1: % Template article for preprint document class `elsart'
2: % SP 2006/04/26
3:
4: \documentclass{elsart}
5:
6: % Use the option doublespacing or reviewcopy to obtain double line spacing
7: % \documentclass[doublespacing]{elsart}
8:
9: % if you use PostScript figures in your article
10: % use the graphics package for simple commands
11: % \usepackage{graphics}
12: % or use the graphicx package for more complicated commands
13: % \usepackage{graphicx}
14: % or use the epsfig package if you prefer to use the old commands
15: % \usepackage{epsfig}
16:
17: % The amssymb package provides various useful mathematical symbols
18: % \usepackage{amssymb}
19: \usepackage{graphicx, subfigure}
20: \usepackage{psfrag}
21:
22: % The lineno packages adds line numbers. Start line numbering with
23: % \begin{linenumbers}, end it with \end{linenumbers}. Or switch it on
24: % for the whole article with \linenumbers.
25: % \usepackage{lineno}
26:
27: % \linenumbers
28: \begin{document}
29:
30: \begin{frontmatter}
31: % Title, authors and addresses
32:
33: % use the thanksref command within \title, \author or \address for footnotes;
34: % use the corauthref command within \author for corresponding author footnotes;
35: % use the ead command for the email address,
36: % and the form \ead[url] for the home page:
37: % \title{Title\thanksref{label1}}
38: % \thanks[label1]{}
39: % \author{Name\corauthref{cor1}\thanksref{label2}}
40: % \ead{email address}
41: % \ead[url]{home page}
42: % \thanks[label2]{}
43: % \corauth[cor1]{}
44: % \address{Address\thanksref{label3}}
45: % \thanks[label3]{}
46:
47: \title{NMscatt: a program for calculating inelastic scattering
48: from large biomolecular systems
49: using classical force-field simulations}
50:
51: % use optional labels to link authors explicitly to addresses:
52: % \author[label1,label2]{}
53: % \address[label1]{}
54: % \address[label2]{}
55:
56: \author[nic]{Franci Merzel\corauthref{cor}},
57: \ead{franc@cmm.ki.si}
58: \author[ill,delft]{Fabien Fontaine-Vive} and
59: \author[ill]{Mark R. Johnson\corauthref{cor}}
60: \corauth[cor]{Corresponding authors.}
61: \ead{johnson@ill.fr}
62:
63: \address[nic]{National Institute of Chemistry, Hajdrihova 19, 1000
64: Ljubljana, Slovenia}
65: \address[ill]{Institute Laue Langevin, BP156, 38042 Grenoble cedex
66: 9, France}
67: \address[delft]{Radiation, Reactors and Radionuclides Department,
68: Faculty of Applied Sciences, Delft University of Technology,
69: Mekelweg 15, 2629 JB Delft, The Netherlands}
70:
71: \begin{abstract}
72: Computational tools for normal mode analysis, which are widely used in
73: physics and materials science problems, are designed here in
74: a single package called \texttt{NMscatt} (Normal Modes \& scattering)
75: that allows arbitrarily large systems to be
76: handled. The package allows inelastic neutron and X-ray scattering
77: observables to be calculated, allowing comparison with experimental
78: data produced at large scale facilities. Various simplification
79: schemes are presented for analysing displacement vectors, which are
80: otherwise too complicated to understand in very large systems.
81: \end{abstract}
82:
83: % keywords here, in the form: keyword \sep keyword
84: \begin{keyword}
85: vibrational analysis\sep phonons\sep
86: atomic force-field simulations\sep inelastic neutron/X-ray
87: scattering\sep dynamical structure factor
88:
89: % PACS codes here, in the form: \PACS code \sep code
90: \PACS 87.15.-v\sep 87.15.Aa\sep 63.20.Dj\sep 61.10.Dp
91: \end{keyword}
92: \end{frontmatter}
93:
94: %{\bf Program summary}\\
95: %{\it Title of the progam:}\texttt{NMscatt}\\
96: %{\it Catalog identifier:}\\
97: %{\it Program summary URL:}\\
98: %{\it Program obtainable from:} CPC Program Library, Queen's University
99: % of Belfast, N. Ireland\\
100: %{\it Computer:} x86 PC\\
101: %{\it Operating system:} GNU/Linux, UNIX\\
102: %{\it Programming language used:} FORTRAN77\\
103: %{\it Memory required:} Depends on the system size to be simulated.\\
104: %{\it No. of bits in a word:} 32 or 64\\
105: %{\it No. of processors used:} 1\\
106: %{\it Parallelized?:} No\\
107: %{\it No. of lines in distributed program, including test data, etc.:}
108: %2 500\\
109: %{\it No. of bytes in distributed program, including test data, etc.:}
110: %74 333\\
111: %{\it Distribution format:} tar.gz\\
112: %{\it Typical running time:} About 7 hours per one k-point evaluation
113: %in sampling all modes dispersion curves for a system containing 3550
114: %atoms in the unit cell on AMD Athlon 64 X2 Dual Core Processor 4200+.\\
115: %{\it Nature of the physical problem:} Normal mode analysis, phonons
116: %calculation, derivation of incoherent and coherent inelastic
117: %scattering spectra.\\
118: %{\it Method of solution:} Full diagonalization (producing
119: %eigen-vectors and eigen-values) of dynamical matrix which is
120: %obtained from potential energy function derivation using finite
121: %difference method.\\
122:
123:
124: % main text
125: \section{Introduction}
126: At large scale facilities for neutron and X-ray scattering, large
127: quantities of experimental data are produced. For complex, nanoscale
128: systems, understanding this data requires computer models. In the case
129: of inelastic scattering, molecular dynamics (MD) simulations
130: \cite{leach} are
131: widely used to equilibrate structures and explore dynamics as a
132: function of temperature and other experimental parameters. However MD
133: only gives a partial description of vibrational modes through the
134: partial density of states and when knowledge about specific
135: vibrational modes is required, normal mode analysis (NMA) \cite{dusa}
136: has to be
137: performed. For physics and materials science problems, NMA gives a
138: description of the lattice dynamics via the dispersion (k-vector
139: dependence) of the mode frequencies \cite{frank}. For small systems
140: ($<$ 200 atoms in
141: the simulation box) very accurate results can be obtained using
142: density functional theory (DFT) methods to determine interatomic force
143: constants \cite{parlinski1, mark1}. By combining DFT and software that
144: constructs and
145: diagonalises the dynamical matrix and calculates the experimental
146: observables, experimentalists now have sophisticated tools to analyse
147: their data. The PHONON \cite{phonon} package is one of the best examples.
148:
149: Phonon codes are traditionally limited to small systems for a number
150: of reasons. For example, for small unit cells, the reciprocal lattice
151: is big and stronger effects of dispersion are expected. If DFT methods
152: are used to determine force constants then these methods are
153: themselves restricted to a few hundred atoms. However the development
154: of nanoscale structures of (partially) crystalline materials
155: stimulates a need for phonon codes to be extended to much larger
156: systems. Parameterised force fields \cite{ff} can be used to determine the
157: inter-atomic force constants and, as will be seen in the example
158: presented here, strong dispersion effects are observed inspite of the
159: small reciprocal lattice.
160:
161: In biomolecular systems, the need for NMA has long been recognised and
162: codes like CHARMM \cite{charmm} allow the gamma point normal modes to
163: be calculated for moderately big systems. In addition, the neutron
164: scattering quantities can be directly calculated from the
165: simulations using the time-correlation function formalism
166: \cite{vhove}, as implemented in the nMOLDYN program \cite{nmoldyn}.
167: A combination of neuton scattering experiments and atomic detail computer
168: simulations has proven to be a powerful technique for studying
169: internal molecular vibrations \cite{jeremy1,jeremy2,jeremy3}. In this approach
170: one can validate the applied numerical models, i.e. force field
171: parametrizations, depending on the agreement between the experimental
172: and calculated spectra.
173:
174: In this paper we present a software package that extends the
175: functionality of codes like PHONON \cite{phonon} and Climax
176: \cite{climax} to arbitrarily large systems and
177: extends the gamma point only analysis already available for larger
178: systems to include k-vector dependence. The software reads a Hessian
179: matrix of force constants, constructs and diagonalises the dynamical
180: matrix for any k-vector and calculates neutron and X-ray scattering
181: observables. The computational bottleneck remains the diagonalisation
182: of correspondingly large dynamical matrices and we comment on
183: approximations that have to be used when the Brillouin zone cannot be
184: sampled at a large number of points. In large systems, atomistic
185: detail in the displacement vectors can be difficult to interpret due
186: to the large number of degrees of freedom and we present two methods
187: for simplifying this information. The first entails summing
188: displacement vectors over atoms in user-defined beads, while the
189: second involves a reduction of the degrees of freedom in the dynamical
190: matrix by summing over force constants, which has the advantage of
191: reducing the number of modes to be examined.
192:
193: \section{Theoretical background}
194:
195: The standard approach, also called a direct method \cite{parlinsk2},
196: to the lattice vibration problem of
197: crystals is based on the explicit knowledge of the interaction
198: between all atom-pairs in the system. Subsequently, one deduces
199: the corresponding force constants, and constructs and diagonalizes the
200: dynamical matrix for any k-vector in order to obtain the frequencies
201: of the normal modes. A reasonable atomic detail description of interactions
202: within large biomolecular systems are provided using empirical
203: force fields.
204:
205: In the following we will briefly summarize the aspects
206: of the classical theory of lattice vibrations \cite{latt} and proceed to the
207: description of the explicit phonon calculations.
208:
209: The individual atomic positions in the crystal can be assigned as
210: \begin{equation}
211: \vec R_{n\mu}(t) = \vec R_n+\vec r_\mu+\vec u_{n\mu}(t),
212: \end{equation}
213: where $\vec R_n$ is unit cell lattice vector and $\vec u_{n\mu}(t)$ is
214: displacement of atom $\mu$ from its equilibrium positions $\vec
215: r_\mu$. Within the harmonic approximation we concentrate on expansion
216: of the small differences of potential energy $V$ due to the
217: small changes in atom positions:
218: \begin{equation}
219: V({\bf u})\approx V_0+\sum_{n\mu\alpha}\frac{\partial V}{\partial
220: u_{n\mu\alpha}}u_{n\mu\alpha}+\frac{1}{2}\sum_{n\mu\alpha,m\nu\beta}
221: u_{n\mu\alpha}{\bf D}_{n\mu\alpha,m\nu\beta}u_{m\nu\beta}+\dots,
222: \end{equation}
223: where the second derivative defines the force constant between
224: the atoms $\mu$ and $\nu$:
225: \begin{equation}
226: {\bf D}_{n\mu\alpha,m\nu\beta}=\frac{\partial^2 V}{\partial
227: u_{n\mu\alpha}\partial u_{m\nu\beta}},\qquad(\alpha,\beta=x,y,z)
228: \label{le1}
229: \end{equation}
230: As each unit cell is identical to every other unit cell in the
231: crystal, the displacement pattern of a normal mode has to be identical
232: to that in any other cell to within a phase difference $\vec
233: k(\vec R_n-\vec R_m)$. The representation of the atom displacement
234: is chosen to be a plain wave ansatz of the form:
235: \begin{equation}
236: \vec u_{n\mu}(\vec k,t) = \frac{u_0}{\sqrt{M_\mu}}\vec e_{\mu\vec
237: k}\exp(i[\vec k\vec R_{n}-\omega_{\vec k}t]),
238: \label{le2}
239: \end{equation}
240: where $\vec e_{\mu\vec k}$ is the polarization vector and $M_\mu$ is
241: the mass of the atom $\mu$. We omit writing
242: Cartesian component subscripts. Solving the equation of motion with ansatz
243: (\ref{le2}) is equivalent to the eigen-value problem
244: \begin{equation}
245: \omega_{\vec kj}^2\vec e_{\mu\vec kj}=\sum_\nu{\cal D}_{\mu \nu}(\vec k)\vec
246: e_{\nu\vec kj},
247: \label{le3}
248: \end{equation}
249: where
250: \begin{equation}
251: {\cal D}_{\mu \nu}(\vec k)=\sum_m\frac{1}{\sqrt{M_\mu M_\nu}}{\bf
252: D}_{n\mu,m\nu}\exp[i\vec k(\vec R_m-\vec R_n)]
253: \label{le4}
254: \end{equation}
255: is the so called dynamical matrix.
256:
257: The form of the dynamical matrix (\ref{le4}) requires
258: the atom pairs for which one atom belongs to a different unit cell,
259: {\it i.e.} $(m\ne n)$, to be identified. These terms contribute to the so called
260: Bloch-factor $\exp[i\vec k(\vec R_m-\vec R_n)]$ and make the dynamical
261: matrix complex. But in case of applying periodic boundary conditions
262: (PBC) as implemented in computer simulation programs, the potential
263: energy of a crystal is given as an explicit function of only the atom
264: positions in the primary unit cell. As a consequence, we obtain the
265: second derivative matrix $\bf H$ in which the contributions from the
266: inter-cell atomic pairs are mapped and added to the corresponding image atom
267: pairs in the primary unit cell:
268: \begin{equation}
269: \left[\sum_m{\bf D}_{n\mu,m\nu}\right]\to{\bf H}_{\mu,\nu}=
270: \frac{\partial^2 V^{PBC}}{\partial
271: u_{\mu}\partial u_{\nu}}
272: \label{le5}
273: \end{equation}
274: One can directly obtain ${\bf D}_{n\mu\alpha,m\nu\beta}$ by
275: increasing the size of the unit cell by one or more layers of
276: periodically arranged image cells and calculate force constants
277: in the extended supercell. However, this approach is unfavourable when
278: dealing with very large systems.
279:
280: A similar approach is to decompose the potential energy $V^{PBC}$ in
281: equation (\ref{le5}) into individual contributions from the image
282: cells, $V^{PBC}=V_0+\sum_mV_m$ and evaluate the second derivative
283: matrix for each term upon the same minimized structure.
284:
285: The situation is simpler if the interaction is truncated at some
286: cutoff distance $r_{cut}$ so that the ``minimum image
287: convention'' (MIC) is obeyed. The MIC states that each atom interacts
288: at most with one image of every other atom in the system (which is
289: repeated to fully enclose the primary unit cell with the periodic
290: boundary conditions). This has the effect of limiting the interaction
291: cutoff, for example, to no more than half the length of the minimum side when
292: simulating the orthorhombic cell, $r_{cut} < min\{a/2,b/2,c/2\}$. It
293: should be noted
294: that the size of nanoscale crystals usually far exceeds the
295: spatial range of forces between atoms ($\sim$ $<$12 \AA) allowing physically
296: reasonable cutoff radii to be introduced.
297:
298: \begin{figure}[ht]
299: \begin{center}
300: \includegraphics[width=0.5\textwidth]{pbc_e.eps}
301: \caption{A 2D periodic system representing the effect of the minimum
302: image convention while constructing dynamical matrix.
303: If a distance between atom $\nu$ (hollow blue circle) and
304: $\mu$ (full black circle) is larger than the cut-off distance (red
305: circle around the atom $\mu$), and if the force constant
306: ${\bf H}_{\mu\nu}$ turns out to be nonzero, it means that the
307: value of ${\bf H}_{\mu\nu}$ has to be related to the atom $\nu$ in the
308: image cell $m$ (full blue circle), with the position vector
309: $\vec r_\nu+\vec l_x$ ($\vec l_x=\vec R_m-\vec R_n$), instead of to
310: the atom $\vec r_\nu$ in the primary cell $n$. The corresponding lattice
311: translation vector $\vec l_x$ is thus determined for the element ${\bf
312: H}_{\mu\nu}$.}
313: \label{fg1}
314: \end{center}
315: \end{figure}
316:
317: According to the MIC we can conclude that there is always only one
318: translation $\vec t$ per atom pair $(\nu,\mu)$ giving rise to the
319: minimum distance $|(\vec r_\nu+\vec t)- \vec r_\mu|$:
320: \begin{eqnarray}
321: \exists s_\alpha &=& \{-1,0,1\}_{\alpha=1,2,3}:
322: \vec t=s_1\vec l_1 + s_2\vec l_2 + s_3\vec l_3;\\
323: && {\textrm{so
324: that:}\quad |(\vec r_\nu+\vec t)- \vec r_\mu|=min,}\nonumber
325: \end{eqnarray}
326: where $\vec l_\alpha$ are lattice translation vectors, c.f. Fig. \ref{fg1}.
327: In the case of a nonvanishing ${\bf H}_{\mu,\nu}$ we get the
328: following expression for a dynamical matrix element
329:
330: \begin{equation}
331: {\cal D}_{\mu \nu}(\vec k)=\frac{1}{\sqrt{M_\mu M_\nu}}\left\{
332: \begin{array}{ll}
333: {\bf H}_{\mu,\nu}, & {\textrm{if : }} |\vec r_\mu-\vec
334: r_\nu|<r_{cut}\\
335: {\bf H}_{\mu,\nu}\exp(i\vec k\vec t), & {\textrm{if : }} |\vec r_\mu-\vec
336: r_\nu|>r_{cut}\\
337: \end{array}\right.
338: \end{equation}
339:
340: According to equation (\ref{le3}) the diagonalization of
341: matrix ${\cal D}(\vec k)$ yields the phonon
342: frequencies $\omega_{\vec kj}$ and corresponding polarization
343: vectors $\vec e_{\mu\vec kj}$ for a given phonon wave vector $\vec k$.
344: A complete solution leads to the phonon dispersion relations. The
345: subscript $j$ denotes a branch in the phonon dispersion. In a
346: crystal of $N$ atoms, there are $3N$ branches.
347:
348: \vskip 0.2cm
349: {\bf Inelastic scattering}
350: \vskip 0.2cm
351:
352: The dynamic structure factor $S(\vec q,\omega)$ contains information
353: about the structure and dynamics of the sample. It can be split into
354: a coherent part arising from the cross-correlations of atomic motions
355: and an incoherent part describing self-correlations of single atom
356: motions. According to the standard theory \cite{bee, lovesey},
357: which is based on the harmonic approximation, we obtain the following
358: expressions for the coherent and incoherent dynamical structure factors:
359: \begin{eqnarray}
360: S(\vec q,\omega)_{coh} &=& \sum_{\vec G}\sum_{\vec
361: k,j}\frac{\hbar}{2\omega_{\vec kj}}
362: \left|\sum_\mu\sigma_\mu^{coh}\frac{\vec q \cdot\vec e_{\mu\vec
363: kj}}{\sqrt{M_\mu}}\exp(-W_\mu(\vec q)+i\vec q\vec r_\mu)
364: \right|^2\times\nonumber\\
365: &\times&
366: (n(\omega_{\vec kj})+1)\delta(\omega-\omega_{\vec kj})
367: \delta(\vec q+\vec k-\vec G),
368: \label{eqn_coh}
369: \end{eqnarray}
370: and
371: \begin{eqnarray}
372: S(\vec q,\omega)_{inc} &=& \sum_\mu\sigma_\mu^{inc}\sum_{\vec
373: k,j}\frac{\hbar}{2M_\mu\omega_{\vec kj}}|\vec q \cdot\vec e_{\mu\vec
374: kj}|^2(n(\omega_{\vec kj})+1)\times\nonumber\\
375: &\times& \exp[-2W_\mu(\vec
376: q)]\delta(\omega-\omega_{\vec kj}),
377: \end{eqnarray}
378: where $\vec q$ is the scattering vector, $\sigma_\mu$ is the
379: corresponding atomic scattering
380: length, $(n(\omega)+1)$ refers to the
381: phonon creation process (absorption spectrum) and $n(\omega)$ is the
382: mean number of phonons of frequency $\omega$ at temperature $T$
383: according to the Bose-Einstein statistics
384: \begin{equation}
385: (n(\omega)+1)=\frac{\exp(\hbar\omega/k_BT)}{\exp(\hbar\omega/k_BT)-1}
386: \end{equation}
387: The factor $\exp[-2W_\mu(\vec q)]$ is called the Debye-Waller factor:
388: \begin{equation}
389: W_\mu(\vec q)=\frac{\vec{q}\cdot{\bf B}(\mu)\cdot\vec{q}}{2},\qquad
390: B_{\alpha\beta}(\mu)=\langle u_{\mu\alpha}u_{\mu\beta}\rangle_T,
391: \end{equation}
392: where ${\bf B}(\mu)$ is a $3\times 3$ symmetric tensor representing
393: the thermodynamic mean square displacement of an atom $\mu$, which can
394: be expressed by the partial atomic phonon density of states
395: $g_{\alpha\beta,\mu}(\omega)$:
396: \begin{equation}
397: B_{\alpha\beta}(\mu)=\frac{{3N}\hbar}
398: {2M_\mu}\int_0^{\infty}\frac{d\omega}{\omega}
399: g_{\alpha\beta,\mu}(\omega)
400: \coth\left(\frac{\hbar\omega}{2k_BT}\right).
401: \label{le6}
402: \end{equation}
403: The partial atomic density of states is a weighted distribution of
404: normal modes
405: \begin{equation}
406: g_{\alpha\beta,\mu}(\omega)=\frac{1}{3Nn}\sum_{\vec
407: k,j}^{n3N}e_{\alpha\mu\vec kj}e^*_{\beta\mu\vec kj}
408: \delta(\omega-\omega_{\vec kj}),
409: \end{equation}
410: where $n$ is the number of sampling $\vec k$-points in the first
411: Brillouin zone.
412:
413: The evaluation of the Debye-Waller $B$ factors
414: using equation (\ref{le6}) requires extra attention due to the
415: ``zero-phonon'' term resulting from
416: the singularity caused by phonon (acoustic) branches where
417: $\omega(\vec q,j)=0$. The contributions of the 3 acoustic
418: modes is treated separately by using the Debye approximation for the
419: density of states, {\it i.e.} $g_{ac}(\omega)\propto\omega^2$, and
420: normalization $\int_0^{\omega_{max}}g_{ac}(\omega)d\omega=3$,
421: where $\omega_{max}$ is the maximum frequency up to which the
422: acoustic dispersion curve is linear
423: \begin{eqnarray}
424: B_{\alpha\beta}(\mu) &=& \sum_{j=1}^3\frac{3\hbar}
425: {M_\mu\omega_{j\,max}^3}\int_0^{\omega_{j\,max}}
426: \omega e_{\alpha\mu\vec kj}e^*_{\beta\mu\vec kj}
427: \left(\frac{1}{\exp(\hbar\omega/k_BT)-1}+\frac{1}{2}\right)d\omega
428: \nonumber\\
429: &=& \sum_{j=1}^3\frac{3\hbar e_{\alpha\mu\vec kj}e^*_{\beta\mu\vec kj}}
430: {M_\mu\omega_{j\,max}}\left[\left(\frac{k_BT}
431: {\hbar\omega_{j\,max}}\right)^2\int_0^{x_{j\,max}}
432: \frac{x}{\exp(x)-1}dx+\frac{1}{4}\right].
433: \end{eqnarray}
434: A new variable $x_{j\,max}={\hbar\omega_{j\,max}}/{k_BT}$ was
435: introduced in the last equation.
436:
437:
438: \section{Analysing the displacement vectors}
439:
440: For systems containing thousands of atoms (N) the displacement vectors
441: obtained by diagonalising the dynamical matrix can be difficult to
442: understand, especially for low frequency modes which involve the
443: displacement of many (or all) atoms. One simple solution to this
444: problem is to sum over the displacements of atoms within beads, which
445: represent logical coarse grains of the system, for example base
446: molecules in the DNA example below. This treatment allows different
447: bead definitions to be applied to the calculated displacement
448: vectors but has the disadvantage of not reducing the number of
449: displacement vectors from 3N.
450:
451: A related approach is to reduce the atomic level Hessian matrix to
452: lower dimension by mapping the inter-atomic force constants on to
453: inter-bead force constants \cite{venkat}. For N' beads, the resulting dynamical
454: matrices have dimension 3N' and therefore result in 3N' displacement
455: vectors for any k-vector. We note that any reduction in the
456: dimensionality of the system causes a loss in information, which is
457: the rotational degrees of freedom of the beads (rigid bodies).
458:
459:
460: \section{Example}
461:
462: To verify the implemented formalism we have simulated a
463: B-form DNA molecule (right-handed, 10 base-pairs per turn, pitch
464: 33.6\AA) using CHARMM \cite{charmm}. The full crystal
465: environment was generated using periodic boundary conditions for
466: an orthorhombic unit cell containing one helix of DNA.
467: The dimension of the unit cell is $a=32.2$\AA, $b=31.8$\AA\ and
468: $c=33.5$\AA, with $c$ parallel to the helical axis.
469: The starting configuration was obtained by minimization of the
470: potential energy of the crystal structure obtained as the
471: time average over a 1ns MD simulation at 100K. The Hessian matrix of
472: force constants was generated by displacing each atom in turn from
473: equilibrium and calculating the forces induced on all other
474: atoms. Diagonalisation of the resulting dynamical matrices was
475: performed using the the routine {\texttt{zcheev}} from the LAPACK
476: library \cite{LAPACK}.
477:
478: \begin{figure}
479: \psfrag{S(q,w) [arb. units]}{\small $S(\vec q,\omega)$ [arb. units]}
480: \psfrag{energy transfer [cm-1]}{\small energy transfer $\omega [cm^{-1}]$}
481: \begin{center}
482: \includegraphics[width=0.8\textwidth]{coh.eps}
483: \caption{Coherent dynamical structure factor of B-DNA.
484: Scattering vector $\vec q$ is varied along the helical axis (0,0,1).
485: Shown are $S(\vec q,\omega),\ |\vec q|=0.019\dots0.581$\AA$^{-1}$.}
486: \label{fg2}
487: \end{center}
488: \end{figure}
489:
490: The typical coherent spectrum of DNA, Figure \ref{fg2}, obtained from
491: equation \ref{eqn_coh}, shows a well-defined
492: Brillouin peak at small $\omega$ which moves along the frequency
493: dimension upon varying momentum transfer $\vec q$. Fitting the
494: spectral profile with a Gaussian as a function of wave-vector gives
495: the dispersion curve shown in Figure
496: \ref{fg3}, which compares well with the recent experimental
497: results \cite{dna_jcp}.
498: \begin{figure}
499: \psfrag{w [cm-1]}{\small $\omega$ [cm$^{-1}$]}
500: \psfrag{q [A-1]}{\small $q$[\AA$^{-1}$]}
501: \begin{center}
502: \includegraphics[width=0.8\textwidth]{disp.eps}
503: \caption{Dispersion curves obtained by Gaussian fitting of the
504: calculated Brillouin peak for inelastic X-ray scattering using
505: \texttt{NMscatt} (solid curve + circles), and using nMoldyn (dashed
506: curve + squares).}
507: \label{fg3}
508: \end{center}
509: \end{figure}
510: Figure \ref{fg3} also shows the result of the equivalent analysis of a
511: 300K MD simulation on the same model of B-DNA using nMoldyn.
512:
513: In order to gain insight into the nature of low frequency dynamics we
514: can analyse the displacement vectors at an atomic level (see Figure
515: \ref{fg5}(a) for an acoustic mode).
516: By summing over the displacement
517: vectors in terms of beads, where base
518: molecules, sugar molecules and phosphate groups are treated as single
519: units, a simplified picture of the normal modes is obtained. Figure
520: \ref{fg5}(b) shows a high frequency mode which has a pronounced
521: contribution from the phosphate groups.
522: \begin{figure}
523: \begin{center}
524: \subfigure[B-DNA: acoustic mode]{\includegraphics
525: [width=0.40\textwidth]{vmode0.ps}}
526: \qquad
527: \subfigure[B-DNA: beads]{\includegraphics
528: [width=0.454\textwidth]{vmode1.ps}}
529: \caption{Atomic {\it a)} and {\it b)} beads
530: (base molecules, sugar molecules and
531: phosphate groups) displacements of the selected mode.}
532: \label{fg5}
533: \end{center}
534: \end{figure}
535:
536: \section{Conclusion}
537: The new, user-friendly computational package
538: \texttt{NMscatt} presented here enable an efficient
539: atomic detail analysis of different types of inelastic scattering
540: applied to arbitrarily large nanoscale systems. The ability to
541: perform molecular dynamics and phonon calculations on large
542: nano and bio-molecular materials means that one can efficiently pursue
543: the investigation of some poorly understood structural and dynamical
544: features of these systems.
545:
546: \section{Acknowledgements.}
547: The authors are grateful to
548: Dr. Stephane Rols for helpful discussions at the start of this project.
549: MJ acknowledges a long-standing collaboration with Prof. Krzysztof
550: Parlinski. FM acknowledges a financial support from ILL during his
551: stay in Grenoble and support from the Ministry of Higher Education,
552: Science and Technology of Republic of Slovenia under Grant nos.
553: P1-0002, J1-6331 and J1-5115.
554:
555: \vskip 1cm
556: \appendix
557: \section{Program package and data structure}
558:
559: There are four main modules in the \texttt{NMscatt} program package
560: {\it phonon, coh, incoh} and {\it bead}, and the overall
561: \texttt{NMscatt} structure
562: is given in Figure \ref{fg6}. Below are described the corresponding
563: modules.
564: \begin{itemize}
565: \item{\it phonon}: Providing the full Hessian matrix for a given
566: energy-minimized atomic structure within the specified
567: crystallographic unit cell this module constructs dynamical matrix and
568: calculates its eigenvalues and eigenvectors at given wave vector
569: $\vec k$. At input this module requires to specify the Bravais lattice
570: vectors that were previously used in the molecular mechanics/dynamics
571: simulation
572: package to satisfy the periodic boundary conditions while generating
573: the minimized structure. The cut-off radius must be given at which
574: the long range interactions are truncated while calculating the Hessian
575: matrix in the simulation. At output two separate files
576: \texttt{eig\_val\_}$n$ and
577: \texttt{eig\_vec\_}$n$ are written containing
578: eigenvalues and eigenvectors, respectively, where $n$ assigns
579: a consecutive number of the sampling $k$-point in the first Brillouin
580: zone. These files serve as an imput for other modules of
581: \texttt{NMscatt}. The choice $n=0$ is assumed to be reserved for the
582: $\Gamma$-point, $\vec k=(0,0,0)$. The lowest $n$ should sample
583: the vicinity of the $\Gamma$-point. The $k$-points are to be specified
584: in the fractional coordinates with respect to the reciprocal lattice
585: vectors. The elements of the upper triangle of the hessian matrix should be
586: provided in the binary (default) \texttt{hessian\_uf} or alternatively in the
587: ASCII file \texttt{hessian\_f} and the atomic coordinates in the file
588: \texttt{coord} writen in the CHARMM coordinate format.
589:
590: \item{\it incoh}: This module allows to calculate atomic
591: Debye-Waller factors representing the mean square displacements
592: and dynamical structure factor of incoherent one-phonon
593: neutron scattering on monocrystals and from orientationally
594: averaged powder. In the latter case we need to specify an
595: absolute value of momentum transfer $q$, range of the $k$-point
596: index $n$: $0-n_{max}$ for picking up the corresponding $k$-point
597: eigenvalues- and eigenvectors-files generated by {\it phonon},
598: number of random orientations
599: of $q$ vector to provide spherical averaging and absolute temperature.
600: Also, one has to define the $k$-point index $n_a$, for which the
601: $k$-point range $0-n_a$ corresponds to the linear regime of the
602: acoustic-mode-dispersion curves. This is needed for the proper
603: derivation of Debye-Waller factors using Debye approximation.
604: As a result, the function $S(q,\omega)$ is given in the file
605: \texttt{s\_q\_w}. Optionally, one can also obtain density of states
606: (DOS) in this module.
607:
608: \item{\it coh}: This module calculates dynamical structure factor of
609: coherent one-phonon neutron or X-ray scattering on monocrystals.
610: In addition to the input data required by {\it incoh} (except for
611: the spherical averaging), we need to define for module {\it coh} also
612: the type of scattering neutron/X-ray, the number of higher Brillouin
613: zones included for sampling momentum transfer vector, and the range
614: of the $k$-point index $n_1-n_2$ evaluated by module {\it phonon}
615: and assigning the $k$-points in the first Brillouin zone, which lie
616: along the selected direction of the momentum transfer vector $\vec q$.
617: As a result, the function $S(q,\omega)$ is given in the file
618: \texttt{s\_qw\_coh}.
619:
620: \item{\it bead}: This module is used to enable visualization of selected
621: vibrational modes obtained by running {\it phonon} for the $\Gamma$-point,
622: such that atomic displacement vectors are projected on to the beads,
623: which are defined as the center of mass of larger atomic groups
624: of the system, for example residues. The output files are readable
625: by the program \texttt{xmakemol}\cite{xmkml} which enables direct
626: visualisation of the mode.
627:
628: \end{itemize}
629:
630: \begin{figure}[ht]
631: \begin{center}
632: \includegraphics[width=0.7\textwidth]{flowch.eps}
633: \caption{Flow-chart of the \texttt{NMscatt} program package. }
634: \label{fg6}
635: \end{center}
636: \end{figure}
637:
638: \pagebreak
639: \noindent
640: {\bf Compiling}
641:
642: To compile the program package Makefile is provided. It is important
643: to note that 64-bits processors are prerequisite for applying the
644: \texttt{NMscatt} to analysis of the large systems (containing more
645: than 2000 atoms). One needs to install the LAPACK library on the
646: computer beforehand, prior to \texttt{NMscatt}. The fortran compiling
647: switches \texttt{g77 -mcmodel=medium -funroll-all-loops -fno-f2c -O3}
648: are recomended when installed on 64-bits processors running Linux.
649:
650: \noindent
651: {\bf Benchmark results}
652:
653: Bencmark results were obtained on AMD Athlon 64 X2 Dual Core Processor
654: 2.2GHz running Linux for B-form DNA simulated with the CHARMM program.
655:
656: %\noindent
657: %{\bf Sample input}
658:
659: \newpage
660: \clearpage
661:
662: \begin{thebibliography}{00}
663:
664: \bibitem{leach} A.R. Leach, {\it Molecular Modelling: Principles and
665: applications} (Pearson Education, Harlow, 2001).
666:
667: \bibitem{dusa} B.R. Brooks, D. Jane\v zi\v c and M. Karplus, {\it
668: J. Comp. Chem.}, {\bf 16} 1522-1542 (1995).
669:
670: \bibitem{frank} W. Frank, C. Els\" asser and M. F\" ahnle, {\it
671: Phys. Rev. Lett.}, {\bf 74} 1791-1794 (1995).
672:
673: \bibitem{parlinski1} K. Parlinski and G. Chapuis, {\it
674: J. Chem. Phys.}, {\bf 110} 6406-6411 (1999).
675:
676: \bibitem{mark1} M.R. Johnson, K. Parlinski, I. Natkaniec and
677: B.S. Hudson {\it Chem. Phys.}, {\bf 291} 53-60 (2003).
678:
679: \bibitem{phonon} K. Parlinski, PHONON Manual, Version 4.22,
680: {\texttt http://wolf.ifj.edu.pl/phonon/}.
681:
682: \bibitem{ff} A.D.J. MacKerrel, D. Bashford, M. Bellott, et al.,
683: {\it J. Phys. Chem. B.}, {\bf 102}, 3586 (1998).
684:
685: \bibitem{charmm} B. R. Brooks, R. E. Bruccoleri, B. D. Olafson,
686: D. J. States, S. Swaminathan, M. Karplus, J. Comput. Chem. {\bf 4},
687: 187 (1983).
688:
689: \bibitem{vhove} L. van Hove, {\it Phys. Rev.}, {\bf 95}, 249 (1954).
690:
691: \bibitem{nmoldyn} G.R. Kneller, V. Keiner, M. Kneller and M. Schiller,
692: {\it Comput. Phys. Comm.}, {\bf 91} 191-214 (1995).
693:
694: \bibitem{jeremy1}J.C. Smith, {\it Q. Rev. Biophys}, {\bf 24} 227-291 (1991).
695:
696: \bibitem{jeremy2}G.R. Kneller, W. Doster, M. Settles, S. Cusack and
697: J.C. Smith, {\it J. Chem. Phys.}, {\bf 97} 8864-8879 (1992).
698:
699: \bibitem{jeremy3} A.M. Micu, D. Durand, M. Quilichini, M.J. Field and
700: J.C. Smith {\it
701: J. Phys. Chem.}, {\bf 99} 5645-5657 (1995).
702:
703: \bibitem{climax} G.J. Kearley, {\it Nucl. Inst. Meth. Phys. Res. A},
704: {\bf 354} 53-58 (1995).
705:
706: \bibitem{parlinsk2} K. Parlinski, Z. Q. Li and Y Kawazoe,
707: {\it Phys. Rev. Lett.}, {\bf 78}, 4063 (1997).
708:
709: \bibitem{latt} B. Donovan and J.F. Angress, {\it Lattice vibrations}
710: (Chapman and Hall Ltd., London, 1971).
711:
712: \bibitem{bee} M. Bee, {\it Quasielastic Neutron Scattering: Principles
713: and Applications in Solid State Chemistry, Biology and Materials
714: Science} (Hilger, Bristol, 1988).
715:
716: \bibitem{lovesey} S. Lovesey, {\it Theory of Neutron Scattering from
717: Condensed Matter}, international Series of Monographs on physics 72
718: (Oxford Science, Oxford 1984).
719:
720: \bibitem{venkat} G. Venkataraman and V.C. Sahni, {\it
721: Rev. Mod. Phys.}, {\bf 42}, 409 (1970).
722:
723:
724: \bibitem{LAPACK} {\texttt http://www.netlib.org/lapack/}.
725:
726: \bibitem{dna_jcp} Y. Liu, S.H. Chen, D. Berti, P. Baglioni, A. Alatas,
727: H. Sinn, E. Alp and A. Said, {\it J. Chem. Phys.}, {\bf 123}, 214909 (2005).
728:
729: \bibitem{xmkml} {\texttt http://www.nongnu.org/xmakemol/}.
730:
731: \end{thebibliography}
732:
733: % The Appendices part is started with the command \appendix;
734: % appendix sections are then done as normal sections
735: % \appendix
736:
737: % \section{}
738: % \label{}
739:
740: %\begin{thebibliography}{00}
741:
742: % \bibitem{label}
743: % Text of bibliographic item
744:
745: % notes:
746: % \bibitem{label} \note
747:
748: % subbibitems:
749: % \begin{subbibitems}{label}
750: % \bibitem{label1}
751: % \bibitem{label2}
752: % If there is a note, it should come last:
753: % \bibitem{label3} \note
754: % \end{subbibitems}
755:
756: %\bibitem{}
757:
758: %\end{thebibliography}
759:
760: \end{document}
761:
762: