Eq.(\ref{phenpot}) 
295: as in Fig.~\ref{ar237}),
296: Ar\(_2\) Aziz potential (pluses, blue),
297: L-J form \(V_{LJ}\) (Eq.(\ref{LJ})
298: as in Fig.~\ref{ar237})\@.}
299: \label{ar2}
300: \end{figure}
301: \begin{figure}
302: \centering
303: \input KR2374
304: \caption{The potential \(V\) (Eq.(\ref{phenpot})
305: for the Kr-Kr ground state  
306: with \(a = 2499\) eV,
307: \(b = 2.5249\) \AA\(^{-1}\),
308: \(c = 0.2466\) \AA\(^{-1}\),
309: \(d = 78.214\) eV\,\AA\(^6\), and
310: \(e = 199064\) \AA\(^{12}\))
311: (solid, red) fits
312: the Kr\(_2\) Aziz points (pluses, blue)
313: with the correct London tail.
314: When matched at the minimum,
315: the Lennard-Jones form \(V_{LJ}\) 
316: (Eq.(\ref{LJ}) with \(r_0 = 4.008\) \AA\ 
317: and \(V(r_0) = -0.017338\) eV) (dashes, green) 
318: is too low for \(r > 4\) \AA.}
319: \label{kr237}
320: \end{figure}
321: \begin{figure}
322: \centering
323: \input KR24
324: \caption{Positive potential \(V\) (Eq.(\ref{phenpot}) 
325: as in Fig.~\ref{kr237}),
326: Kr\(_2\) Aziz points (pluses, blue),
327: L-J form \(V_{LJ}\) (Eq.(\ref{LJ})
328: as in Fig.~\ref{kr237})\@.}
329: \label{kr2}
330: \end{figure}
331: \par
332: To test whether the 
333: hybrid \(V(r)\)
334: can represent covalent bonds
335: far from equilibrium,
336: we used Gnuplot~\cite{Gnuplot}
337: to fit Eq.(\ref{phenpot}) 
338: to empirical potentials for % Remark: we had 'of,' not 'for.'
339: molecular H\(_2\), N\(_2\), and O\(_2\)
340: obtained from spectroscopic 
341: data~\cite{Weissman1963,Krupenie1977,Krupenie1972}
342: by the RKR (Rydberg~\cite{Rydberg1931}, Klein~\cite{Klein1932},
343: Rees~\cite{Rees1947}) method,
344: setting \(d\) equal to the London values.
345: Figure~\ref{rkr} shows
346: that the hybrid potential of Eq.(\ref{phenpot}) (solid, red)
347: goes through the RKR points for H\(_2\) (pluses, blue)
348: from 0.5 to 4 \AA\@.
349: Fitted to the minimum,
350: the harmonic potential (\ref{Ha}) 
351: (dashes, green)
352: and the Lennard-Jones potential (\ref{LJ})
353: (dashes, blue)
354: are accurate only
355: near the minimum at 0.74 \AA\@.
356: Figures~\ref{rkrn2} \& \ref{rkro2}
357: show that the hybrid potential \(V(r)\)
358: fits the N\(_2\) and O\(_2\) 
359: RKR points~\cite{Krupenie1977,Krupenie1972}
360: and the N\(_2\) first-order (FO)
361: points~\cite{Ermler1982}
362: between 1 and 1.8 \AA\@.
363: The Lennard-Jones and harmonic potentials
364: of Eqs.(\ref{LJ}--\ref{Ha}) fit only near the minima.
365: \par
366: How does the hybrid form compare
367: with the class-2 potentials
368: of Eqs.(\ref{Morse}--\ref{HH})?
369: Figures~\ref{h2rev}, \ref{n2rev}, \&
370: \ref{o2rev} show that for 
371: \( 1.4 < r < 2 \) \AA\ 
372: the hybrid \( V(r) \)
373: is closer than (\ref{Morse}--\ref{HH})
374: to the RKR points.
375: A useful estimate of how well a 
376: particular potential \( V_P(r) \)
377: fits \(N\) data points \( V_D(r_i) \)
378: is the dimensionless error 
379: \beq
380: \delta = \frac{1}{|V_D(r_0)|} \,
381: \left[ \frac{1}{N} \,
382: \sum_{i = 1}^N 
383: \left( V_P(r_i) - V_D(r_i) \right)^2 \,
384: \right]^{1/2}.
385: \label{delta}
386: \eeq
387: For \(H_2\),
388: \(N_2\), and \(O_2\), the average
389: error \(\delta\) was 59.7 for \(V_H\),
390: 49.3 for \(V_{LJ}\),
391: 0.037 for \(V_M\),
392: 0.031 for \(V_V\), 0.021 for \(V_{HH}\),
393: and 0.0044 for the hybrid \(V\)\@.
394: The hybrid form \(V\) is five times more
395: accurate than the
396: class-2 potentials (\ref{Morse}--\ref{HH})
397: and four orders of magnitude more accurate 
398: than the class-1 potentials (\ref{LJ}--\ref{Ha})\@.
399: \par
400: Can \(V(r)\) also represent 
401: weak noncovalent bonds?
402: Using Gnuplot, we fitted 
403: \(a\), \(b\), \(c\), and \(e\) in Eq.(\ref{phenpot})
404: to Aziz's accurate HFDID1 potential for 
405: Ar\(_2\)~\cite{Aziz1993} and HFD-B potential
406: for Kr\(_2\)~\cite{Aziz1986}
407: and set \(d\) equal
408: to their London-tail coefficients.
409: Figures~\ref{ar237}--\ref{kr2} 
410: show that for \(2 \le r \le 7\)\AA,
411: the hybrid form
412: (\ref{phenpot}) (solid, red) 
413: fits the Aziz potentials for both Ar\(_2\) and
414: for Kr\(_2\) (pluses, blue)\@.
415: The Lennard-Jones \(V_{LJ}\) curves
416: (\ref{LJ}) (dashes, green) 
417: matched at the minima
418: are too deep for \( r > 4.5 \)~\AA\ 
419: (Figs.~\ref{ar237} \& \ref{kr237})
420: and too hard for \( r < 3 \)~\AA\ 
421: (Figs.~\ref{ar2} \& \ref{kr2})\@.
422: %Being accurate at all relevant distances,
423: The potential \(V(r)\) of Eq.(\ref{phenpot}) 
424: represents weak noncovalent bonds
425: better than \(V_{LJ}\)
426: (and \(V_H\))\@.
427: \par
428: Does it matter that \(V_{LJ}(r)\)
429: fails to fit 
430: the Ar--Ar and Kr--Kr interactions?
431: To find out, we used \(V\) and  \(V_{LJ}\)
432: to compute the dimensionless second virial coefficient \(B_2/r_0^3\)
433: of Ar and Kr at room temperature (\(kT = 0.025 \) eV)\@.
434: Here \(r_0\) is the minimum of the potential, and
435: the second virial coefficient \(B_2\) is the integral
436: over all space
437: \beq
438: B_2(T) = - {\thalf} \, \int \! d^3r \, 
439: \left( e^{-\beta V(r)} -1 \right) 
440: \label{B_2}
441: \eeq
442: in which \(\beta = 1/(kT)\)\@.
443: The hybrid potential \(V\) 
444: fitted to the curves of Figs.~\ref{ar237} -- \ref{kr2}
445: gives \(B_2/r_0^3 = -0.499\) for Ar and 
446: \(-1.35\) for Kr,
447: which respectively differ from 
448: the experimental~\cite{Lide1994}
449: values of \(-0.552\) and \(-1.41\) by 9.6\% and 4.1\%\@.
450: The Lennard-Jones potential \(V_{LJ}\)
451: fitted to \(r_0\) and \(V(r_0)\)
452: gives \(B_2/r_0^3 = -0.899\) for Ar and \(-1.92\) for Kr
453: (errors of 63\% and 37\%);
454: it also wags a long-range tail with  
455: London coefficients \(2 \, |V(r_0)| \, r_0^6 \)
456: that are too large by 83\% for Ar and by 84\% for Kr\@.
457: If the parameters \(r_0\) and \(V(r_0)\)
458: in Eq.(\ref{LJ}) for \(V_{LJ}\)
459: are chosen to give the correct 
460: second virial coefficient \(B_2/r_0^3\)
461: for a range of temperatures~\cite{Hill1986},
462: then \(V(r_0)\) 
463: is too shallow
464: by 16 \% for Ar and 15\% for Kr,
465: and the London coefficients of the 
466: long-range tail are too large by 70\% for Ar and by 64\% for Kr\@.
467: With only two parameters,
468: Lennard-Jones fits are procrustean.
469: \par
470: What about additivity (\cite{Axilrod1943,Muto1943,Cole1988})?
471: When three or more atoms interact,
472: their potential energy is not
473: the sum of the three (or more) pair potentials.
474: Is the accuracy of the hybrid form important
475: in the liquid phase
476: where additivity is only approximate?
477: \par
478: To test whether the lack of complete additivity
479: in the liquid phase obscures the 
480: advantages of the hybrid form \(V\) over the
481: Lennard-Jones potential \(V_{LJ}\),
482: we used both to compute  
483: the heats of vaporization \(\Delta_{\mathrm{vap}}H\)
484: of Ar and Kr at their
485: boiling points at atmospheric pressure.
486: In our Monte Carlo simulations,
487: we imposed periodic boundary conditions 
488: to reduce finite-size effects.
489: Our Monte Carlo code is 
490: available at 
491: \href{http://bio.phys.unm.edu/latentHeat/}
492: {bio.phys.unm.edu/latentHeat}.
493: We used it
494: to compute the potential energy \(U\) per
495: atom in the liquid and gas phases.
496: The latent heat of vaporization 
497: \( \Delta_{\mathrm{vap}}H \)
498: is the difference between the potential
499: energies \(U_{\textrm{gas}}\) and \(U_{\textrm{liquid}}\)
500: plus the work done in expanding by \(\Delta V\)
501: against the pressure \(p\)
502: of the atmosphere,
503: \(\Delta_{\mathrm{vap}}H = U_{\textrm{gas}} \,
504: - \, U_{\textrm{liquid}} \, + \, p \, \Delta V \)\@.
505: \par
506: The hybrid form \(V\) 
507: fitted to the curves 
508: of Figs.~\ref{ar237} -- \ref{kr2} gave
509: \(\Delta_{\mathrm{vap}}H = 0.0694\) eV (per atom) for Ar
510: and 0.0982 eV for Kr,
511: which differ from 
512: the experimental values~\cite{Lide1994} of 0.0666
513: and 0.0941 eV by 4.2\% and 4.4\%\@.  % CRC
514: %and 0.0932 eV by 5.0\% and 6.0\%\@. % NIST
515: In equivalent Monte Carlo simulations,
516: the Lennard-Jones potential \(V_{LJ}\) fitted
517: to \(r_0\) and \(V(r_0)\) gave
518: and \(\Delta_{\mathrm{vap}}H = 0.0787\) eV for Ar 
519: and 0.111 eV for Kr
520: (errors of 18\% and 18\%)\@.
521: So the errors due to a lack of additivity
522: are of the order of 4\%,
523: while those due to the Lennard-Jones potential
524: are about 18\%\@.
525: Even in the liquid phase,
526: limited additivity is less of a problem
527: than the defects of the Lennard-Jones potential.
528: \par
529: For a wide class of atom pairs,
530: the hybrid form 
531: (\ref{phenpot})
532: can reliably represent the best
533: spectroscopically determined
534: potentials over all relevant distance scales. 
535: It also yields accurate second virial coefficients
536: and heats of vaporization.
537: Its simplicity recommends it as a teaching tool
538: and as a practical form for computation.
539: Given the differences between it
540: and the 
541: Lennard-Jones, harmonic, Morse,
542: Varnshi, and Hulburt-Hirschfelder potentials, 
543: it would be worthwhile
544: to examine the consequences of
545: these differences in Monte Carlo searches
546: for low-energy states of biomolecules
547: and in numerical simulations
548: of phase transitions and reactions
549: far from equilibrium.
550: \begin{acknowledgments}
551: Thanks to S.\ Valone for conversations 
552: and for RKR data;
553: to S.\ Atlas, C.\ Beckel, 
554: B.\ Brooks, J.\ Cohen, K.\ Dill, D.\ Harries, G.\ Herling,
555: M.\ Hodoscek, R.\ Pastor, 
556: R.\ Podgornik, W.\ Saslow, and C.\ Schwieters for advice.
557: P.~J.\ Steinbach kindly hosted KC\ at NIH,
558: where we used the Biowulf computers.
559: \end{acknowledgments}
560: 
561: \bibliography{chem,cs,physics,vdw}
562: 
563: \end{document}
564: 
565: