cond-mat0203587/arb.tex
1: \documentstyle[psfig]{espart}
2: \setcounter{topnumber}{3}
3: \setcounter{topnumber}{5} 
4: \setcounter{bottomnumber}{5}
5: \setcounter{totalnumber}{5} 
6: \renewcommand{\topfraction}{0.99}
7: \renewcommand{\bottomfraction}{0.99}
8: \renewcommand{\textfraction}{0.01} 
9: \newcounter{refcaption}
10: \newcounter{nofigs}
11: \setcounter{refcaption}{1}
12: \setcounter{nofigs}{0}
13: \begin{document}
14: \begin{frontmatter}
15: 
16: \title{{\small J. Phys. I France, Vol. 5, p. 1431-1455
17: (1995)}\\[1.5cm]
18: Molecular dynamics of arbitrarily shaped granular particles}
19: 
20: \author{Thorsten P\"oschel\thanksref{bylinetp}}
21: 
22: \address{Arbeitsgruppe Nichtlineare Dynamik, Universit\"at Potsdam, Am
23:   Neuen\\ Palais, D--14415 Potsdam, Germany.}
24: 
25: \author{Volkhard Buchholtz\thanksref{bylinevb}}
26: 
27: \address{Humboldt--Universit\"at zu Berlin, Institut f\"ur
28:   Physik, Unter den Linden 6,\\ D--10099 Berlin, Germany}
29: 
30: \thanks[bylinetp]{E-mail: thorsten@hlrsun.hlrz.kfa--juelich.de}
31: \thanks[bylinevb]{E-mail: volkhard@itp02.physik.hu--berlin.de}
32: \date{16 July 1995}
33: 
34: \maketitle
35: 
36: \begin{abstract}
37:   We propose a new model for the description of complex granular
38:   particles and their interaction in molecular dynamics simulations of
39:   granular material in two dimensions. The grains are composed of
40:   triangles which are connected by deformable beams. Particles are
41:   allowed to be convex or concave. We present first results of
42:   simulations using this particle model.
43: \end{abstract}
44: \end{frontmatter}
45: 
46: \section{Introduction}
47: Molecular dynamics has been proven to be a well suited method to
48: investigate the dynamic and static behavior of granular matter. The
49: concept of molecular dynamics was initially used to simulate the
50: dynamics of atoms and molecules, i.e.~of particles without inner
51: degrees of freedom. Alder and Wainwright who belong to the inventors
52: of this method investigated already 1957 numerically the low density
53: approximation of the entropy $S=-N~k_B~\langle \ln~f(p,t)\rangle$ ($f$
54: is the one--particle probability density in momentum space) in a hard
55: sphere system consisting of $N=100$
56: particles~\cite{AlderWainwright:1957ug}. There are applications of
57: molecular dynamics in many fields of interest and molecular dynamics
58: became one of the standard methods in computational physics. Recent
59: molecular dynamics simulations are of very high algorithmic
60: complexity, and systems consisting of up to $10^9$ particles
61: (with simple interaction forces) have been simulated (see
62: e.g.~\cite{BeazleyLomdahl:1993,BeazleyLomdahlJensenTamayo:1994,BeazleyEtAl:1995}). A
63: very interesting overview on the historical development is given in
64: the first part of Hoover's book~\cite{Hoover:1992}.
65: 
66: In molecular dynamics of granular particles, sometimes called
67: ``granular dynamics'', each of the ``molecules'' is a macroscopic body
68: typically with a diameter of the order $D\approx 0.1\dots 1~mm$ which
69: has its own thermodynamic properties, i.e.~it has internal degrees of
70: freedom and hence it can dissipate mechanical energy when it is
71: subjected to outer forces. In some special cases as in simulations of
72: planetary rings~\cite{Salo:1992,Weidenschilling} the diameter can also
73: be of the order $D\approx 1~mm\dots 10~m$. In the most simple case one
74: can express the inelastic nature of the collisions of the grains by
75: introducing restitution coefficients which describe the relative
76: velocities in normal and tangential direction $\vec{v}_{ij}^{\,N}$ and
77: $\vec v_{ij}^{\,T}$ of two granular particles $i$ and $j$ after a
78: collision as functions of the relative velocities before the collision
79: $\vec V_{ij}^N$ and $\vec V_{ij}^T$
80: %\begin{mathletters}
81: \begin{eqnarray}
82:   \vec v_{ij}^{\,N} &=& -\epsilon_N\cdot \vec V_{ij}^N \hspace{0.5cm}
83:   \left( 0\le \epsilon_N \le 1\right) 
84:   \label{resitutionEQa}\\ 
85:    \vec v_{ij}^{\,T} &=&
86:     -\epsilon_T\cdot \vec V_{ij}^T \hspace{0.5cm} \left( -1\le
87:     \epsilon_T \le 1\right) ~.
88:   \label{resitutionEQb}
89: \end{eqnarray}
90: %\end{mathletters}
91: Here the relative velocity is given by
92: \begin{equation}
93:   \vec{v}_{ij} = \dot{\vec{r}}_i -\dot{\vec{r}}_j 
94:   \label{relvelocity1}
95: \end{equation}
96: or, including the particle rotation with angular velocities $\vec
97: \omega_i$ and $\vec \omega_j$ , by
98: \begin{equation}
99:   \vec v_{ij} = \left( \dot{\vec r}_i + \omega_i\times R_i \left(-\vec{n}\right)
100: \right) - \left( \dot{\vec r}_j + \omega_j\times R_j 
101: \vec{n} \right)
102:   \label{relvelocity2}
103: \end{equation}
104: with the unity vectors
105: %\begin{mathletters}
106:   \begin{eqnarray}
107:     \vec{n} &=& \frac{\vec r_i -\vec r_j }{\left| \vec r_i -\vec r_j
108:     \right|}\\ 
109:     \vec{t} &=& \left( \begin{tabular}{ll} 0 & -1 \\ 1 & 0
110: \end{tabular} \right) \cdot \frac{\vec r_i -\vec r_j }{\left| \vec r_i
111:   -\vec r_j \right|}~.
112: %    \label{unityvectors}
113:   \end{eqnarray}
114: %\end{mathletters}
115: Hence the relative velocities at the point of contact are
116: %\begin{mathletters}
117:   \begin{eqnarray}
118:     \vec v_{ij}^{\,N} &=& \vec n \left(\vec v_{ij} \cdot \vec n \right)\\
119:     \vec v_{ij}^{\,T} &=& \vec t \left(\vec v_{ij} \cdot \vec t \right) ~.
120: %    \label{relvelocities}
121:   \end{eqnarray}
122: %\end{mathletters}
123: In general the normal and tangential restitution coefficients $
124: \epsilon_N $ and $ \epsilon_T $ are themselves functions of the
125: relative impact velocities $\epsilon_N=\epsilon_N\left(
126: V_{ij}^N,~V_{ij}^T\right) $ and $\epsilon_T=\epsilon_T\left(
127: V_{ij}^N,~V_{ij}^T\right) $. For the case that the particles are rough
128: homogeneous spheres the coefficients $\epsilon_N$ and $\epsilon_T$ can
129: be calculated
130: analytically~\cite{BrilliantovSpahnHertzschPoeschel:1994} solving the
131: viscoelastic equations of the particles. Using this model one neglects
132: the complicated interaction between particles during the collision.
133: The approximation is justified if collisions are relatively rare
134: events, i.e. if the mean free path between collisions is long and one
135: can restrict the calculation to two--particle--interactions only. If a
136: system is in this regime one calls it a ``granular gas''. Event driven
137: algorithms have been applied in many papers, recently e.g.
138: in~\cite{GoldhirschZanetti:1993,McNamaraYoung:1992} for an
139: investigation of inelastic clustering of granular particles (using
140: constant restitution coefficients) and by Luding et al.~to simulate
141: vertically shaken
142: material~\cite{LudingClementBlumenRajchenbachDuran:1994DISS,LudingClementBlumenRajchenbachDuran:1994STUDIES}.
143: For sophisticated implementations of such algorithms
144: see~\cite{Rapaport:1980,Lubachevsky:1991}.
145: 
146: In many cases of interest, however, the particles collide very often
147: or they touch each other permanently, i.e. the system reveals static
148: properties. Examples for such systems are flows through pipes and
149: hoppers, ball mills, shaken materials, sand heaps and many others. In
150: the model proposed by Cundall and Strack~\cite{CundallStrack:1979} and
151: Haff and Werner~\cite{HaffWerner:1986} the particles feel restoring
152: forces in normal and tangential direction, and they lose kinetic
153: energy during the collisions. If the particles $i$ and $j$ at the
154: positions $\vec{r}_i$ and $\vec{r}_j$ with radiuses $R_i$ and $R_j$
155: touch each other they feel the force
156: %\begin{mathletters}
157: \begin{equation}
158:   \vec{F}_{ij}= F_{ij}^N\, \vec{n} + F_{ij}^T\, \vec{t}
159: \end{equation}
160: with 
161: \begin{eqnarray}
162:   F_{ij}^N &=& Y \cdot \left(R_i+R_j- \left| \vec{r}_i-\vec{r}_j\right| 
163:   \right)^{\frac{3}{2}} - m_{ij}^{eff} \cdot \gamma_N \cdot 
164:   (\dot{\vec{r}}_i - \dot{\vec{r}}_j) \cdot\vec{n}
165: \label{fnormal}\\ 
166: F_{ij}^T &=& \mbox{sign} (v_{ij}^{rel}) \cdot \min\left( m_{ij}^{eff} \gamma_T \left|
167: v_{ij}^{rel}\right| , \mu \left| F_{ij}^{N}\right| \right) 
168: \label{ftang} \\
169: v_{ij}^{rel} &=& (\dot{\vec{r}}_i - \dot{\vec{r}}_j)  \cdot
170: \vec{t}+ R_i \cdot \Omega_i + R_j \cdot \Omega_j 
171: \label{surfvelocity}\\ 
172: m_{ij}^{eff} &=& \frac{m_i \cdot m_j}{m_i + m_j} ~.
173: \label{meff}
174: \end{eqnarray}
175: %\end{mathletters}
176: $Y$ is the Young modulus, $\gamma_N$ and $\gamma_T$ are the damping
177: coefficients in normal and tangential direction and $\mu$ is the
178: Coulomb friction coefficient. Eq.~(\ref{surfvelocity}) describes the
179: relative velocity of the surfaces of the particles at the point of
180: contact and eq.~(\ref{meff}) gives the effective mass.
181: Eq.~(\ref{ftang}) takes the Coulomb friction law into account, saying
182: that two particles slide on top of each other if the shear force
183: overcomes $\mu$ times the normal force. Eq.~(\ref{fnormal}) comes
184: from the Hertz law \cite{Hertz:1882ug} for the force between two rigid
185: spheres which are in contact with each other.
186: 
187: This model has been successfully applied to simulate the behavior of
188: dry granular material in many works describing various physical
189: phenomena. It developed to the standard method for calculations of
190: granular dynamics. Some of these problems are size segregation and
191: convection in vibrated material in two and three dimensions
192: (e.g.~\cite{GallasHerrmannSokolowski:1992PRL,Taguchi:1992PRL,GallasHerrmannPoeschelSokolowski:1994,PoeschelHerrmann:1994}),
193: the flow in hoppers
194: (e.g.~\cite{RistowHerrmann:1994,FormKohringMelinPuhlTillemans:1994,CampbellPotapov:1993})
195: and pipes (e.g.~\cite{Poeschel:1994,Lee:1994}), the flow in rotating
196: cylinders
197: (e.g.~\cite{Ristow:1994,RistowCantelaubeBideau:1994ug,PoeschelBuchholtz:1993CSF}),
198: the motion of granular material on an inclined
199: surface~\cite{Walton:1991,Poeschel:1993JDP}, sound propagation in
200: granular material~\cite{Melin:1994}, the onset of
201: turbulence~\cite{Taguchi:1992JDP,Taguchi:1994} and many others.  Many
202: experimental results of many authors could be reproduced numerically
203: using this type of molecular dynamics simulations. There have been
204: developed very efficient algorithms for the case of short range
205: interaction with large force gradients as it is typical for granular
206: materials,
207: e.g.~\cite{BuchholtzPoeschel:1993JMPC,FormItoKohring:1993,EveraersKremer:1993}.
208: The model by Cundall and Strack yields reliable results in more
209: dynamic systems, i.e. when the static friction of the particles does
210: not play a major role. When the static properties of the material
211: govern the behavior, however, the model might fail.
212: In~\cite{BuchholtzPoeschel:1994PA} is shown that one cannot build up a
213: stable sand heap with a finite inclination of such particles but the
214: heap dissolves under its own mass and the angle becomes smaller with
215: rising number of particles.  Lee~\cite{Lee:1993} simulated static
216: friction effects using spheres which are connected by springs. When
217: two particles touch each other for a certain time a spring ``grows''
218: between these particles and keeps the particles together when forces
219: are applied which would separate them. When the separating forces
220: become too large the springs break and the particles move freely.
221: Using this model Lee reproduced the finite angle of repose in a sand
222: heap, however, the growing and breaking springs have no direct
223: equivalent in nature, and hence the model seems to be quite
224: artificial.
225: 
226: Another simulation method which corresponds to the previous one for
227: the case of very large damping has been introduced by Visscher and
228: Bolsterli~\cite{VisscherBolsterli:1972ug}. In each iteration step all
229: of the grains are moved according to the motion of a vessel containing
230: the granular material. Then the particles are released one by one,
231: beginning with the lowest one (with respect to the direction of
232: gravity). Each particle moves until it reaches the next local minimum
233: of the potential energy when it hits either the wall or other
234: previously released grains. When a particle reached its local minimum
235: it stays in that position, i.e.~a falling
236: particle cannot cause the motion of another previously released
237: particle. This behavior means that inertia of the moving particles is
238: neglected as soon as they hit another particle or the wall, and hence
239: it corresponds to the case of very large damping. The advantage of
240: this algorithm is its numerical simplicity, i.e.~it is possible to
241: investigate much larger systems than using molecular dynamics.
242: Recently it has been applied in simulations of several problems as
243: size
244: segregation~\cite{JullienMeakinPawlovitch:1992}
245: and the motion of particles in a rotating
246: cylinder~\cite{BaumannJobsWolf:1993,BaumannJanosiWolf:1994}. Obviously
247: the simplification of infinite damping is not valid in each case, and
248: hence one has to carefully investigate whether it is justified to apply
249: this algorithm (see
250: e.g.~\cite{BarkerMehta:1995}).
251: 
252: There are some more simulation methods for the dynamics of granular
253: material which we want to mention only: Peng and Herrmann applied a
254: Lattice Gas Automaton to investigate density waves in a
255: pipe~\cite{PengHerrmann:1994}. Caram and Hong investigated granular 
256: flows using a Random Walk technique~\cite{CaramHong:1991}.
257: 
258: 
259: There are some models to simulate complex shaped grains which are
260: composed of spheres. In the model by Gallas and Soko{\l}owski
261: \cite{GallasSokolowski:1993} the grains consist of two spheres
262: connected with each other by a stiff bar. Walton and Braun applied
263: more complicated particles consisting of four or eight spheres rigidly
264: connected with each other~\cite{WaltonBraun:1993}. Using this model
265: they examined the transition from stationary to sliding flow and the
266: transition from sliding to raining flow in a rotating drum. P\"oschel
267: and Buchholtz~\cite{PoeschelBuchholtz:1993,BuchholtzPoeschel:1994PA}
268: describe grains built up of five spheres where one of them is located
269: in the center of the grain and four identical spheres are at the
270: corners of a square. Each pair of neighboring spheres is connected by
271: a damped spring. The latter model was applied to the rotating cylinder
272: \cite{PoeschelBuchholtz:1993CSF} and it was shown that the simulation
273: results agree much better with the experiment than equivalent
274: calculations with spheres. Especially it was shown that one can
275: reproduce stick--slip motion and avalanches which was not possible in
276: simulations using spheres. The inclination of the surface of the
277: material in the rotating cylinder and the dependency of the
278: inclination on the angular velocity, however, did not agree well with
279: the experiment.
280: 
281: Mustoe and DePorter~\cite{MustoeDePorter:1993} proposed a particle
282: model where the boundary geometry of the particles is defined (in
283: local coordinates) by
284: \begin{equation}
285: f_i\left( x_i,y_i\right) =\left[\frac{\left| x_i \right|}{a_i}\right]^{n_i} + 
286:              \left[\frac{\left| y_i \right|}{b_i}\right]^{n_i} - 1 = 0.
287:  \label{hogue.eq}
288: \end{equation}
289: Varying $n_i$ between 2 and $\infty$ the shape of the $i$th particles
290: varies continuously from elliptical to rectangular. To detect contacts
291: of pairs of particles described by eq.~(\ref{hogue.eq}) one has to
292: solve numerically equations of high order for each pair of possibly
293: touching particles in each iteration step. This requires an iteration
294: technique which converges slowly for higher exponents $n_i$ and which
295: makes hence the algorithm numerically complicated. Hogue and
296: Newland~\cite{HogueNewland:1993} investigated the flow of granular
297: material on an inclined chute and through a hopper using a model where
298: the boundaries of the convex particles are given by polygons with up
299: to 24 vertices each. To detect whether two particles touch each other
300: one has to calculate the intersections between each pair of vertices.
301: During collisions energy is dissipated according to Stronge's energy
302: dissipation hypothesis~\cite{Stronge:1990} in normal direction, whilst
303: Coulomb's friction law models the energy losses also in tangential
304: direction.
305: 
306: Using grains which consist of interconnected spheres or of particles
307: described by eq.~(\ref{hogue.eq}) it is not possible to simulate
308: particles with sharply formed corners. For some effects it seems to be
309: essential to simulate such particles to reproduce the experimental
310: observed effects. This point is discussed in detail
311: in~\cite{BuchholtzPoeschel:1994PA}. Tillemans and
312: Herrmann~\cite{TillemansHerrmann:1994,Tillemans:1994} proposed a two
313: dimensional particle model where the particles are convex polygons.
314: When two particles $i$ and $j$ touch each other, i.e. when there is an
315: overlap, there acts the force
316: %\begin{mathletters}
317: \begin{eqnarray}
318:   \vec{F}_{ij}&=&F_{ij}^N~\vec{e}^{\,N} + F_{ij}^T~\vec{e}^{\,T}\\
319:   F_{ij}^N &=& \frac{Y~A}{L_c} -
320:   m_{ij}^{eff}~\gamma_N~\left( \vec{v}^{\,rel}\cdot \vec{e}^{\,N} \right) \\ 
321:   F_{ij}^T &=& - \min\left( m_{ij}^{eff} \gamma_{T}
322:   \left|\vec{v}_{rel} \cdot \vec{e}^{\,T} \right|, \mu \left|
323:     F_{ij}^{N} \right|\right)~.
324: %  \label{forcetillemans}
325: \end{eqnarray}
326: %\end{mathletters}
327: $A$ is the compression area of the particles
328: (overlap area), Y is the Young module of the material.
329: The effective mass $m_{ij}^{eff}$ is given by eq.~(\ref{meff}), the
330: relative particle velocity is $\vec{v}^{\,rel}= \dot{\vec{r}}_i
331: -\dot{\vec{r}}_j $, where $\vec{r}_i$ points to the center of mass
332: of the particle $i$, and $L_c$ is the characteristic size of the
333: particle. The unit vectors are in the direction of the line that
334: connects the intersection points of the overlapping particles $i$ and
335: $j$ ($\vec{e}^{\,T}$) and perpendicular to this line
336: ($\vec{e}^{\,N}$). Only convex particles are allowed. This model was
337: used e.g.~in simulations of shear cells, earth quakes and flow through
338: hoppers~\cite{Tillemans:1994}. The results differ significantly from
339: similar calculations using spheres. For the case of the hopper
340: simulation they found clogging and arching which could not be found in
341: simulations with spheres. A similar model was used earlier by
342: Handley~\cite{Handley:1993} who investigated the fracture behavior of
343: brittle granular material. Potapov et
344: al.~\cite{PotapovCampbellHopkins:1994aug,PotapovCampbellHopkins:1994bug}
345: investigated a similar model for solid fracture. Initially they
346: subdivide a macroscopic two dimensional body into many small
347: equilateral triangles (elements). The forces in the body are resolved
348: as forces at inter--element contacts of two different types which they
349: call ``glued'' and ``collisional''. Glued contacts are the joints
350: between elements interior to a solid body which can support stresses.
351: When the tensile stresses exceed a certain value, cracks form and the
352: glued bond breaks. Collisional contacts are contacts between the
353: surface elements of the body (which eventually collides with other
354: bodies or walls), and contacts between triangles in the bulk of the
355: body where glued contacts have been broken. In the three latter
356: models~\cite{TillemansHerrmann:1994,Tillemans:1994,Handley:1993,PotapovCampbellHopkins:1994aug,PotapovCampbellHopkins:1994bug}
357: the convex grains have been Voronoi polyhedra in two
358: dimensions~\cite{Finney:1979}.
359: 
360: In this paper we present a new particle model for molecular dynamics
361: of granular material in two dimensions where the particles are
362: simulated using a similar model which has some advantages. In the
363: following section we describe the model, the forces acting between the
364: grains are derived in sections~\ref{trianglesSEC} and
365: \ref{stressesinbeamsSEC}. Some aspects of the implementation and the
366: performance of the algorithm are discussed in
367: section~\ref{implementationSEC} and some sample results based on this
368: particle interaction model will be briefly discussed in
369: section~\ref{resultsSEC}.
370: 
371: \section{The model}
372: In our model the grains are composed of an arbitrary number of
373: ideal elastic triangles which are connected by beams (see
374: fig.~\ref{grainsFIG}). 
375: \begin{figure}[ht]
376:   \ifcase \value{nofigs}
377:   \begin{center}
378:     \begin{minipage}{4cm}
379:       \psfig{figure=figs/grains.eps,width=4cm,angle=270}
380:     \end{minipage}
381:     \hspace{2cm}
382:     \begin{minipage}{2cm}
383:       \psfig{figure=figs/grain5.eps,width=2cm,angle=270} 
384:       \vspace{0.5cm}
385:       
386:       \psfig{figure=figs/grainstern.eps,width=2cm,angle=270}
387:     \end{minipage}
388:   \end{center}
389:   \vspace{0.5cm}
390:   
391:   \fi
392:   \caption{Examples of grains composed of different numbers of triangles. 
393:     The model is not restricted to convex grains.}
394:   \label{grainsFIG}
395: \end{figure}
396: 
397: A beam in our sense is an deformable damped bar
398: which is subjected to forces in the direction of its axis and
399: transverse to its axis, i.e.~to normal and shear forces, and to
400: moments acting on its ends. Triangles which belong to the same grain
401: do not interact with each other. When two triangles of different
402: grains collide, i.e.~if there is an overlap between both, they feel a
403: restoring force. Hence the force acting upon the triangle $i$
404: belonging to the grain $j$ ($j\in
405: \{1,2,\dots,N\}$) is
406: \begin{equation}
407: \vec{F}_i^{\,j} = \sum_{l=1,l\neq j}^N 
408:       \sum_{k=1}^{n_t(l)} \vec{\Gamma}_{ik}^{jl} 
409:    + \sum_{k=1}^{n_b^j(i)} \vec\Lambda_{ik}^{\,j} 
410:    + \vec\Phi(\vec{r}_{i}^{\,j}, \dot{\vec{r}}_i^{\,j})~.
411:   \label{totalforceEQ}
412: \end{equation}
413: The first sum in the first term in eq.~(\ref{totalforceEQ}) runs over
414: all grains $l$ except the $j$th, the second sum runs over all
415: triangles $k\in\{1,2,\dots,n_t(l)\}$, the $l$th grain consists of.
416: $\vec{\Gamma}_{ik}^{jl}$ is the force which acts between the triangles
417: $i$ and $k$ which belong to different grains $j$ and $l$ (see section~\ref{trianglesSEC}). The sum in
418: the second term runs over all beams connecting the triangle $i$ of the
419: grain $j$ with other triangles of the same grain.
420: $\vec\Lambda_{ik}^{\,j}$ is the force induced by the beam $k$ acting
421: on the $i$th triangle of the grain $j$. They originate from the
422: distortion of the beam that connects the triangle $i$ with another one
423: of the same grain. The acting forces and momenta will be discussed
424: below in section~\ref{stressesinbeamsSEC}. A similar model for the
425: beam forces has been introduced earlier by Herrmann et al. for the
426: fracture of disordered lattices~\cite{HerrmannHansenRoux:1989}.  The
427: last term $\vec\Phi(\vec{r}_{i}^{\,j}, \dot{\vec{r}}_i^{\,\,j})$
428: describes the action of an external force, as e.g.~gravity, on the
429: triangles.
430: 
431: During its deformation a beam dissipates energy similar to a linearly
432: damped spring, i.e.~proportional to its deformation rate~(see
433: section~\ref{stressesinbeamsSEC}). To avoid time consuming
434: calculations due to Steiner's law each beam is fixed at both ends in
435: the center of mass points of the triangles it connects. When the beam
436: is not deformed by eventually applied forces (the beam is at rest) it
437: is perpendicular to the neighboring edges of the triangles which it
438: connects.  Fig.~\ref{grainsFIG} shows some examples of grains. The
439: only restriction concerning the number, the shape and the position of
440: the triangles is caused by the condition that the beams are fixed in
441: the center of mass.
442: 
443: The next section~\ref{trianglesSEC} gives a detailed description of the
444: force that acts when triangles collide, in
445: section~\ref{stressesinbeamsSEC} the forces due to deformed beams are
446: derived.
447: 
448: \section{The interaction of the triangles}
449: \label{trianglesSEC}
450: In this section we describe the calculation of the term
451: $\vec{\Gamma}_{ik}^{jl}$ in eq.~(\ref{totalforceEQ}) originating from
452: the compression of two triangles $i$ and $k$ which belong to the
453: grains $j$ and $l$, respectively. When the triangles are compressed
454: there exists a virtual ``overlap area'' which leads to an elastic restoring
455: force $\vec{\Gamma}_{ik}$. For simplicity here and in the following we
456: drop the upper indexes of the variables. Areas of triangles $XYZ$ and
457: quadrangles $WXYZ$ we denote by $\Delta(XYZ)$ and $\Box(WXYZ)$.
458: Fig.~\ref{standard.fig} shows the most frequently found type of a
459: collision between the triangles $A_iB_iC_i$ and $A_kB_kC_k$, i.e.~a
460: corner of one triangle deforms a side of another one. 
461: 
462: \begin{figure}[ht]
463:   \ifcase \value{nofigs}
464: \centerline{\psfig{figure=/figs/standard.eps,width=6cm,angle=270}}
465: \vspace{0.5cm}
466: \fi
467: \caption{The ``standard'' type of a collision 
468:   between two triangles $A_iB_iC_i$ and $A_kB_kC_k$. The forces
469:   $\pm\vec{\Gamma}_{ik}$ are directed perpendicular to the
470:   intersection line $\overline{S_1S_2}$.  Moments act with respect to
471:   the center of mass points $G_i$ and $G_j$.  The absolute value of
472:   the interaction force $\left|\vec\Gamma_{ik}\right|$ is proportional
473:   to the shadowed intersection area $\Delta \left(S_1S_2A_k\right)$.}
474: \label{standard.fig}
475: \end{figure}
476: 
477: The absolute
478: value of the restoring force $\vec{\Gamma}_{ik}$ is given by the
479: shadowed area $\Delta\left(S_1S_2A_k\right)$ of the triangle
480: $S_1S_2A_k$ times the Young module $Y$. Its direction is
481: perpendicular to the intersection line $\overline{S_1S_2}$ (Poisson
482: hypothesis, see e.g.~\cite{FormKohringMelinPuhlTillemans:1994}). Hence
483: resulting momenta acting upon the triangles $A_iB_iC_i$ and
484: $A_kB_kC_k$ read
485: %\begin{mathletters}
486: \begin{eqnarray}
487: \vec{M}\left(A_iB_iC_i\right) &=& \vec{HG_i} \times \vec{\Gamma}_{ik} \label{momentaA}
488: \\
489: \vec{M}\left(A_kB_kC_k\right) &=& \vec{HG_k} \times \vec{\Gamma}_{ki} = 
490:   -\vec{HG_k}\times \vec{\Gamma}_{ik}~.
491: \label{momentaB}
492: \end{eqnarray}
493: %\end{mathletters}
494: The vector $\vec{HG_i}$ points from the middle point of the line
495: $\overline{S_1S_2}$ to the center of mass of the triangle $A_iB_iC_i$.
496: 
497: Although the case shown in fig.~\ref{standard.fig} is the mostly
498: occurring type of collisions there are several other types which
499: treatment has to be discussed below.
500: 
501: Classifying the possible interactions of the triangles we define five
502: different types of collisions which are slightly different from those
503: in~\cite{PotapovCampbellHopkins:1994aug}. The detection of overlapping
504: polygons is a very important problem in computer graphics too, where
505: one frequently has to decide whether two objects cover each other, and
506: which of them is in the foreground or in the background, respectively.
507: Therefore we found it helpful and inspiring first to have a look into
508: advanced methods in computer graphics,
509: e.g.~\cite{BentleyOttmann:1979ug,PreparataShamos:1985ug}. In all
510: figures~\ref{standard.fig}--\ref{schnit5} the overlapping area is
511: drawn extremely exaggerated. Typically the overlapping area of two
512: interacting triangles does not exceed a few tenth of a percent of the area
513: of the triangles.
514: 
515: \begin{enumerate}
516: \begin{figure}[ht]
517:   \ifcase \value{nofigs}
518:   \centerline{\psfig{figure=/figs/schnitt1a.eps,width=6cm,angle=270}\psfig{figure=/figs/schnitt1b.eps,width=6cm,angle=270}}
519:   \vspace{0.5cm}
520: \fi
521: \caption{The first type of collisions: There are two intersection 
522: points $S_1$ and $S_2$. Both lie at the same edge of one of the 
523: triangles.}
524: \label{schnit1}
525: \end{figure}
526: 
527: \item{For the first type (fig.~\ref{schnit1}) there are two
528:     intersection points $S_1$ and $S_2$ lying at the same edge of one
529:     of the interacting triangles. One of the situations drawn in
530:     fig.~\ref{schnit1} corresponds to the standard collision
531:     (fig.~\ref{standard.fig}). We calculate the area $\Delta \left( S_1
532:       S_2 A_i\right)$ (shadowed area) given by the intersection points
533:     $S_1$ and $S_2$ and the included point $A_i$. Because the
534:     overlapping area is definitely smaller than half the area of the
535:     triangles the force is proportional to the minimum between $\Delta
536:     \left( S_1 S_2 A_i\right)$ and $\Delta \left( A_i B_i C_i\right) -
537:     \Delta \left( S_1 S_2 A_i\right)$
538: \begin{equation}
539:   \left|\vec{\Gamma}_{ik}\right| = \Gamma_{ik} = 
540:   Y \cdot \min \left\{\Delta \left( S_1 S_2 A_i\right) , \Delta \left(
541:   A_i B_i C_i\right) - \Delta \left( S_1 S_2 A_i\right) \right\}~.
542: \end{equation}
543: The force acts perpendicular to the line between the intersection
544: points $\overline{S_1 S_2}$.}
545: 
546: \item The second type (fig.~\ref{schnit2}) is characterised by two
547: intersection points lying on different edges for both triangles. We
548: calculate the areas $\Delta \left( S_1 S_2 A_i\right) $ (dark gray)
549: and $\Delta \left( S_1 S_2 A_k\right) $ (light gray). The force is
550: proportional to the overlap
551: \begin{eqnarray}
552: \left|\vec{\Gamma}_{ik}\right| = \Gamma_{ik} = Y \cdot \left[
553: \min \left\{\Delta \left( S_1 S_2 A_i\right) , \Delta 
554: \left( A_i B_i C_i\right) - \Delta \left( S_1 S_2 A_i\right) 
555: \right\} + \right.\nonumber \\
556: \left. \min \{\Delta \left( S_1 S_2 A_k\right) , \Delta 
557: \left( A_k B_k C_k\right) - \Delta \left( S_1 S_2 A_k\right) \}
558: \right]~.
559: \end{eqnarray}
560: It acts perpendicular to the line between the intersection points
561: as in the previous case.
562: \begin{figure}[ht]
563:   \ifcase \value{nofigs}
564: \centerline{\psfig{figure=/figs/schnitt2a.eps,width=6cm,angle=270}\hspace{0.5cm}\psfig{figure=/figs/schnitt2b.eps,width=6cm,angle=270}}
565: \vspace{0.5cm}
566: \fi
567: \caption{The second type of collisions: There are two intersection 
568: points $S_1$ and $S_2$ lying on different edges for both triangles.}
569: \label{schnit2}
570: \end{figure}
571: 
572: 
573: \item {The third type (fig.~\ref{schnit3}) is characterised by four
574:     intersection points, each pair of two points lies on one edge of
575:     both triangles. Therefore the third edge of both interacting
576:     triangles does not intersect the edges of the other triangle.
577:     There are two pairs of forces proportional to the quadrangular
578:     area $\Box \left( S_1 S_2 S_3 S_4\right) $ of the overlap
579: \begin{equation}
580: \left|\vec{\Gamma}_{ik}\right| = \Gamma_{ik}  = 
581: \frac{1}{2} \cdot Y \cdot \Box \left( S_1 S_2 S_3 S_4\right)~, 
582: \end{equation}
583: one acting perpendicular to the line $\overline{S_1 S_2}$ and the
584: other perpendicular to the line $\overline{S_1 S_4}$. The pre--factor
585: $0.5$ is due to two pairs of acting forces instead of one for the first two
586: collision types.}
587: \begin{figure}[ht]
588:   \ifcase \value{nofigs}
589: \centerline{\psfig{figure=/figs/schnitt3.eps,width=6cm,angle=270}}
590: \vspace{0.5cm}
591: \fi
592: \caption{The third type of collisions: There are four intersection 
593: points $S_1$, $S_2$, $S_3$ and $S_4$. Each pair of points lies on 
594: one edge of either the first or the second triangle.}
595: \label{schnit3}
596: \end{figure}
597: 
598: \item{The fourth type (fig.~\ref{schnit4}) is characterised by four
599:     intersection points too, but here one of the triangles has
600:     intersection points at all its edges. Because interactions of this
601:     type occur extremely seldomly we did not implement an exact
602:     calculation for this case, but we calculate two interactions of
603:     type $1$ with the intersection points $S_1$--$S_2$ or $S_3$--$S_4$
604:     respectively, instead of solving the correct problem. In real
605:     simulation the fraction of this type occurs approximately
606:     $10^{-5}$ of the number of collisions.  Nevertheless one has to
607:     deal with these extremely seldom events since otherwise they might
608:     cause problems from which the system cannot recover.  If one does
609:     not care about these events usually the system gains in a few time
610:     steps after the event occurs a huge amount of kinetic energy,
611:     i.e.~the system explodes.}
612: \begin{figure}[ht]
613:   \ifcase \value{nofigs}
614: \centerline{\psfig{figure=/figs/schnitt4.eps,width=6cm,angle=270}}
615: \vspace{0.5cm}
616: \fi
617: \caption{The fourth type of collisions: There are four intersection 
618: points $S_1$, $S_2$, $S_3$ and $S_4$. One of the triangles intersects 
619: with all its three edges.}
620: \label{schnit4}
621: \end{figure}
622: 
623: \item{The fifth and last type of interaction (fig.~\ref{schnit5}) is
624: characterised by six intersection points $S_1$ -- $S_6$. Equivalent to
625: the forth type we do not calculate the exact interaction but
626: substitute this calculation by solving three interactions of the first
627: type with pairs of interaction points $S_1$--$S_2$, $S_3$--$S_4$ and
628: $S_5$--$S_6$}.
629: \end{enumerate}
630: According to the forces $\vec{\Gamma}_{ik}$ we derive forces in
631: parallel to the axes of the Cartesian coordinate system and moments
632: $\vec{M}$ acting with respect to the center of mass points $G_i$ and
633: $G_k$ as described in equations~(\ref{momentaA}, \ref{momentaB}).
634: \begin{figure}[ht]
635:   \ifcase \value{nofigs}
636: \centerline{\psfig{figure=/figs/schnitt5.eps,width=6cm,angle=270}}
637: \fi
638: \caption{The fifth type of collisions: There are six intersection
639: points $S_1$,\dots,$S_6$.\hspace{2cm}}
640: \label{schnit5}
641: \end{figure}
642: 
643: \section{Stresses in beams}
644: \label{stressesinbeamsSEC}
645: When torques and forces are applied to a beam the beam deforms. Since
646: all deformations are assumed to be small compared with the size of the
647: beam one can superpose the deformations originating from different
648: applying forces and
649: torques~\cite{TimoshenkoGere:1972,TimoshenkoLessells:1928}.
650: Fig.~\ref{bending.fig} shows an (infinitesimally) deformed beam with
651: radius of curvature $\rho$. We find the approximation
652: \begin{equation}
653: \tan\left( \frac{d\Theta}{2}\right) =\frac{1}{2 \cdot \rho}\cdot
654: dx \hspace{0.5cm}\mbox{or}\hspace{0.5cm} 
655: \rho=\frac{dx}{d\Theta}
656: \end{equation}
657: with $\tan(d\Theta) = d\Theta$.
658: \begin{figure}[ht]
659:   \ifcase \value{nofigs}
660: \centerline{\psfig{figure=/figs/bending.eps,width=8cm,angle=270}}
661: \vspace{0.5cm}
662: \fi
663: \caption{An infinitesimally deformed beam with radius of curvature 
664: $\rho $. The fibres below or above the neutral fibre $\overline{SS}$ 
665: are in tension or in compression. For a beam with quadratic cross 
666: section $b^2$ the deformation leads to the resulting moment 
667: $M=\frac{E\,b^4}{12\,\rho}$.}
668: \label{bending.fig}
669: \end{figure}
670: 
671: The dashed line $\overline{ss}$ is the neutral surface length of which
672: does not change during the deformation. All other fibres below and
673: above the neutral surface are in tension or in compression,
674: respectively. The length of a fibre in distance $y$ from the neutral
675: surface is $\left( \rho+y\right) d\Theta=\left( 1+y/\rho\right) dx$
676: and hence the strain reads $\epsilon=y/\rho$. With Hooke's law
677: $\sigma=E\cdot\epsilon$ one finds $\sigma=Ey/\rho$. Finally the
678: resulting momentum $M$ is
679: \begin{equation}
680: M= \int_A\sigma y dA = \frac{EI}{\rho}
681: \label{M=EI/rho}
682: \end{equation}
683: where the moment of inertia $I=\int y^2 dA$ depends on the geometrical
684: shape of the cross section of the beam. In the present paper we assume
685: that the beams have quadratic cross section of width $b$ and hence
686: $I=b^4/12$. When we express the radius of curvature $\rho$ by
687: \begin{equation}
688: \frac{1}{\rho} = \frac{d^2\,v}{dx^2}~,
689:   \label{curvature}
690: \end{equation}
691: where $v$ is the deflection of the beam from its initial position
692: (with the approximation $\tan{\Theta\approx \Theta}$).
693: From eqs.~(\ref{M=EI/rho}) and (\ref{curvature}) one finds the basic
694: equation for the deflection of a beam:
695: \begin{equation}
696: \frac{d^2\,v}{dx^2} = v'' = +\frac{M}{E\,I}
697:   \label{basicequation}
698: \end{equation}
699: 
700: This equation has to be solved to find the forces and moments which
701: act when a beam is deformed. The integration constants are used to
702: satisfy the boundary conditions.
703: 
704: There are three different deformation modes for beams: elongation,
705: shearing and bending. In the following three subsections
706: \ref{elongation.subs}, \ref{shearing.subs} and \ref{bending.subs} we
707: will discuss the deformation rules in detail.
708: 
709: \subsection{Elongation of beams}
710: \label{elongation.subs}
711: Since the exact notation of the acting forces and moments as vector
712: functions of the coordinates of the triangles the beam connects, is
713: not very instructive but confusing due to the length of the formulae
714: expressions we discuss for simplicity of notation the deformation of
715: the beams here and in the following in local coordinates. Instead of
716: providing the exact vector notation we rather discuss absolute values.
717: In all cases the directions in which the forces and moments act are
718: very clear, moreover they are given for each case explicitly in the
719: figures~\ref{fig:beamatrest}--\ref{beamdoppelt.fig} and in
720: figure~\ref{shearbend.fig}. For the implementation of the algorithm,
721: however, one should note that it is necessary to transform the given
722: expressions for the forces and momenta into the coordinate system of
723: the triangles. These transformations require a considerable part of
724: the computation time.
725: \begin{figure}[htbp]
726:   \ifcase \value{nofigs}
727: \centerline{\psfig{figure=/figs/beamatrest.eps,width=8cm,angle=270}}
728: \vspace{0.5cm}
729: \fi
730:     \caption{The shape and the location of a beam at rest. 
731: Its length is assumed to be $L$, its ends $A$ and $B$ lie 
732: on the $x$--axis. When the beam is not deformed there do 
733: not act any forces $F_A$, $F_B$ or moments $M_A$, $M_B$. 
734: In the following 
735: \ifcase \value{refcaption}
736: \ref{bending.fig}--\ref{beamdoppelt.fig} 
737: \or
738: figures~8-11
739: \fi
740: and in 
741: \ifcase \value{refcaption}
742: \ref{shearbend.fig} 
743: \or
744: fig.~14
745: \fi
746: the $x$--axis is drawn as a thin horizontal line.}
747: 
748: \label{fig:beamatrest}
749:     
750: \end{figure}
751: 
752: Fig.~\ref{fig:beamatrest} displays the shape and the location of a
753: beam of length $L$ at rest, i.e.~when no deforming forces $F_A$,
754: $F_B$ or moments $M_A$, $M_B$ at the ends $A$ and $B$ of the beam
755: apply. The $x$--axis is drawn with a thin solid line, i.e.~at rest the
756: beam lies on the $x$--axis.
757: 
758: For the elongation deformation we find according to the
759: linear Hooke law the restoring force
760: \begin{equation}
761: \left| F^{el} \right| = \left| \Delta L \right| \cdot E .
762: \end{equation}
763: 
764: \subsection{Shearing}
765: \label{shearing.subs}
766: We want to calculate the moments and forces $M_A$, $M_B$, $F_A$ and
767: $F_B$ of a sheared beam as drawn in Fig.~\ref{beamscherung.fig}.  For
768: a beam of length $L/2$ which is fixed at one end (at $x=0$) and where
769: acts the shear force $F$ (in negative $y$--direction) at the other end
770: (at $x=L/2$) we find for the local moment
771: $M(x)=F\,\left(\frac{L}{2} - x \right)$ and hence
772: eq.~(\ref{basicequation}) reads
773: \begin{equation}
774: v''=\frac{F}{E\,I} \left(\frac{L}{2}-x\right)
775:   \label{shearingeq}
776: \end{equation}
777: with the boundary conditions $v(0)=0$ and $v'(0)=0$.
778: We find for the vertical deviation $\Delta/2$ at the point $x=L/2$
779: \begin{equation}
780: \Delta/2= \frac{FL^3}{24\,EI}
781: \end{equation}
782: force which acts at the free end of the beam.
783: \begin{figure}[ht]
784:   \ifcase \value{nofigs}
785: \centerline{\psfig{figure=/figs/beamscherung.eps,width=8cm,angle=270}}
786: \vspace{0.5cm}
787: \fi
788: \caption{When the beam in 
789: \ifcase \value{refcaption}
790: fig.~\ref{fig:beamatrest} undergoes a shear deformation 
791: $\Delta$ there act the forces $F_A$ and $F_B$ and the moments 
792: $M_A$ and $M_B$. The forces and moments are given in 
793: eqs.~(\ref{forcesshearing.eq}, \ref{momshearing.eq}).
794: %\ref{eq.shearing}).
795: \or 
796: fig.~9 undergoes a shear deformation $\Delta$ there act 
797: the forces $F_A$ and $F_B$ and the moments $M_A$ and $M_B$. 
798: The forces and moments are given in eqs.~(22).
799: \fi
800: }
801: \label{beamscherung.fig}
802: \end{figure}
803: 
804: One can assume that the sheared beam in Fig.~\ref{beamscherung.fig} is
805: built up of two of such beams of length $L/2$ as described above. 
806: Finally one finds
807: \begin{equation}
808: \Delta = \frac{F\,L^3}{12\,EI}
809:   \label{delta}
810: \end{equation}
811: and hence
812: %\begin{mathletters}
813: \begin{eqnarray}
814: F_A &=& -F_B = F = \frac{12 E I}{L^3}\Delta\label{forcesshearing.eq}\\
815: M_A &=& M_B = F \cdot \frac{L}{2} = \frac{6 E I}{L^2} \Delta~.
816: \label{momshearing.eq} 
817: \end{eqnarray}
818: %\end{mathletters}
819: 
820: \subsection{Bending}
821: \label{bending.subs}
822: When a beam undergoes bending deformation
823: (fig.~\ref{beamdoppelt.fig}), the resulting forces and moments can be
824: found by superposing the forces and moments that would act if the beam
825: would be bent at each side separately. Therefore we can restrict
826: ourself on calculating the forces and moments according to the
827: single sided deformation drawn in fig.~\ref{beameinseitig.fig}. 
828: \begin{figure}[ht]
829:   \ifcase \value{nofigs}
830: \centerline{\psfig{figure=/figs/beamdoppelt.eps,width=8cm,angle=270}}
831: \vspace{0.5cm}
832: \fi
833: \caption{A bent beam. The forces $F_A$ and $F_B$ and the moments $M_A$ 
834: and $M_B$ can be calculated by superposing the forces and moments of two 
835: single sided bent beams 
836: \ifcase \value{refcaption}
837: (fig.~\ref{beameinseitig.fig}), one of them bent at the side $A$, 
838: the other one at the side $B$. The results are given in 
839: eqs.~(\ref{momentsbending.eqA}, \ref{momentsbending.eqB}) and (\ref{forcesbending.eq}).
840: \or
841: (fig.~12), one of them bent at the side $A$, the other one at the 
842: side $B$. The results are given in eqs.~(32) and (33).
843: \fi
844: }
845: \label{beamdoppelt.fig}
846: \end{figure}
847: 
848: The deformation in fig.~\ref{beameinseitig.fig}, however, can be
849: understood as a deformation of a beam with free ends according to a
850: single sided moment $M$ (fig.~\ref{auflieger.fig}) superposed by a
851: moment $M_b$ (fig.~\ref{beameinseitig.fig}) that assures that the
852: angle $\Theta_B$ (fig.~\ref{beameinseitig.fig}) results to zero.
853: \begin{figure}[ht]
854:   \ifcase \value{nofigs}
855:   \centerline{\psfig{figure=/figs/beameinseitig.eps,width=8cm,angle=270}}
856: \vspace{0.5cm}
857: \fi
858: \caption{A single sided bent beam. The forces and moments result from the 
859: superposition of the deformation drawn in 
860: \ifcase \value{refcaption}
861: fig.~\ref{auflieger.fig} and the restoring moment $M_B$ 
862: (fig.~\ref{beameinseitig.fig}) which assures the angle 
863: $\Theta_B$ to be zero. The equations for the forces and 
864: moments are given in eqs.~(\ref{onesidebending}) and 
865: (\ref{onesidebendingf}).
866: \or
867: fig.~13 and the restoring moment $M_B$ (fig.~12) which assures the 
868: angle $\Theta_B$ to be zero. The equations for the forces and 
869: moments are given in eqs.~(27) and (29).
870: \fi
871: }
872: \label{beameinseitig.fig}
873: \end{figure}
874: 
875: First we consider one sided bending of the beam
876: (fig.~\ref{beameinseitig.fig}). For a beam as drawn in
877: fig.~\ref{auflieger.fig} where the moment $M$ acts at one of 
878: its both sides one finds easily
879: %\begin{mathletters}
880: \begin{eqnarray}
881: \Theta_A^* &=& \frac{ML}{3EI}
882: \label{thetafree.eqA}
883: \\
884: \Theta_B^* &=& - \frac{ML}{6EI}.
885: \label{thetafree.eqB}
886: \end{eqnarray}
887: %\end{mathletters}
888: \begin{figure}[ht]
889:   \ifcase \value{nofigs}
890:   \centerline{\psfig{figure=/figs/auflieger.eps,width=8cm,angle=270}}
891: \vspace{0.5cm}
892: \fi
893: \caption{If the ends of the beam are not fixed but freely moving, the 
894: applied moment $M$ causes bending due to the angles $\Theta_A$ and 
895: $\Theta_B$ given in 
896: \ifcase \value{refcaption}
897: eqs.~(\ref{thetafree.eqA}, \ref{thetafree.eqB}).
898: \or
899: eqs.~(23).
900: \fi
901: }
902: \label{auflieger.fig}
903: \end{figure}
904: 
905: Since the beam in Fig.~\ref{beameinseitig.fig} has the angle
906: $\Theta_B=0$ we find the moments acting at the ends of the beam in
907: fig.~\ref{beameinseitig.fig} by superposing the moment
908: $M_B$, which causes a virtual angle $\Theta_B^{**}$, fulfilling the
909: condition $\Theta_B^{**}+\Theta_B^{*}=0$:
910: \begin{equation}
911: M_B=\frac{3EI}{L}(- \Theta_B^*)= \frac{M}{2}.
912: \end{equation} 
913: The moment $M_B$ causes the angle 
914: \begin{equation}
915: \Theta_A^{**}=-\frac{\frac{M}{2}L}{6EI}=-\frac{ML}{12EI}
916: \end{equation}
917: and the resulting angle $\Theta_A$ (fig.~\ref{beameinseitig.fig}) is
918: \begin{equation}
919: \Theta_A=\Theta_A^{*}+\Theta_A^{**}=\frac{M_AL}{4EI}~.
920: \end{equation}
921: Hence, finally we find for the different moments and forces for the bending
922: deformation (fig.~\ref{beameinseitig.fig})
923: %\begin{mathletters}
924: \begin{eqnarray}
925: M_A &=& - \frac{4EI}{L} \Theta_A \label{onesidebendingA}
926: \\
927: M_B &=& - \frac{2EI}{L} \Theta_A \label{onesidebendingB}
928: \end{eqnarray}
929: %\end{mathletters}
930: and with 
931: \begin{equation}
932: M_A = - F_BL - M_B
933:   \label{addmoments}
934: \end{equation}
935: \begin{equation}
936: F_B=-F_A=\frac{6EI}{L^2} \Theta_A~.
937: \label{onesidebendingf}
938: \end{equation}
939: (Note that the moments $M_A$ and $M_B$ in eq.~(\ref{addmoments}) have
940: to be added as if they would apply to the same end of the beam.)
941:  
942: If the beam is bent at both ends by angles $\Theta_A$ and $\Theta_B$
943: (Fig.~\ref{beamdoppelt.fig}) the resulting moments $M_A$ and $M_B$ and
944: forces $F_A$ and $F_B$ can be calculated by a superposition of the
945: forces and moments according to two independent bending deformations
946: of the type discussed in the previous case.
947: 
948: The angle $\Theta_A$ generates the moments $M_A^{*}$ and $M_B^{*}$
949: according to eqs.~(\ref{onesidebendingA}, \ref{onesidebendingB})
950: %\begin{mathletters}
951: \begin{eqnarray}
952: M_A^{*} &=& - \frac{4EI}{L} \Theta_A\\
953: M_B^{*} &=& - \frac{2EI}{L} \Theta_A
954: \end{eqnarray}
955: %\label{twosidebendingm}
956: %\end{mathletters}
957: and the angle $\Theta_B$ causes the moments
958: %\begin{mathletters}
959: \begin{eqnarray}
960: M_A^{**} &=& - \frac{2EI}{L} \Theta_B \\
961: M_B^{**} &=& - \frac{4EI}{L} \Theta_B~.
962: \end{eqnarray}
963: %\end{mathletters}
964: Hence we find the resulting moments
965: %\begin{mathletters}
966: \begin{eqnarray}
967: M_A=M_A^{*}+M_A^{**} &=& - \left( \frac{4EI}{L} \Theta_A +
968:   \frac{2EI}{L} \Theta_B \right)
969: \label{momentsbending.eqA}
970: \\
971: M_B=M_B^{*}+M_B^{**} &=& - \left( \frac{2EI}{L} \Theta_A + \frac{4EI}{L} \Theta_B~\right),
972: \label{momentsbending.eqB}
973: \end{eqnarray}
974: %\end{mathletters}
975: and with $M_A+M_B=- F_BL$ the resulting forces
976: \begin{equation}
977: F_B = -F_A =\frac{6EI}{L^2}\left( \Theta_A + \Theta_B \right)~.
978: \label{forcesbending.eq}
979: \end{equation}
980: 
981: Comparing the results from sections \ref{shearing.subs} and
982: \ref{bending.subs} one can show that each deformation of a beam can be
983: expressed by bending and elongation, i.e.~the shear deformation need
984: not to be considered. The proof is given in Appendix \ref{redundanz}.
985: 
986: The deformation of the beams is damped proportionally to the
987: deformation rates
988: %\begin{mathletters}
989: \begin{eqnarray}
990:   M_A^{(d)} &=& -\frac{\gamma~I}{L} ~\dot{\Theta}_A , \\ 
991:   M_B^{(d)} &=& -\frac{\gamma~I}{L} ~\dot{\Theta}_B , \\ 
992:   F_A^{N(d)} &=& -\gamma\left( \vec{v}_A - \vec{v}_B \right)
993:     \cdot \stackrel{\longrightarrow}{AB} ,\\
994:   F_B^{N(d)} &=& \gamma\left( \vec{v}_A - \vec{v}_B \right)
995:     \cdot \stackrel{\longrightarrow}{AB} ,
996: %  \label{dampingmom}
997: \end{eqnarray}
998: %\end{mathletters}
999: where $\gamma$ is the damping coefficient of the beam material.  As
1000: proven in the appendix the deformation of the beam is in linear
1001: approximation completely determined by the angles $\Theta_A$ and
1002: $\Theta_B$ and the length $L$. Hence there is no damping force acting
1003: in shear direction.
1004: 
1005: \section{Implementation of the algorithm}
1006: \label{implementationSEC}
1007: We implemented a molecular dynamics algorithm using the particle model
1008: described in the previous sections in FORTRAN. A Gear predictor
1009: corrector scheme of fourth
1010: order~\cite{AllenTildesley:1987,Gear:1966ug} was applied to integrate
1011: Newton's equations of motion. We have to proceed both the forces
1012: caused by the elastic deformation of the triangles as well as the
1013: forces induced by the beams for each iteration step
1014: (eq.~\ref{totalforceEQ}). Since every particle can interact with each
1015: other one causing a high algorithmic complexity, we applied a Verlet
1016: neighborhood list~\cite{Verlet:1967ug} to decrease the amount of
1017: computer time. As discussed below (see table~\ref{table}) the
1018: calculation of the triangle intersections is the most time intensive
1019: part of the calculation. Hence we applied an advanced version of the
1020: neighborhood list method to keep the fraction of the computation of
1021: the triangle--interactions as small as possible.
1022: 
1023: Our list method acts in two steps. In the first step we prepare
1024: neighborhood lists for each triangle to reduce the number of possible
1025: interactions which has to checked. Because we have to investigate all
1026: possible triangle interactions the time for this step rises as the
1027: square of the total number of triangles $T\sim \left( N \cdot
1028: n_t\right)^2$. (For simplicity we assumed for this estimate that all
1029: $N$ grains consist of $n_t$ triangles each. Hence there are $\left( N
1030: \cdot n_t\right) \cdot \left( \left( N-1\right) \cdot n_t\right) $
1031: possible triangle--interactions.)  In the second step we check for
1032: each entry of the constructed neighborhood list whether the
1033: corresponding triangles interact indeed or not. If they interact,
1034: i.e.~if there is an overlap area, the type of interaction is
1035: determined according to the classification scheme in
1036: section~\ref{trianglesSEC}.
1037: 
1038: Finally we construct three lists for the interactions of the first
1039: three types of the classification scheme. The interactions of the
1040: forth and fifth type are replaced by interactions of the first type as
1041: described above. The computer time for this step rises linearly with
1042: the number of particles $N$ and with the number of triangles $n_t$
1043: belonging to each particle. Different from usual Verlet tables these
1044: lists contain only triangles which definitely do interact, and hence
1045: we call the most time consuming subroutines for the calculation of the
1046: forces only for triangles which do interact in one of the five manners
1047: described above (section~\ref{trianglesSEC}).
1048: 
1049: The calculation of the forces induced by the beam deformations follows straight
1050: forward the formulas described in section~\ref{stressesinbeamsSEC}.
1051: 
1052: In table~\ref{table} we present a detailed performance analysis of
1053: the algorithm. The numbers in the table refer to simulations on a DEC
1054: 3000/700 workstation. As visible from the data most of the time
1055: (approximately $70 \%$) is necessary to construct the interaction
1056: list, because we check here every pair of neighbored triangles for
1057: intersections. The construction of this lists give the possibility to
1058: classify the interactions and therefore to use for each kind of
1059: interaction its own optimised algorithm. This prevents us from
1060: calculating useless intersection points, areas and other data.
1061: 
1062: The computer time needed for the predictor--corrector integration is
1063: very low ($3.7 \%$). Thus it seems to be not useful to optimise the
1064: integration procedure, or to decrease the accuracy of the integration
1065: to safe computer time. The calculation of the beam forces is not
1066: time expensive as well. 
1067: 
1068: \begin{table}[htbp]
1069:   \caption{The performance analysis of the algorithm. The 
1070:     calculation has been done on a DEC--3000/700 workstation
1071:     (explanation in the text).}
1072:   \begin{tabular}{llll}
1073:     Basic algorithm & portion of & subroutine & portion of\\
1074:     &computation&&computation\\
1075:     \hline
1076:     Construction of the lists & 73.7 \% & Verlet list & 2.07 \% \\
1077:     \cline{3-4}
1078:     && Classifying interaction & 70.62 \% \\ \cline{3-4}
1079:    \hline
1080:     Calculation of the & 7.87 \% & TYPE 1 & 7.54 \% \\ \cline{3-4}
1081:     triangle interaction & & TYPE 2 & 0.24 \% \\ \cline{3-4}
1082:     & & TYPE 3 & 0.09 \% \\
1083:     \hline
1084:     Calculation of the & & & 13.33 \% \\
1085:     beam forces & & &\\
1086:     \hline
1087:     integration & 3.71 \% & predictor & 1.16 \% \\ \cline{3-4}
1088:     && corrector & 2.55 \%\\
1089:     \hline
1090:     all others & & & 1.4 \% 
1091:   \end{tabular}
1092:   \label{table}
1093: \end{table}
1094: 
1095: From the data in table~\ref{table} it is obvious that triangle
1096: interactions of the first type dominate. Approximately $98 \%$ of the
1097: occurring interactions are of this type. Interactions of the fourth and
1098: fifth type, where our program yields approximative results only, occur
1099: extremely seldomly. We found them only a few times during all our
1100: simulations.
1101: 
1102: \section{First results}
1103: \label{resultsSEC}
1104: \subsection{The collision of a two particles}
1105: \label{twoparticles}
1106: To demonstrate the non--trivial behavior of colliding non--spherical
1107: particles and to check the correctness of our implemented model we
1108: want to present the results of an experiment where a quadratic grain
1109: collides with another resting one in the absence of gravity. The grains
1110: consist of four equal triangles each, forming a simple quadratic grain
1111: as shown in fig.~\ref{grainsFIG}.  The parameters of the materials
1112: are: $Y=2\cdot 10^{7}\, g/(cm\, sec^2)$, $E = 1\cdot 10^{5}
1113: \,g/(sec^2)$, $I=1\cdot 10^{-4} cm^{3}$ and $\gamma=9$ $g/sec$.
1114: The value of the Young--module $Y$ agrees with most of the simulations
1115: of spheres which can be found in the literature.
1116: 
1117: Figs.~\ref{fig:twoparticles} (left figures) show the time evolution
1118: of the particles as a sequence of stroboscopic snapshots for different
1119: relative velocities of the particles a)~$v_{rel}=10\,cm/sec$,
1120: b)~$v_{rel}=50\,cm/sec$ and c)~$v_{rel}=100\,cm/sec$.
1121: \begin{figure}[htbp]
1122:   \ifcase \value{nofigs}
1123:     \centerline{\psfig{figure=figs/twopart.10.ps,width=4cm}\hspace{0.5cm}\psfig{figure=figs/e10.ps,width=7cm,angle=270}}
1124:     \centerline{\psfig{figure=figs/twopart.50.ps,width=4cm}\hspace{0.5cm}\psfig{figure=figs/e50.ps,width=7cm,angle=270}}
1125:       \centerline{\psfig{figure=figs/twopart.100.ps,width=4cm}\hspace{0.5cm}\psfig{figure=figs/e100.ps,width=7cm,angle=270}}
1126: \vspace{0.5cm}
1127: \fi
1128:   \caption{The collision of a horizontally from left to right 
1129:     moving grain (black drawn) with a resting one (grey) for different
1130:     impact rates $v_{rel}=10 \,cm/sec$ (top), $v_{rel}=50
1131:     \,cm/sec$ (middle), and $v_{rel}=100 \,cm/sec$
1132:     (bottom). The stroboscopic snapshots of the grains are take at
1133:     equidistant time intervals. An animated
1134:     sequence of the collisions can be accessed via World Wide
1135: %???Web~\cite{www}.
1136: Web~[67].
1137:     The right hand figures display the kinetic, rotational and total
1138:     energies of the particle collisions over time.}
1139:   \label{fig:twoparticles}
1140: \end{figure}
1141: 
1142: 
1143: One of the triangles of each grain has been filled to visualise the
1144: rotation of the grains. Initially the velocity of the left particle is
1145: $\vec{v}=\left({v_{rel},0}\right) \,cm/sec$ and the right
1146: particle rests. For low velocities $v<v_{tc}\approx 70 cm/sec$ the
1147: grains collide twice within a very short time. Therefore the traces of
1148: the particles that undergo soft collisions with velocity values
1149: $v<v_{tc}$ differ qualitatively from the traces of particles colliding
1150: with higher relative velocity. Since it is hard to display the
1151: complicated motion of the grains in detail we refer to our World Wide
1152: Web--site URL
1153: http://summa.physik.hu-berlin.de:80/$\sim$thorsten/MDASGP.html~\cite{www}
1154: where one can find animated sequences of various problems mentioned in
1155: the current paper.  The right hand side in fig.~\ref{fig:twoparticles}
1156: shows the rotational and kinetic energies and the total energy as a
1157: function of time. The initial kinetic energy of the translational
1158: motion of the left particle results in kinetic energy of both
1159: particles due to translational motion and rotation. A part of the
1160: mechanical energy is lost due to dissipation in the beams. The
1161: relative amount of rotational energy depends strongly on the impact
1162: velocity as expected from the discussion above. The dependence of the
1163: extinction of the rotational degree of freedom on the impact rate is
1164: demonstrated in fig.~\ref{rotvel} too. This figure shows the time
1165: evolution of the angular velocity of the grains $\omega =
1166: \frac{1}{n_t}\,\left| \sum\limits_{i=1}^{n_t} \left( \vec{v}_i -
1167:     \vec{v}_g \right) \times \vec{s}_i\right| $ over time for
1168: different values of the impact velocity. $\vec{v}_g = \frac{1}{n_t}
1169: \sum\limits_{i=1}^{n_t} \vec{v}_i$ is the velocity of the grain and
1170: $\vec{s}_i$ is the vector pointing from the center of mass point of
1171: the grain to the center of mass point of the triangle $i$. $n_t$ is
1172: the number of triangles the grain consists of. 
1173: \begin{figure}[htbp]
1174:   \ifcase \value{nofigs}
1175:   \centerline{\psfig{figure=figs/angvel.ps,width=8cm,angle=270}}
1176: \vspace{0.5cm}
1177: \fi
1178:   \caption{The angular velocities of the particles over the
1179:     time. For lower velocities the particles collide two times,
1180:     therefore those collisions results in a small angular velocity of
1181:     the grains.}
1182:   \label{rotvel}
1183: \end{figure}
1184: 
1185: For high velocity one
1186: observes an abrupt change of the rotational motion at the time of the
1187: collision and a short damped oscillation according to the excited
1188: inner degrees of freedom of the grain (distortion of beams) and their
1189: relaxation.  After the collision the particles move on with constant
1190: translational and rotational velocities. For low impact rate we find a
1191: quite different behavior: First the particles collide as in the
1192: previous case, resulting in a certain rotational motion. Short time
1193: after the first collision, however, they collide a second time with
1194: different edges. Finally we find for the latter case of slow
1195: collisions a very small resulting angular velocity because the second
1196: collision causes angular moments in opposite directions, hence the
1197: resulting moment is small.
1198: 
1199: Observing the motion of a moving particle colliding with a resting one
1200: (see~\cite{www}), i.e.~the simplest possible contact of particles, one
1201: can remark that even for this system there is a complex behavior.  For
1202: rigid spheres the velocities after a collision as a function of the
1203: impact velocities~(eq.~(\ref{resitutionEQa}, \ref{resitutionEQb})) can be calculated
1204: analytically solving the viscoelastic equations of the
1205: spheres~\cite{BrilliantovSpahnHertzschPoeschel:1994}. We guess that
1206: this will not be possible for the case of our particles.
1207:       
1208: \subsection{Outflow of a hopper}
1209: \label{sec:hopper}
1210: The outflow of a hopper is of high interest not only because of the
1211: technological importance of this process, but also because of the
1212: exciting phenomena as clogging and density waves which have been
1213: discovered recently (e.g.~\cite{BaxterBehringerFagertJohnson:1989}).
1214: Density waves and clogging have been investigated using molecular
1215: dynamics in two and three dimensions in various papers,
1216: e.g.~\cite{FormKohringMelinPuhlTillemans:1994,LangstonTuezuenHeyes:1993,RistowHerrmann:1995}.
1217: There is at least one other interesting and surprising phenomenon,
1218: detected recently by Evesque and
1219: Meftah~\cite{EvesqueMeftah:1992,EvesqueMeftah:JDP}: They found that a
1220: hour glass ``ticks'' much slower if it is subjected to vertical
1221: vibrations $y=A\cos\left(\omega t\right)$. The effect seems to depend
1222: on the frequency $f=\omega/2\,\pi$ and on the acceleration of the
1223: vibration $\Gamma = A\,\omega^2$. For frequencies between $40\,Hz\le f
1224: \le 60\,Hz$ surprisingly they observed for some accelerations that the
1225: flow almost stops.
1226: 
1227: We investigated this effect using different particle models and we
1228: found that the effect could be reproduced with the new particle model
1229: only. The results will be described in detail in a forthcoming
1230: paper~\cite{BuchholtzPoeschel:1995hour}.
1231: 
1232: Fig.~(\ref{fig:hopper}) shows snapshots of the system with $N=400$
1233: complex particles. For an animated sequence we refer to our
1234: WWW--site~\cite{www}. The left figure shows the outflowing hopper, the
1235: flow varies irregularly and the grey scale codes for the particle
1236: velocities. When animating the figures one clearly observes density
1237: waves~\cite{www}. The right hand figure shows the same system a few
1238: moments later. The flow has stopped due to clogging.
1239: \begin{figure}[htbp]
1240:   \ifcase \value{nofigs}
1241: \centerline{\psfig{figure=figs/start.ps,width=4cm,angle=0,bbllx=170bp,bblly=130bp,bburx=430bp,bbury=415bp,clip=}\psfig{figure=figs/end.ps,width=4cm,angle=0,bbllx=170bp,bblly=130bp,bburx=430bp,bbury=415bp,clip=}}
1242: \fi
1243:   \caption{Snapshots of the simulation of an outflowing 
1244:     hopper. The left figure shows regular outflow, the right figure
1245:     show a clogging situation. An animated sequence is available via
1246:     World Wide 
1247: %???Web~\cite{www}.
1248: Web~[67].
1249: }
1250:   \label{fig:hopper}
1251: \end{figure}
1252: %\clearpage
1253: 
1254: \subsection{Granular flow in a rotating cylinder}
1255: \label{sec:drehtrommel}
1256: The flow of granular material in a rotating cylinder is one of the
1257: most popular problems in the field of granular materials. It has been
1258: investigated experimentally and theoretically by many authors using
1259: various techniques (see
1260: e.g.~\cite{Rajchenbach:1990,Rolf:1993KUG,Nakagawa:1994,ZikEtAl:1994,Ristow:1994,BaumannJanosiWolf:1994,PoeschelBuchholtz:1993}
1261: and many others). The results shall not be discussed here. We applied
1262: the algorithm to this problem too. For the detailed description of the
1263: results see~\cite{BuchholtzPoeschelTillemans:1994}. Here we only want
1264: to demonstrate the abilities of the method and to mention the main
1265: results briefly.
1266: 
1267: Fig.~\ref{snap.fig} shows snapshots of the simulation of a slowly
1268: rotating cylinder. The grey scale codes for the velocity of the
1269: grains, black means high velocity, black codes for high velocity.
1270: For low angular velocity of the driven cylinder one finds
1271: experimentally stick--slip flow (e.g.~\cite{Rajchenbach:1990}),
1272: i.e.~the material moves downwards not homogeneously but it forms
1273: avalanches. In the right figure one observes an avalanche on the top
1274: of the material indexed by light grey shadowed grains. The left figure shows
1275: the same system immediately before the avalanche. Avalanches are
1276: relatively seldom events. Most of the time the particles rest with
1277: respect to each other, i.e.~they move only due to the rotation of the
1278: cylinder (left side of fig.~\ref{snap.fig}). When increasing the
1279: angular velocity of the cylinder there is a relatively sharp
1280: transition between the stick--slip flow and the homogeneous regime.
1281: This transition could be found in the simulation using non--spherical
1282: grains~\cite{BuchholtzPoeschelTillemans:1994}. Animated sequences for
1283: both regimes, stick--slip and continuous flow, can be accessed via
1284: World Wide Web~\cite{www}.
1285: \begin{figure}[ht]
1286:   \ifcase \value{nofigs}
1287: \centerline{\psfig{figure=figs/snap.ps,width=8cm,angle=0}}
1288: \fi
1289: \caption{\em Snapshots of the simulation with $N=500$ complex
1290:   particles with $P\in [0.1\,cm; 0.2\,cm]$ in a rotating cylinder of
1291:   diameter $D=4\,cm$. The angular velocity is $\Omega=0.1\,sec^{-1}$.
1292:   The wall particles have not been drawn. The right snapshot shows an
1293:   avalanche, the left one has been taken short time before. The grey
1294:   scale codes for the particle velocity.}
1295: \label{snap.fig}
1296: \end{figure}
1297: 
1298: 
1299: Fig.~\ref{inclination.fig} shows the dependence of the difference
1300: between the inclination of the material and the angle of repose on the
1301: angular velocity of the cylinder. The line displays the behavior found
1302: by Rajchenbach $\Theta-\Theta_c \sim
1303: \Omega^2$~\cite{Rajchenbach:1990}. The simulation data are in good
1304: agreement with this observation. 
1305: \begin{figure}[ht]
1306:   \ifcase \value{nofigs}
1307: \centerline{\psfig{figure=figs/thetaomega.ps,width=10cm,angle=270}}
1308: \vspace{0.5cm}
1309: \fi
1310: \caption{\em The inclination $\Theta$ of the material surface over 
1311:   the angular velocity $\Omega$. The dotted line displays the function
1312:   which has been measured experimentally by
1313: %???  \nobreak{Rajchenbach}~\cite{Rajchenbach:1990}.
1314:   \nobreak{Rajchenbach}~[73].
1315: }
1316: \label{inclination.fig}
1317: \end{figure}
1318: 
1319: Both results mentioned, the sharp transition between the flow regimes
1320: and the inclination as a function of the angular velocity of the
1321: cylinder, agree with experimental observations. We have not been
1322: able to find these effects using simple spheres or using particles
1323: composed of
1324: spheres~\cite{PoeschelBuchholtz:1993,PoeschelBuchholtz:1993CSF} in
1325: molecular dynamics simulations.
1326: 
1327: \section{Conclusion}
1328: \label{ConclusionSEC}
1329: We presented a model for the simulation of the particles of a granular
1330: material. Each particle consists of triangles which are connected by
1331: deformable beams. There are no restrictions concerning the shape of the
1332: grains (convex or concave), nor the number or the shape of the
1333: triangles a grain consists of. To preserve calculation time it is
1334: favourable to chose arrangements of the triangles so that the beams
1335: are fixed at the center of mass points of the triangles. Otherwise one needs
1336: additional computation time caused by the Steiner law. 
1337: 
1338: When two triangles of different grains collide there acts an elastic
1339: restoring force proportionally to the compression of the particles.
1340: During the collision of the triangles the energy is preserved. A
1341: collision of a triangle with another one causes moments and forces
1342: acting on this triangle and hence a deformation of the beams which
1343: connect the triangles to neighboring ones. The deformation of the
1344: beams causes forces and moments acting on the neighboring triangles.
1345: 
1346: Beams can be deformed in axial direction (elongation) and in shear
1347: direction and they can be bent. When a beam is deformed it dissipates
1348: energy proportionally to the deformation rate.
1349: 
1350: In the present paper we described beams that recover completely while
1351: dissipating mechanical energy when the deforming moments and forces
1352: vanish. Our model, however, is not restricted to this case.
1353: Interesting phenomena are plastic deformation and wear which can be
1354: easily simulated by introducing thresholds for the beam forces and
1355: moments or for the deformations. When a force or a moment or a
1356: deformation, respectively, exceeds the threshold the beam breaks or
1357: deforms permanently. Similar simulations with other models have been
1358: reported in the
1359: literature~\cite{Tillemans:1994,Handley:1993,PotapovCampbellHopkins:1994aug}.
1360: 
1361: The proposed algorithm has been implemented in FORTRAN. We have
1362: demonstrated the behavior of a system of grains applying it to three
1363: examples of granular assemblies. In a simple collision simulation of
1364: two grains we discussed the trajectories of the grains as well as the
1365: evolution of the kinetic and rotational energies depending on the
1366: impact rate. Sample results for the flow out of a hopper and the
1367: motion of granular material in a rotating cylinder have been
1368: presented. In the latter case we found quantitative agreement of the
1369: simulation with the experiment that could not be found so far, neither
1370: with spherical grains nor with non--spherical grains of other type.
1371: 
1372: \ack We thank Hans J.~Herrmann, Stefan Schwarzer and
1373: Hans--J\"urgen Tillemans for many helpful discussions and comments.
1374: This work has been supported by the Deutsche Forschungsgemeinschaft
1375: (grant Ro 548/5-1).
1376: 
1377: \begin{appendix}
1378: \section{Shear deformation can be expressed by bending}
1379: \label{redundanz}
1380: The deformation of a beam in the two dimensional space has three
1381: degrees of freedom: the distance of the relative positions of the ends
1382: of the beams $L=\sqrt{\left( x_A-x_B\right) ^2+\left( y_A-y_B\right)
1383:   ^2}$, and the angles $\Theta_A$ and $\Theta_B$. These values are the
1384: parameters in eqs.~(\ref{momentsbending.eqA}, \ref{momentsbending.eqB}) and
1385: (\ref{forcesbending.eq}). The shear parameter $\Delta$ in
1386: eqs.~(\ref{forcesshearing.eq}, \ref{momshearing.eq})
1387: %\ref{eq.shearing}) 
1388: is a fourth
1389: parameter, i.e.~one of them must be redundant. In the following we
1390: will show that shear deformation can be expressed by bending, at least
1391: in our approach of a linear deformation--force relation.
1392: 
1393: \begin{figure}[ht]
1394:   \ifcase \value{nofigs}
1395: \centerline{\psfig{figure=/figs/shearbend.eps,width=8cm,angle=270}}
1396: \vspace{0.5cm}
1397: \fi
1398: \caption{The deformation of a beam in shear direction $\Delta$ (upper 
1399: part) can be represented by bending deformation (lower part). The 
1400: appendix contains the proof that one need not to consider shear 
1401: {\em and} bending deformation in the linear approximation which 
1402: is justified for small deformations.}
1403: \label{shearbend.fig}
1404: \end{figure}
1405: Suppose we have a deformed beam as drawn in Fig.~\ref{shearbend.fig}
1406: (bottom). Then applying eqs.~(\ref{forcesshearing.eq}, \ref{momshearing.eq})
1407: %\ref{eq.shearing}) 
1408: we find for the
1409: forces and moments according to shearing
1410: %\begin{mathletters}
1411: \begin{eqnarray}
1412: M_A^s &=& M_B^s = \frac{6EI}{\left(L^*\right)^2}\, \Delta\\
1413: F_A^s &=& -F_B^s = \frac{12EI}{\left(L^*\right)^3}\, \Delta
1414: %\label{Shearing*}
1415: \end{eqnarray}
1416: %\end{mathletters}
1417: and according to bending (eqs.~(\ref{momentsbending.eqA},\ref{momentsbending.eqB}) and
1418: (\ref{forcesbending.eq}))
1419: %\begin{mathletters}
1420: \begin{eqnarray}
1421: M_A^b &=& -\frac{2EI}{L^*}\,\Theta_B^*\\
1422: M_B^b &=& -\frac{4EI}{L^*}\,\Theta_B^*\\
1423: F_A^b &=& -F_B^b = -\frac{6EI}{\left(L^*\right)^2}\,\Theta_B^*~.
1424: %\label{Bending*}
1425: \end{eqnarray}
1426: %\end{mathletters}
1427: The total forces and moments read
1428: %\begin{mathletters}
1429: \begin{eqnarray}
1430: M_A^*&=& M_A^s + M_A^b = \frac{6EI}{\left(L^*\right)^2}\,\Delta - \frac{2EI}{L^*}\,\Theta_B^* 
1431: \label{Total*A}
1432: \\
1433: M_B^*&=& M_B^s + M_B^b = \frac{6EI}{\left(L^*\right)^2}\,\Delta - \frac{4EI}{L^*}\,\Theta_B^*
1434: \label{Total*B}
1435:  \\
1436: F_A^*&=&-F_B^* = \frac{12EI}{\left(L^*\right)^3}\, \Delta - \frac{6EI}{\left(L^*\right)^2}\,\Theta_B^*~.
1437: \label{Total*C}
1438: \end{eqnarray}
1439: %\end{mathletters}
1440: Turning the entire system by the angle $\Theta_A\approx\tan \Theta_A =
1441: -\frac{\Delta}{L^*}$ we obtain the new system as drawn in the top of
1442: fig.~\ref{shearbend.fig} with the new angles
1443: $\Theta_A=-\frac{\Delta}{L}$ and $\Theta_B=\Theta_B^*-\frac{\Delta}{L}$.
1444: With eqs.~(\ref{momentsbending.eqA}, \ref{momentsbending.eqB}) and (\ref{forcesbending.eq}) we find
1445: %\begin{mathletters}
1446: \begin{eqnarray}
1447: M_A &=& -\frac{4EI}{L}\,\Theta_A- \frac{2EI}{L}\,\Theta_B = \frac{6EI}{L^2}\,\Delta - \frac{2EI}{L}\,\Theta_B^*
1448: \label{TotalA}
1449: \\
1450: M_B &=& -\frac{2EI}{L}\,\Theta_A- \frac{4EI}{L}\,\Theta_B = \frac{6EI}{L^2}\,\Delta - \frac{4EI}{L}\,\Theta_B^*
1451: \label{TotalB}
1452: \\
1453: F_A &=& -F_B = -\frac{6EI}{L^2}\,\left(\Theta_A+\Theta_B \right) = \frac{12EI}{L^3}\,\Delta - \frac{6EI}{L^2}\,\Theta_B^*
1454: \label{TotalC}
1455: \end{eqnarray}
1456: %\end{mathletters}
1457: Approximating $L$ by $\sqrt{L^2+\Delta^2} = L^*$ evidently the results
1458: in eqs.~(\ref{TotalA}-\ref{TotalC}) coincide with eqs.~(\ref{Total*A}-\ref{Total*C}). Thus we do
1459: not need to consider shear deformation in our simulations since it is
1460: redundant here.
1461: \end{appendix}
1462: 
1463: \begin{thebibliography}{10}
1464: 
1465: \bibitem{AlderWainwright:1957ug}
1466: B.~J.~Alder and T.~E.~Wainwright, 
1467: {\em J. Chem. Phys.}, {\bf 27} 1208 (1957).
1468: 
1469: \bibitem{BeazleyLomdahl:1993}
1470: D.~M. Beazley and P.~S. Lomdahl,
1471: {\em Parallel Computing}, {\bf 20}, 173 (1993).
1472: 
1473: \bibitem{BeazleyLomdahlJensenTamayo:1994}
1474: D.~M. Beazley, P.~S. Lomdahl, N.~Gr{\o}nbech--Jensen, and P.~Tamayo,
1475: In {\em Proceedings 8. Intern. Parallel Processing Symposium,
1476:   Canc{\'u}n, Mexico}, pages 800--809, Los Alamos, 1994. IEEE Computer Society.
1477: 
1478: \bibitem{BeazleyEtAl:1995}
1479: D.~M. Beazley, P.~S. Lomdahl, N.~Gr{\o}nbech--Jensen, R.~Giles, 
1480: and P.~Tamayo, In D.~Stauffer, {\em Annual Reviews of Computational Physics}, 
1481: Vol.~III, World Scientific, (Singapore, 1995) (in press).
1482: 
1483: \bibitem{Hoover:1992}
1484: W.~G. Hoover,
1485: {\em Nuclear Physics A}, {\bf 545}, 523c (1992).
1486: 
1487: \bibitem{Salo:1992}
1488: H.~Salo,
1489: {\em Icarus}, {\bf 96}, 85 (1992).
1490: 
1491: \bibitem{Weidenschilling}
1492: S.~J.~Weidenschilling, C.~R.~Chapman, D.~R.~Davis and R.~Greenberg,
1493: In: R.~Greenberg and A.~Brahic (Eds.),{\em Planetary Rings} pages
1494: 367--415, Arizona, 1994.
1495: 
1496: \bibitem{BrilliantovSpahnHertzschPoeschel:1994}
1497: N.~V. Brilliantov, F.~Spahn, J.--M. Hertzsch, and T.~P\"oschel,
1498: {\em preprint HLRZ 17/94} (1994).
1499: 
1500: \bibitem{GoldhirschZanetti:1993}
1501: I.~Goldhirsch and G.~Zanetti,
1502: {\em Phys. Rev. Lett.}, {\bf 70}, 1619 (1993).
1503: 
1504: \bibitem{McNamaraYoung:1992}
1505: S.~McNamara and W.~R. Young,
1506: {\em Phys. Fluids A}, {\bf 4}, 496 (1992).
1507: 
1508: \bibitem{LudingClementBlumenRajchenbachDuran:1994DISS}
1509: S.~Luding, E.~Cl\'ement, A.~Blumen, J.~Rajchenbach, and J.~Duran,
1510: {\em Phys. Rev. E.}, {\bf 50}, 4113 (1994).
1511: 
1512: \bibitem{LudingClementBlumenRajchenbachDuran:1994STUDIES}
1513: S.~Luding, E.~Cl\'ement, A.~Blumen, J.~Rajchenbach, and J.~Duran,
1514: {\em Phys. Rev. E.}, {\bf 49}, 1634 (1994).
1515: 
1516: \bibitem{Rapaport:1980}
1517: D.~C. Rapaport,
1518: {\em J. Comp. Phys.}, {\bf 34}, 184 (1980).
1519: 
1520: \bibitem{Lubachevsky:1991}
1521: B.~D. Lubachevsky,
1522: {\em J. Comp. Phys.}, {\bf 94}, 255 (1991).
1523: 
1524: \bibitem{CundallStrack:1979}
1525: P.~A. Cundall and O.~D.~L. Strack,
1526: {\em G\'eotechnique}, {\bf 29}, 47 (1979).
1527: 
1528: \bibitem{HaffWerner:1986}
1529: P.~K. Haff and B.~T. Werner,
1530: {\em Powder Techn.}, {\bf 48}, 239 (1986).
1531: 
1532: \bibitem{Hertz:1882ug}
1533: H.~Hertz,
1534: {\em J. f. reine u. angewandte Math.}, {\bf 92}, 156 (1882).
1535: 
1536: \bibitem{GallasHerrmannSokolowski:1992PRL}
1537: J.~A.~C. Gallas, H.~J. Herrmann, and S.~Soko{\l}owski,
1538: {\em Phys. Rev. Lett.}, {\bf 69}, 1371 (1992).
1539: 
1540: \bibitem{Taguchi:1992PRL}
1541: Y--h. Taguchi
1542: {\em Phys. Rev. Lett.}, {\bf 69}, 1367 (1992).
1543: 
1544: \bibitem{GallasHerrmannPoeschelSokolowski:1994}
1545: J.~A.~C. Gallas, H.~J. Herrmann, T.~P\"oschel, and S.~Soko{\l}owski,
1546: {\em J. Stat. Phys., in press} (1995).
1547: 
1548: \bibitem{PoeschelHerrmann:1994}
1549: T.~P\"oschel and H.~J. Herrmann,
1550: {\em Europhys. Lett.}, {\bf 29}, 123 (1995).
1551: 
1552: \bibitem{RistowHerrmann:1994}
1553: G.~H. Ristow and H.~J. Herrmann,
1554: {\em Phys. Rev. E.}, {\bf 50}, R5 (1994).
1555: 
1556: \bibitem{FormKohringMelinPuhlTillemans:1994}
1557: W.~Form, G.~A. Kohring, S.~Melin, H.~Puhl, and H.--J. Tillemans,
1558: {\em preprint HLRZ 75/93} (1994).
1559: 
1560: \bibitem{CampbellPotapov:1993}
1561: C.~Campbell and A.~V. Potapov,
1562: {\em in \cite{IinoyaTsujiMasudaHigashitani:1993}}, pages 27--31, (1993).
1563: 
1564: \bibitem{Poeschel:1994}
1565: T.~P\"{o}schel,
1566: {\em J. Physique}, {\bf 4}, 499 (1994).
1567: 
1568: \bibitem{Lee:1994}
1569: J.~Lee,
1570: {\em Phys. Rev. E.}, {\bf 49}, 281 (1994).
1571: 
1572: \bibitem{Ristow:1994}
1573: G.~H. Ristow,
1574: {\em Europhys. Lett.}, {\bf 28}, 97 (1994).
1575: 
1576: \bibitem{RistowCantelaubeBideau:1994ug}
1577: G.~H. Ristow, F.~Cantelaube--Lebec, and D.~Bideau,
1578: {\em preprint} (1994).
1579: 
1580: \bibitem{PoeschelBuchholtz:1993CSF}
1581: T.~P\"{o}schel and V.~Buchholtz,
1582: {\em Chaos, Solitons \& Fractals, in press} (1993).
1583: 
1584: \bibitem{Walton:1991}
1585: O.~R. Walton,
1586: In {\em Second US--Japan Seminar on Micromechanics of Granular
1587:   Materials}, pages 1--9, Potsdam, NY (1991).
1588: 
1589: \bibitem{Poeschel:1993JDP}
1590: T.~P\"oschel,
1591: {\em J. Physique}, {\bf 3}, 27 (1993).
1592: 
1593: \bibitem{Melin:1994}
1594: S.~Melin,
1595: {\em Phys. Rev. E.}, {\bf 49}, 2353 (1994).
1596: 
1597: \bibitem{Taguchi:1992JDP}
1598: Y-h. Taguchi,
1599: {\em J. Physique}, {\bf 2}, 2103 (1992).
1600: 
1601: \bibitem{Taguchi:1994}
1602: Y-h. Taguchi,
1603: {\em preprint TITCMT-94-4} (1994).
1604: 
1605: \bibitem{BuchholtzPoeschel:1993JMPC}
1606: V.~Buchholtz and T.~P\"{o}schel,
1607: {\em J. Mod. Phys. C}, {\bf 4}, 1049 (1993).
1608: 
1609: \bibitem{FormItoKohring:1993}
1610: W.~Form, N.~Ito, and G.~A. Kohring,
1611: {\em J. Mod. Phys. C}, {\bf 4}, 1085 (1993).
1612: 
1613: \bibitem{EveraersKremer:1993}
1614: R.~Everaers, and K.~Kremer,
1615: {\em preprint} (1993).
1616: 
1617: \bibitem{BuchholtzPoeschel:1994PA}
1618: V.~Buchholtz and T.~P\"{o}schel,
1619: {\em Physica A}, {\bf 202}, 390 (1994).
1620: 
1621: \bibitem{Lee:1993}
1622: J.~Lee,
1623: {\em J. Physique I}, {\bf 3}, 2017 (1993).
1624: 
1625: \bibitem{VisscherBolsterli:1972ug}
1626: W.~M. Visscher and M.~Bolsterli,
1627: {\em Nature}, {\bf 239}, 504 (1972).
1628: 
1629: \bibitem{JullienMeakinPawlovitch:1992}
1630: R.~Jullien, P.~Meakin, and A.~Pawlovitch,
1631: {\em Phys. Rev. Lett.}, {\bf 69}, 640 (1992).
1632: 
1633: \bibitem{BaumannJobsWolf:1993}
1634: G.~Baumann, E.~Jobs, and D.~E. Wolf,
1635: In M.~Matsuhita, M.~Shlesinger, and T.~Vicsek, editors, {\em Fractals
1636:   in Natural Sciences, Budapest 1993}, World Scientific (Singapore, 1993).
1637: 
1638: \bibitem{BaumannJanosiWolf:1994}
1639: G.~Baumann, I.~J\'anosi, and D.~E. Wolf,
1640: {\em Europhys. Lett.}, {\bf 27}, 203 (1994).
1641: 
1642: \bibitem{JullienMeakinPawlovitch:1993ug}
1643: R.~Jullien, P.~Meakin, and A.~Pawlovitch,
1644: {\em Europhys. Lett.}, {\bf 22}, 523 (1993).
1645: 
1646: \bibitem{BarkerMehta:1995} 
1647: G.~C. Barker and A.~Mehta, 
1648: Comment on \cite{JullienMeakinPawlovitch:1993ug},  
1649: {\em Europhys. Lett.}, {\bf 29}, 61 (1995); 
1650: R.~Jullien, P.~Meakin, and A.~Pavlovitch, 
1651: Reply to \cite{BarkerMehta:1995}, 
1652: {\em Europhys. Lett.}, {\bf 29}, 63 (1995).
1653: 
1654: \bibitem{PengHerrmann:1994}
1655: G.~Peng and H.~J.~Herrmann,
1656: {\em Phys. Rev. E.}, {\bf 49}, 1796 (1994).
1657: 
1658: \bibitem{CaramHong:1991}
1659: H.~S. Caram and D.~C. Hong,
1660: {\em Phys. Rev. Lett.}, {\bf 67}, 828 (1991).
1661: 
1662: \bibitem{GallasSokolowski:1993}
1663: J.~A.~C. Gallas, and S.~Soko{\l}owski,
1664: {\em J. Mod. Phys. B}, {\bf 7}, 2037 (1993).
1665: 
1666: \bibitem{WaltonBraun:1993}
1667: O.~R. Walton and R.~L. Braun,
1668: {\em Joint DOE/NSF Workshop on flow of particulates and fluids}
1669:   (1993).
1670: 
1671: \bibitem{PoeschelBuchholtz:1993}
1672: T.~P\"{o}schel and V.~Buchholtz,
1673: {\em Phys. Rev. Lett.}, {\bf 71}, 3963 (1993).
1674: 
1675: \bibitem{MustoeDePorter:1993}
1676: G.~G. Mustoe and G.~DePorter,
1677: In C.~Thornton, editor, {\em Powders and Grains'93}, pages
1678:   421--427, Balkema (Rotterdam, 1993).
1679: 
1680: \bibitem{HogueNewland:1993}
1681: C.~Hogue and D.~E. Newland,
1682: In C.~Thornton, editor, {\em Powders and Grains'93}, pages
1683:   413--417, Balkema (Rotterdam, 1993).
1684: 
1685: \bibitem{Stronge:1990}
1686: W.~Stronge,
1687: {\em Proc. Roy. Soc.}, {\bf A431} (1990).
1688: 
1689: \bibitem{TillemansHerrmann:1994}
1690: H.~J. Tillemans and H.~J. Herrmann.
1691: {\em Physica A}, {\bf 217}, 261 (1994).
1692: 
1693: \bibitem{Tillemans:1994}
1694: H.~J. Tillemans,
1695: {\em Molekulardynamik--Simulationen beliebig geformter Teilchen in
1696:   zwei Dimensionen},
1697: PhD thesis, Universit{\"a}t K{\"o}ln (1994).
1698: 
1699: \bibitem{Handley:1993}
1700: M.~F. Handley,
1701: {\em An Investigation onto the Constitutive Behaviour of Brittle
1702:   Granular Media by Numerical Experiment},
1703: PhD thesis, University of Minnesota, Minneapolis (1993).
1704: 
1705: \bibitem{PotapovCampbellHopkins:1994aug}
1706: A.~V. Potapov, C.~S. Campbell, and M.~A. Hopkins,
1707: {\em preprint} (1994).
1708: 
1709: \bibitem{PotapovCampbellHopkins:1994bug}
1710: A.~V. Potapov, C.~S. Campbell, and M.~A. Hopkins,
1711: {\em preprint} (1994).
1712: 
1713: \bibitem{Finney:1979}
1714: J.~L. Finney,
1715: {\em J. Comp. Phys.}, {\bf 32}, 137 (1979).
1716: 
1717: \bibitem{HerrmannHansenRoux:1989}
1718: H.~J. Herrmann, A. Hansen, and S. Roux,
1719: {\em Phys. Rev. B.}, {\bf 39}, 637 (1989).
1720: 
1721: \bibitem{BentleyOttmann:1979ug}
1722: J.~L. Bentley, and T.~Ottmann.
1723: {\em IEEE Trans. Comp. C}, {\bf 28}, 9 (1979).
1724: 
1725: \bibitem{PreparataShamos:1985ug}
1726: F.~P. Preparata and M.~I. Shamos,
1727: {\em Computational Geometry: An Introduction},
1728: Springer--Verlag (Berlin, 1985).
1729: 
1730: \bibitem{TimoshenkoGere:1972}
1731: S.~P. Timoshenko and J.~M. Gere
1732: {\em Mechanics of Materials},
1733: van Nordstrand Reinhold Company (New York, 1972).
1734: 
1735: \bibitem{TimoshenkoLessells:1928}
1736: S.~P. Timoshenko and I.~M. Lessells,
1737: {\em Festigkeitslehre},
1738: Julius Springer (Berlin, 1928).
1739: 
1740: \bibitem{AllenTildesley:1987}
1741: M.~P. Allen and D.~J. Tildesley,
1742: {\em Computer Simulations of Liquids},
1743: Clarendon Press (Oxford, 1987).
1744: 
1745: \bibitem{Gear:1966ug}
1746: C.~W. Gear,
1747: {\em The numerical integration of ordinary differential equations of
1748:   various orders},
1749: Technical Report ANL 7126, Argonne National Laboratory (1966).
1750: 
1751: \bibitem{Verlet:1967ug}
1752: L.~Verlet,
1753: {\em Phys. Rev.}, {\bf 159}, 98 (1967).
1754: 
1755: \bibitem{www} %[67]
1756:   The animated sequences mentioned in this paper can be found via
1757:   world wide web under URL
1758:   http://summa.physik.hu-berlin.de:80/$\sim$thorsten/MDASGP.html~. The
1759:   following 9 ``movies'' can be accessed: a) the collision of two
1760:   dice, one resting, the other one moving with 5 different velocities,
1761:   b) the flow through a hopper, the clogging process, c) the flow of
1762:   granular material in a rotating cylinder in the stick--slip regime,
1763:   and in the continuous regime (1995).
1764: 
1765: \bibitem{BaxterBehringerFagertJohnson:1989}
1766: G.~W. Baxter, R.~P. Behringer, T. Fagert, and G.~A. Johnson,
1767: {\em Phys. Rev. Lett.}, {\bf 62}, 2825 (1989).
1768: 
1769: \bibitem{LangstonTuezuenHeyes:1993}
1770: P.~A. Langston, U.~T\"uz\"un, and D.~M. Heyes,
1771: {\em in \cite{IinoyaTsujiMasudaHigashitani:1993}}, pages 43--48 (1993).
1772: 
1773: \bibitem{RistowHerrmann:1995}
1774: G.~Ristow and H.~J. Herrmann,
1775: {\em Physica A}, {\bf 213}, 474 (1995).
1776: 
1777: \bibitem{EvesqueMeftah:1992}
1778: P.~Evesque and W.~Meftah,
1779: {\em C. R. Acad. Sci. Paris II}, {\bf 314}, 1125 (1992).
1780: 
1781: \bibitem{EvesqueMeftah:JDP}
1782: P.~Evesque and W.~Meftah,
1783: {\em Int.Journ. of Mod. Phys. B}, {\bf 17}, 1799 (1993).
1784: 
1785: \bibitem{BuchholtzPoeschel:1995hour}
1786: V.~Buchholtz and T.~P\"{o}schel,
1787: {\em in preparation} (1995).
1788: 
1789: \bibitem{Rajchenbach:1990}
1790: J.~Rajchenbach,
1791: {\em Phys. Rev. Lett.}, {\bf 65}, 2221 (1990).
1792: 
1793: \bibitem{Rolf:1993KUG}
1794: L.~Rolf,
1795: {\em Zement--Kalk--Gips}, {\bf 46}, 117 (1993).
1796: 
1797: \bibitem{Nakagawa:1994}
1798: M.~Nakagawa,
1799: {\em Chemical Engineering Science}, {\bf 49}, 2540 (1994).
1800: 
1801: \bibitem{ZikEtAl:1994}
1802: O.~Zik, D.~Levine, S.~G. Lipson, S.~Shtrikman, and J.~Stavans,
1803: {\em Phys. Rev. Lett.}, {\bf 73}, 644 (1994).
1804: 
1805: \bibitem{BuchholtzPoeschelTillemans:1994}
1806: V.~Buchholtz, T.~P\"oschel, and H.~J. Tillemans,
1807: {\em Physica A}, {\bf 216}, 199 (1995).
1808: 
1809: \bibitem{IinoyaTsujiMasudaHigashitani:1993}
1810: K.~Iinoya, Y.~Tsuji, H.~Masuda, and K.~Higashitani, editors.
1811: {\em First Nisshin Engineering Particle Technology},
1812: Nisshin Engineering Co. Ltd. (Osaka, 1993).
1813: 
1814: \end{thebibliography}
1815: 
1816: \end{document}
1817: 
1818: