+ All Categories
Home > Documents > Experimentally validated computations of flow,...

Experimentally validated computations of flow,...

Date post: 15-Jul-2020
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
21
Experimentally validated computations of flow, mixing and segregation of non-cohesive grains in 3D tumbling blenders Maher Moakher, Troy Shinbrot & Fernando J. Muzzio* Dept. of Chemical and Biochemical Engineering Rutgers University, Piscataway, NJ 08854 *[email protected] Granular mixing is a vital operation in food, chemical, and pharmaceutical industries. Although the tumbling blender is by far the most common device used to mix grains, surprisingly little is known about mixing or segregation in these devices. In this paper, we report the first fully three-dimensional particle dynamics simulations of granular dynamics in two standard industrial tumbling blender geometries: the double-cone and the V-blender. Simulations for both monodisperse and bidisperse (segregating) grain sizes are performed and compared with experiment. Mixing and transport patterns are studied, and we find in both tumblers that the dominant mixing mechanism, azimuthal convection, contends against the dominant bottleneck, axial dispersion. The dynamics of blending, on the other hand, differs dramatically between the two tumblers: flow in the double-cone is nearly continuous and steady, while flow in the V-blender is intermittent and consists of two very distinct processes. Keywords : Granular, Mixing, Simulation, Tumblers 1. Introduction The importance of granular mixing to the US economy is immense. Goods ranging from pharmaceuticals and polymers to semiconductors and ceramics increasingly depend on highly reliable granular flow and highly uniform granular mixing. Specific examples are easily found. The energy reserves of coal in the United States exceed the combined oil reserves in the rest of the world, yet due to processing limitations, in excess of 2 billion tons of coal fines are currently stored as hazardous wastes in the continental US [Killmeyer, 1993]. Likewise, the annual cost of inefficient industrial mixing in the US has been estimated to be as high as $10 billion, with roughly 60% of blended products typically being powdered or granular [Nienow, 1992]. The most common batch mixers in industrial use are tumbling blenders, where grains are caused to flow by a combination of the action of gravity and the rotating motion of the mixer. Two of the most common industrial tumbling blenders are the double-cone and the V-blender (depicted in Fig. 1), which together comprise a high fraction of all batch blending applications. For a review on solid mixing devices, see [Fan 1972; Fan 1990; Carley-Macauley 1962, 1964] and references therein. The flow pattern within these mixers is believed to consist of a thin, rapid flow region near the surface, a nearly non-deforming region beneath that rotates with the container as a solid body, and a narrow transition region between, which is characterized by high shear and density gradients. This picture, however, is best understood in the context of quasi-two- dimensional disk or drum blenders [Hogg 1972; Khakhar 1997]. The analytic foundation for the understanding of flow patterns in fully three-dimensional tumblers such as the double-cone or V-blender is much weaker, and indeed the work that we discuss here - 1 -
Transcript
Page 1: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Experimentally validated computations of flow, mixingand segregation of non-cohesive grains in 3D tumbling

blenders

Maher Moakher, Troy Shinbrot & Fernando J. Muzzio*Dept. of Chemical and Biochemical Engineering

Rutgers University, Piscataway, NJ 08854*[email protected]

Granular mixing is a vital operation in food, chemical, and pharmaceutical industries.Although the tumbling blender is by far the most common device used to mix grains,surprisingly little is known about mixing or segregation in these devices. In this paper, wereport the first fully three-dimensional particle dynamics simulations of granular dynamicsin two standard industrial tumbling blender geometries: the double-cone and the V-blender.Simulations for both monodisperse and bidisperse (segregating) grain sizes are performedand compared with experiment. Mixing and transport patterns are studied, and we find inboth tumblers that the dominant mixing mechanism, azimuthal convection, contends againstthe dominant bottleneck, axial dispersion. The dynamics of blending, on the other hand,differs dramatically between the two tumblers: flow in the double-cone is nearly continuousand steady, while flow in the V-blender is intermittent and consists of two very distinctprocesses.

Keywords: Granular, Mixing, Simulation, Tumblers

1. Introduction

The importance of granular mixing to the US economy is immense. Goods ranging frompharmaceuticals and polymers to semiconductors and ceramics increasingly depend onhighly reliable granular flow and highly uniform granular mixing. Specific examples areeasily found. The energy reserves of coal in the United States exceed the combined oilreserves in the rest of the world, yet due to processing limitations, in excess of 2 billiontons of coal fines are currently stored as hazardous wastes in the continental US[Killmeyer, 1993]. Likewise, the annual cost of inefficient industrial mixing in the US hasbeen estimated to be as high as $10 billion, with roughly 60% of blended products typicallybeing powdered or granular [Nienow, 1992].

The most common batch mixers in industrial use are tumbling blenders, where grains arecaused to flow by a combination of the action of gravity and the rotating motion of themixer. Two of the most common industrial tumbling blenders are the double-cone and theV-blender (depicted in Fig. 1), which together comprise a high fraction of all batchblending applications. For a review on solid mixing devices, see [Fan 1972; Fan 1990;Carley-Macauley 1962, 1964] and references therein.

The flow pattern within these mixers is believed to consist of a thin, rapid flow region nearthe surface, a nearly non-deforming region beneath that rotates with the container as a solidbody, and a narrow transition region between, which is characterized by high shear anddensity gradients. This picture, however, is best understood in the context of quasi-two-dimensional disk or drum blenders [Hogg 1972; Khakhar 1997]. The analytic foundationfor the understanding of flow patterns in fully three-dimensional tumblers such as thedouble-cone or V-blender is much weaker, and indeed the work that we discuss here

- 1 -

Page 2: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

indicates that this quasi-2D picture does not carry over well into truly three-dimensionaltumblers. To date, the design and control of realistic, three-dimensional blenders has beenbased more on trial and error than on quantitative or analytic methods. Even quantitativecharacterizations of mixing performance as a function of the most basic parameters, such asvessel speed or filling level, are scarce in the literature [Brone 1998a; Brone 1997a; Muzzio1997; Brone 1998b].

In this paper, we apply particle dynamics techniques to study flow and mixing in double-cones and V-blenders. While this approach is currently limited in particle numbers toO(104), it is deemed adequate for freely flowing materials, which are commonly present inindustries ranging from fertilizers and foods to alloys and plastics, and which often presentprofound mixing challenges due to strong segregational tendencies. Cohesive materials,which exhibit different phenomenology, usually require high shear to mix homogeneouslyand are often processed in other types of equipment.

The plan for this paper is as follows. A brief review of granular flow models is presentedin §2, and the computational method used is summarized in §3. Numerical results arepresented and compared with experiment in §4 for the double-cone tumbler and in §5 forthe V-blender. Finally, brief conclusions are presented in §6.

2. Granular flow modeling

Models of granular flows can be broadly divided into three categories: continuum, kinetictheory, and discrete.

Continuum models neglect the discrete nature of grains and assume a continuous variationof matter that obeys conservation laws of mass and momentum. The behavior of thematerial is thus assumed to be described by constitutive equations, frequently ad hoc, thatrelate kinematic, mechanical, and thermal field variables. Examples of continuum modelsare those adapted from soil plasticity [e.g. Spencer 1982] and fluid mechanics [e.g.Goodman 1972]. Some continuum models incorporate microstructural parameters thathave the ability to describe characteristics and mechanisms peculiar to granular materials,such as the solid fraction distribution function and dilatancy phenomena [e.g. Mehta 1990].While many of these models are limited to high speed flows where either the Coulombicfriction or collisional interactions dominate, models that incorporate both types of behaviorhave also been proposed [e.g. Sayed 1981; Johnson 1987; McTigue 1987].

Kinetic theory models exploit similarities between interacting grains and collidingmolecules in a dense gas [Chapman 1970]. These models incorporate energy dissipationthat is characteristic of granular flows and can be applied in certain circumstances, such asat the surface of an agitated granular mass [Esipov 1997]. The fluctuational velocitydistribution is in that case taken to be approximately Maxwellian, and continuumhydrodynamic forms of conservation laws of mass, momentum, and fluctuational kineticenergy (temperature) are then established by averaging methods [Savage 1981; Jenkins1985].

Discrete models admit numerous subclassifications, but all take the constituent grains to bedistinct and to move according to prescribed rules. Examples of discrete models includeMonte Carlo methods [Rosato 1986; Barker 1990; Hopkins 1992], which applyprobabilistic rules; cellular automata [Baxter 1991; Shinbrot 1997a], which usedeterministic (and often simplified) rules; and particle dynamics, whose rules are derivedfrom first principles. The use of the latter methods for granular flows is supported by thesuccess of molecular dynamic simulations for gas and liquid systems. Furthermore, with

- 2 -

Page 3: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

the increase in computational power, the use of particle dynamics techniques hasproliferated in recent years [e.g. Walton 1986; McNamara 1992; Goldhirsch 1993; Luding1994a; Constantin 1995; Gallas 1996; Ristow 1996; Muguruma 1997; Bizon 1998].

Two types of particle dynamics methods are most common: 'hard-particle' methods, inwhich collisions are instantaneous and binary, and 'soft-particle' methods, in whichcollisions can be lasting and multiple. Hybrid [Hopkins 1991; Louge 1994; McCarthy1998] and mean-field [Shinbrot 1997b] methods have also been proposed. Typically,hard-particle methods are most useful in rapid granular flows, where collisions are discreteand distinct. In this case, one determines the outcome of each individual collision andevolves particle trajectories ballistically between collisions. Computational load is occupiedin this case by re-ordering future events (collisions) following each collision. Soft-particlemethods, by contrast, are used in situations where contacts are enduring, rather thandistinct. In this situation, particles are permitted to suffer minute deformations, and thesedeformations are used to compute restoring elastic, plastic, and frictional forces. Ofnecessity, particle contacts persist in tumbling blenders, so we adopt the soft-particleapproach here.

3. Soft-particle method

The soft-particle method was developed by Cundall [Cundall 1971; Cundall 1979], and hasbeen used to simulate chute flow [Dippel 1996], heap formation [Luding 1997], hopperdischarge [Thomson 1991; Zhang 1992; Ristow 1994a], vibrated beds [Gallas 1992;Luding 1996; Bizon 1998], and flows in rotating drums [Ristow 1996; Khakhar 1997;Wightman 1997]. Several excellent reviews are available, for example Mehta [1994]Ristow [1994b], Schäfer [1996] and Dippel [1998].

In the present study, the granular material is idealized as a collection of frictional andinelastic spherical particles. Each particle may interact with its neighbors or with theboundary of the tumbler through both normal and tangential forces. The elastic modulusand computational timestep are chosen so that deformations of particles remain small whencompared with their displacements and diameters [cf. Luding 1994b].

The normal forces that develop between particles in contact are calculated using the"partially latching spring" model of Walton and Braun [1986]. This model approximatesthe elastic-plastic behavior of contacting spheres observed in laboratory experiments[Goldsmith 1964; Hopkins 1991] and in finite element calculations [Walton 1993b]. Thenormal force F ij

n acting on particle i, with position vector ri and radius Ri, resulting fromits interaction with particle j, with position vector rj and radius Rj, is taken to be thefollowing function of the overlap αij = |ri-rj| - (Ri+Rj) for loading and unloadingrespectively:

F ijn =

K1αijnij , ˙ α ij ≥ 0 ( loading)

K2(αij − αo )nij , ˙ α ij ≤ 0 (unloading)

, (1)

where nij = ( r i − r j) r i − r j is the unit vector joining the centers of the two particles,and K1 and K2 are the normal stiffness coefficients for loading and unloading,respectively1. The stiffness coefficient for unloading is taken to be larger than that ofloading, after Walton [1986]. For simple, binary collisions, this model produces a

1 In our simulations, we use K1 = 3000 N/m, K2 = K1/ε2, ε = 0.6

- 3 -

Page 4: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

coefficient of restitution, ε = K1/K2, that is independent of the relative velocity of impact[Schäfer 1996].

Tangential forces are calculated using Walton's [1993a] three-dimensional "incrementallyslipping" model. This model is based on Mindlin's [1949] theoretical work on frictionalcontact. After contact occurs, static frictional force builds up nonlinearly withdisplacements in the tangent plane of contact. Once a threshold of maximum staticfrictional force is reached, bodies slide with constant dynamic friction coefficient .

We track particle interactions using linked list algorithms described elsewhere [e.g. Grest1989], and integrate the resulting equations of motion using a 'leap-frog' algorithm2 [Allen1987]. The equations of motion are written as a system of two sets of ordinary differentialequations: one for positions and the other for velocities. The latter is integrated first andthen the former. After this, the accelerations (forces) are evaluated for the next time step.

Particles in a tumbling blender can be in relative motion with respect to the blender'sboundary, which in turn rotates with respect to a fixed frame. To simulate particle motionswithin vessels undergoing a general rotational motion (not necessarily a simple rotationabout a fixed axis), it is useful to write the equations of motion in a frame attached to thevessel. The advantage of this choice is that contact detection between particles and theblender's inner wall, and computation of the resulting particle-wall forces, are considerablysimplified.

Figure 1 - geometries of double cone & V-blender

For the containers in our simulations, we use smooth but frictional boundaries, defined asfollows. Referring to Fig. 1, for the double-cone we use two sizes. In monodispersesimulations, we use a tumbler with cone angle A = 45°, diameter 2R = 7.62 cm, coneheight D = 3.24 cm, and cylinder height 2L = 2.67 cm, while in bidisperse simulations weuse a tumbler with cone angle A = 45°, diameter 2R = 25.4 cm, cone height D = 10.795cm, and cylinder height 2L = 8.9 cm. For the V-blender, we compose two identicalcylindrical shells joined at an angle a = 90°, where each shell has diameter d = 7.62 cm,maximum length l = 11.43 cm, and distance perpendicular to the axis of rotation from theaxis to the intersection between shell central axes s = 2.69 cm. Our experimental containers

2 The advantage of using the leap-frog integrating scheme is that it is second order accurate, explicit,simple, and does not require the storage of auxillary variables. Predictor-corrector integrating schemesby contrast can provide more accurate solution to the equations of motion if at least two or threecorrection steps are used. This requires expensive force evaluations, and consequently these schemes arerarely used in molecular and particle dynamics simulations [Allen 1987]. This is also true for higherorder methods such as Runga-Kutta algorithms.

- 4 -

Page 5: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

are made of Plexiglas™ and are of identical geometry, with the exceptions that the physicaldouble-cone ends are rounded, and the physical V-blender arms meet at an angle a = 84°.

4. Case 1: Mixing and segregation in the double-cone blender

Mixing Mechanism:

Prior experimental studies [Brone 1997, 1998a] of mixing and segregation in the double-cone have laid the groundwork for this research. The qualitative picture painted by thesestudies is that blenders with a reflectional symmetry suffer impeded transport across thesymmetry plane. This picture is reinforced and quantified in our particle dynamics study.A synopsis3 of our results appears in Fig. 2. Here we follow 15,000 identical spheres ofdiameter 2 mm (fill 44% of total volume) through 6 revolutions of the blender at 15 rpm.Particles are initially loaded from above in random, non-overlapping positions and areallowed to fall to rest before the simulation is begun. After one revolution, we sketch theapproximate boundary of the free surface with a dashed line in the figure. Particles initiallyin the upper right quadrant are colored yellow, and roughly speaking it is observed that theyellow particles travel from top to bottom once per cycle as material is turned over in thetumbler. This is an effect of the level of fill (to be discussed in future papers): a less-filledtumbler would turn its contents over more times in a single revolution that a more-filledtumbler.

Nevertheless, the simulations reveal the essential convective motion that one would naivelyexpect to find in a tumbling blender. Beyond this, Fig. 2 shows that after 6 revolutions,the yellow particles appear to be uniformly dispersed within the right half of the tumbler,but very little transport is evident between the two halves of the blender. For comparison,we show snapshots of an experiment in which identical but colored 1.6 mm spherical glassbeads are tumbled in a double-cone at 16 rpm for 6 revolutions. In both simulation andexperiment, particle motion evidently consists of an azimuthal convective flowaccompanied by a distinct axial diffusive transport, with a significant bottleneck to mixingacross the symmetry plane of the tumbler. We quantify these observations next.

Figure 2 - Top: time sequence of front view from particle dynamicsimulations of monodisperse particles differing only by color in double-cone blender. N represents the number of revolutions experienced bythe tumbler. The top of the tumbler is rotated frontward, and the freesurface is outlined by a dashed line at N = 1. Bottom: correspondingexperimental snapshots for nearly identical initial and operatingconditions in a transparent tumbler. The tumbler is stopped at theindicated time for each snapshot and is emptied and re-loaded forsubsequent snapshots.

3 For simulation, see http://sol.rutgers.edu/~shinbrot/Moakher/DC.mpg.

- 5 -

Page 6: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

In Figure 3, we display the transport of initially vertical colored stripes of spheres. Theloading and subsequent motion of particles are as before; we have merely applied differentcolors to particles whose centers pass through specified regions of space. At N = 0.5 weagain outline the free surface with a dashed line.

Figure 3 - Time sequence of initially vertical colored stripes ofparticles in the same double-cone simulation shown in Fig. 2. Mixingis comparatively rapid within either half of the tumbler, but mixingacross the symmetry plane is slow.

Quantification of Mixing Rate:

By tracking the axial coordinates of particles initially within each vertical stripe, we canevaluate the axial mixing flux in the double-cone, as shown in Fig. 4. In the three upperinsets, we plot the concentrations of the outermost (light gray), central (gray) andinnermost (black) particles as a function of axial position after 1, 3 and 6 revolutions of thetumbler. The colored horizontal lines in the upper insets correspond to uniformconcentrations4. In the main plot, we display the logarithm of the variance of concentrationas a function of the number of tumbler revolutions for each of these three sets of particles.

From the main plot of Fig. 4, we note that the upper curve -- corresponding to particlesinitially near the axial center of the tumbler -- has a nearly exponential scaling region,extending from about 0.5 to 6 revolutions. The lower curve -- corresponding to particlesinitially near the axial extremes -- is more curved, indicating that the mixing is nearlyexponential -- i.e. diffusive [Brone 1998b] -- near the axial center of the blender, but ismore complex toward the axial extremes. This fits with the qualitative picture mentionedearlier: mixing within either half of the blender occurs through a combination of convectionand diffusion, while mixing between halves is chiefly diffusive. Likewise the upper insetsto Fig. 4 show the evolution of concentrations of particles initially in the colored verticalstripes and reflect the same behavior. Practically speaking, the implication is clear:breaking the symmetry of the blender is key to accelerating mixing [cf. Brone 1997].

4 Each of these three regions contains a different number of particles, hence concentrations are normalizeddifferently within each region. For this reason, the indicated concentration corresponding to a uniformblend differs in each inset.

- 6 -

Page 7: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Figure 4 - Main plot: logarithm of variance of volume elements indouble-cone blender. Three sets of initial conditions -- outermost (lightgray), central (gray) and innermost (black) -- are displayed in thesecalculations, as identified in the lower inset. Upper insets showconcentrations of particles originating in outermost, central andinnermost regions after 0, 3, and 6 revolutions.

Velocity Fields:

A great advantage of computational models is that one can easily extract useful results thatwould be difficult to obtain experimentally. In Fig. 5, we display the mean velocity fieldfor the double-cone at 8 phases of the rotation cycle. Each velocity vector is formed byaveraging over both space and time: over space, the velocities of particles lying withinsmall cubic volume elements are averaged, and over time, velocities of particles at the phaseof rotation indicated are averaged over the final 4 cycles of tumbler motion. Apparently ateach phase there is a distinct flowing layer near the surface and region beneath that movesslowly as a single unit. This is in keeping with the current understanding of tumbler flowsdescribed in §2, however this is not always the case in 3D tumblers, as will be shownshortly.

A top view of the velocity field (right of Fig. 5) shows two additional effects: first, there isa sizeable gradient in velocities, with the highest velocities near the center of the flowinglayer, and second the velocity field curves to follow the profile of the container sides. Thiscurvature leads to significant mixing and segregation effects. This is another facet of thephenomenon described in Fig. 4, indicating that convective axial mixing is enhancedfurther from the axial center of the tumbler. Moreover, we can infer segregationalimplications from this velocity field: large, fast moving, spheres should travel more easilyalong straight paths than smaller spheres. Correspondingly, the velocity field shown inFig. 5 suggests that for polydisperse mixtures in the flow regime studied here, largerparticles may migrate toward the high speed rectilinear flow region in the center of thetumbler, leaving smaller particles to migrate to follow the slower, curved pathlines near theaxial extremes.

- 7 -

Page 8: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Figure 5 - Left plots: side view of velocity field at successive phases, τ= N mod 1, of rotation of double-cone blender. These views show aslice through the axial center of the tumbler; each plot displays thevelocity field averaged over space and time as described in text. Rightplot: top view of velocity field taken of horizontal slice throughblender at τ = 7/8. Magnitude of velocities are coded by length ofvectors shown.

Segregation in the Double-cone:

This suggestion is only a heuristic, yet it appears to be followed in this system in the flowregime studied. In Fig. 6(a), we display a top view of a simulation5 using a bidispersemixture of approximately equal masses of large and small beads. Here we have used11,500 small (6 mm diameter, blue) spheres initially randomly mixed with 1500 large (12mm diameter, red) spheres, and we have rotated the blend at 30 rpm for 12 revolutions.For comparison, also in Fig. 6(b), we display a snapshot of an experiment in which werotated an equal mass blend of 1.6 mm (blue) and 4 mm (red) diameter glass beads at 16rpm in a 1 quart double-cone blender. The banded segregation pattern shown begins to beformed in the first few rotations of the tumbler; in the experiment we have the luxury ofcontinuing to rotate indefinitely and we find that the pattern changes little after the firstseveral revolutions. In both experiment and simulation, the tumbler is initially filled to50% of its total volume. In Fig's 6(c)-(d), we also show the segregation patterns in theinterior of the simulation and experiment respectively. These figures display a view fromthe side of the blender intersecting a vertical plane through the blender's rotation axis(indicated by green arrows in Fig's 6(a)-(b)). Simulation and experiment show that thesmall grains form a contiguous core parallel to the axis of rotation.

5 For simulation, see http://sol.rutgers.edu/~shinbrot/Moakher/DCseg.mpg.

- 8 -

Page 9: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Figure 6 - (a) Top view of segregation in double-cone tumbler of 12mm (red) and 6 mm (blue) spheres after 12 revolutions at 30 rpm. (b)Top view of experiment for comparison, using 4 mm (red) and 1.6 mm(blue) glass spheres. (c) Slice through vertical plane passing throughaxis of rotation of blender, showing interior structure in simulation.(d) Slice through corresponding vertical plane in experiment.

To produce the experimental view in Fig. 6(d), we inserted a thin metal plate along thevertical plane of interest to divide the blender into equal halves. We then poured meltedparaffin wax (indicated in Fig. 6(d)) onto the top of one half of the blender and vacuumedthe grains away from the other half. Once the paraffin set, it held the grains in place so thatthe tumbler could be tipped with the metal plate facing up. The plate was then removed andthe topmost exposed layer vacuumed to reveal the interior structure shown.

Quantitatively, we evaluate the segregation rate in the bidisperse system by plotting theintensity of segregation as a function of the number of revolutions (Fig. 7). The intensityof segregation, I, is defined as:

I = (Ci − C )2

i=1

n∑n −1

, (2)

where Ci is the concentration of one species of particle, C is its mean concentration, n isthe number of samples taken, and σr is the standard deviation expected from the samenumber of samples taken from a random mixture. To evaluate I, we subdivide theoccupied volume into cubic elements and eliminate elements containing uncharacteristicallyfew particles; in this way we prevent spurious results caused by sampling anomalouselements near boundaries. The fit shown in gray is to the exponential function,

- 9 -

Page 10: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

I = Io-A exp(-k N), in keeping with suggestions from the literature that segregation may bedescribed by first order kinetics [Zik 1994; Cantelaube 1997; Choo 1997; Aranson 1998].

Figure 7 - Growth of intensity of segregation over time in simulationof double-cone. Gray curve is best fit to exponential function definedin text.

5. Case 2: Mixing and segregation in the V-blender

Mixing Mechanism:

The mixing mechanism that we have seen for the double-cone is straightforward: there is aco-existence between comparatively rapid convective mixing and much slower dispersivemixing. Furthermore, the velocity field is symmetric and (though significantly curved)apparently well behaved and amenable to straightforward analytic treatment.

In the V-blender this picture changes in several respects6. We begin our treatment of thisproblem by examining mixing of initially vertical colored stripes of identical particles. InFig. 8 (top), we display the evolution of these stripes over the first revolution of thetumbler; beneath these plots we display the evolution of initially horizontal stripes. Westress two observations from these plots. First, particles in the yellow and cyan verticalstripes, beginning at the outward tips of the blender, end up spread along its front surfacesafter one revolution; likewise the green particles beginning at the top horizontal stripe of thetumbler are brought to its front and center. This is representative of the perhaps obviousfact that the geometry of the blender enforces convective motion that contains a large axial,as well as radial, component (discussed in greater depth shortly). Second, the overallmotion of grains over one cycle of rotation of this blender consists of two differentoperations. Particles in either shell of the tumbler are split apart at N ≅ 0.5, and are mergedtogether at N ≅ 1. This fact, while perhaps intuitively obvious, has subtle, yet significant,consequences for mixing.

6 For simulation, see http://sol.rutgers.edu/~shinbrot/Moakher/VB.mpg.

- 10 -

Page 11: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Figure 8 - Top: evolution of initially vertical colored stripes ofidentical spheres in V-blender. Bottom: evolution of the same spheres,initially tagged in horizontal colored stripes.

Let us examine front and back views of identical beads colored red to the left of the planeseparating the two shells of the blender and blue to the right. After two blenderrevolutions, front and back views are shown in Fig. 9 for the V-blender simulation (a), fora companion experiment (b) and for the double-cone simulation (c). Evidently, in the V-blender, both simulation and experiment exhibit substantially more mixing across thesymmetry plane in the back view than in the front. This front/back asymmetry is notobserved in the double-cone. To our knowledge, this asymmetry has not been reportedpreviously, yet its cause appears to be central to the operation of the V-blender.

The result of the merging operation is seen in the front view of the blender, while the resultof the splitting phase is seen in the back view. Thus the splitting and merging of the twohalves of this blender commented upon earlier have the effect of forcing material near thesymmetry plane to choose, presumably nearly at random, which blender shell to enterduring every other half-period of rotation, when the shells are facing down (N = 0.5 inFig. 8). It is during this phase of motion of the blender that mixing between the halvesoccurs. This holds lessons -- to be discussed elsewhere -- for blender designimprovements.

Beyond this potentially applicable result, we have also detected possible signatures ofintriguing dynamical structure in the V-blender's mixing flow. Although the significanceof this structure remains to be determined at this stage, recent results indicating the presenceof deterministic structure in other granular mixing problems [Shinbrot 1998], suggest thatthis observation may nonetheless be worth reporting. This structure is revealed in Fig. 10,which shows a side-view of the same horizontally layered initial state discussed previously(Fig. 8, bottom). In the side-view of this simulation, we see two structures of potentialimportance. At N = 0.5, in the center of the blend, we see a spiral shape, which is likelyjust a reflection of the overturning of the granular mass about a central core. Nevertheless,the presence of this spiral indicates that there must be significant shear in this region. Sinceshear tends to expel larger particles [Lacey 1943, Leighton 1987], we expect to find highconcentrations of fines here, and indeed this is what we observe in experiments usingpolydisperse particles. Second, at N = 1, we observe a diagonal swath of connected blueparticles, which originated at the bottom of the tumbler. Since the swath remainscontiguous, despite being evidently stretched along a particular direction, this reinforces thesuggestion made elsewhere [Shinbrot, 1998] that properties of chaotic mixing of fluids --such as asymptotic directionality [Alvarez, 1998] and the view of mixing as beingrepresentable by repeated iterations of simple topological operations [Aref, 1984] -- may

- 11 -

Page 12: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

carry over into the study of mixing of grains. Clearly more work is needed to substantiatethis physical picture, but our simulations seem consistent with this view.

Figure 9 - Top: front views of V-blender and double-cone; bottom: backviews. (a) V-blender simulation after 6 revolutions, with identicalparticles initially blue to the right of the symmetry plane and red to theleft. (b) Experiment in transparent V-blender, again after 6 revolutions.In both simulation and experiment, more mixing is apparent in viewfrom back than from front. (c) By comparison, simulation of double-cone reveals nearly symmetric mixing from front and back after 6revolutions.

Figure 10 - Evolution of identical spheres initially layered horizontally.These views show a slice through the center of one shell of thetumbler.

Velocity Fields:

As with the double-cone blender, it is instructive to examine the velocity field within the V-blender. This is shown in Fig. 11, where each snapshot represents velocities of particleswhose centers lie within a square lattice of cells averaged over the same phase during thefinal 4 cycles of tumbler motion. On the left, we see the velocity field in planar slicestraveling through the center of one shell of the blender. Importantly, the velocity field heredoes not consist of a flowing layer above a solid-body region, as was seen in the double-cone and as one might surmise from existing literature. Instead, transport consists of

- 12 -

Page 13: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

phases of motion during which nearly all of the grains are in motion, either splitting fromthe blender center toward its arms (upper right of Fig. 11) or merging from arms to center(lower right). This is in keeping with the picture described earlier in this section, but thevelocity fields in addition reveal that between these two stages, the granular bed is largelystatic. Thus flow in the V-blender consists of intermittent periods of near-universal motionof particles alternating with periods of near-universal stationarity. This agrees withexperimental observations, in which we have used transparent vessels, and where similarlywe see rapid sloshing motions of particles alternating with periods of substantial calm.

We quantify this behavior in the inset to Fig. 15 for both double-cone and V-blender. Herewe plot the fraction of particles traveling with speed V > Vc as a function of the phase ofrotation of the tumbler, τ. For this plot, we choose a value for the cut-off speed, Vc, of 3times the time-averaged speed of all particles. From this inset plot, we confirm thatthroughout the rotation of the double-cone, a small (8%) and nearly constant fraction ofparticles flow 'rapidly' by this measure, while for the V-blender, there are strongfluctuations between a state in which < 5% of particles flow rapidly and a state in which >25% of particles flow rapidly. If Vc is taken to be the mean particle speed, the sloshingmotion in the V-blender causes up to 68% of particles to periodically move with speed >Vc.

By examining the snapshots to the right of Fig. 11, taken at opposite phases of the tumblerrotation during which collective particle speeds are near their maximum, we see directly thealternating splitting and merging of the granular bed that appears to govern mixing in thisflow.

Figure 11 - Left plots: side view of velocity field at successive phases,τ = N mod 1, of rotation of V-blender. These views show a slicethrough the center of one shell of the tumbler; each plot displays thevelocity field averaged over space and time as described in text. Rightplots: top views of velocity field taken on horizontal slices throughblender at τ = 1/4 and τ = 3/4, when mean speeds are maximal. Thearrows in the right plots indicate the plane through which velocities inthe left plots are taken.

- 13 -

Page 14: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Quantification of Mixing Rate:

Different as this velocity field is from simpler blenders such as the drum or double-cone, itcomes as little surprise that the mixing behavior of the V-blender is unique. In Fig. 12, weplot the change in variance with time in a monodisperse simulation of 20,000 3.2 mmdiameter particles tumbled at 30 rpm. Unlike the double-cone (Fig. 4), this tumbler seemsto lack a simple exponential scaling region, and the merging/splitting action is manifestedby strong oscillations in the variance. On the other hand, neglecting the barrier to transportacross the symmetry plane commented on earlier, the mean rate of mixing is much faster inthe V-blender than in the double-cone: the same decrease in variance seen for the double-cone in 6 revolutions is achieved in the V-blender in under 1 revolution. Likewise, by thethird revolution the V-blender has nearly achieved its asymptotic mixed state. Note that thelight gray curve in the main plot has a different asymptote than the other curves becausethere are fewer light-colored particles than represented by the other shades.

Figure 12 - Main plot: logarithm of variance of volume elements in V-blender. Three sets of initial conditions -- outermost (light gray),central (gray) and innermost (black) -- are displayed in thesecalculations, as identified in the lower inset. Upper insets showconcentrations of particles originating in outermost, central andinnermost regions after 0, 3, and 6 revolutions.

Segregation in the V-blender:

We can see the effect of splitting and merging on dissimilar particles7 in Fig. 13(a). Herewe display the evolution of an initially randomly placed arrangement of 16000 small (2.5mm diameter, blue) and 2000 large (5 mm diameter, red) spheres. After 6 revolutions at 15rpm, an asymmetric segregated state is established. From the front, in this state the smallerparticles appear to have migrated outward. In a view from the back, however, it becomesclear that something more complicated is going on: the smaller particles are seen toaccumulate near the center. This is seen in the enlarged view in Fig. 13(b). A slice

7 For simulation, see http://sol.rutgers.edu/~shinbrot/Moakher/VBseg.mpg.

- 14 -

Page 15: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

through the center plane of the tumbler (Fig. 13(c)) shows that the small particles actuallyoccupy a horizontally oriented, V-shaped, region.

Figure 13 - (a) Front view of segregation in V-blender of 5 mm (red)and 2.5 mm (blue) spheres. (b) Enlarged rear view of segregated stateafter N = 6 revolutions. (c) Slice through center of N = 6 state.

Comparisons between simulation and experiment in both upright and inverted orientationsof the tumbler are shown in Fig. 14. For the experiments, we half filled a transparent 1quart V-blender with equal masses of 1.6 mm (blue) and 4 mm (red) glass spheres androtated the blend at 16 rpm until an asymptotic segregation pattern was formed. As withthe double-cone experiment, the segregation pattern forms rapidly and changes little afterseveral revolutions of the tumbler. Evidently, in the upright configuration, Fig's 14(a)-(b),the horizontal V-shaped segregation pattern is seen in both experiment and simulation. Asanticipated earlier (cf. Fig. 10), in both experiment and simulation, a region of finerparticles emerges close to the axis of rotation. There is likewise correspondence betweenexperiment and simulation for the inverted orientation of the tumbler (Fig's 14(c)-(d). Inthis orientation, however, the smaller particles arrange themselves into two symmetricwedges at the extremes of either shell.

Figure 14 - Top: front views of V-blender; bottom: back views. (a)Experiment with different size particles in upright transparent V-blender. (b) Corresponding simulation after 2 revolutions. (c)-(d)Experiment and simulation in inverted orientation; simulation shownafter 2-1/2 revolutions.

- 15 -

Page 16: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

We have also evaluated the growth in intensity of segregation, I, for the V-blender, asdisplayed in Fig. 15. As before, a fit to the exponential function I = Io-A exp(-k N) isshown in gray. For the V-blender, there appears to be a variation in I with period roughlyone per revolution of the tumbler. In Fig. 15 we indicate approximate times at whichmerging of particles from opposite tumbler shells occurs. These times roughly correspondto minima in I (which are absent from the double-cone plot: Fig. 7). This suggests thatwhile mixing of similar size particles across the symmetry plane occurs predominantlyduring splitting of particles into the blender shells, mixing of dissimilar size particles occursat the opposite phase of motion, during merging from the shells. This is intriguing andindicates the need for more detailed dynamical studies, specifically focused on splitting andmerging as occurs both in tumblers such as the V-blender and in chutes and near baffles.

Figure 15 - Main plot: growth of intensity of segregation over time insimulation of V-blender. Gray curve is best fit to exponential functiondefined in text. Arrows indicate approximate times at which merging ofparticles from tumbler shells occurs. Inset: fraction of rapidly movingparticles vs. phase of tumbler rotation for double cone and V-blender.

6. Conclusion:

We have presented results from the first fully three-dimensional simulations of twoindustrially relevant tumblers: the double-cone and the V-blender. It appears from thesestudies that the detailed dynamics in these blenders are quite different, yet both share thesame mixing bottleneck: dispersion across a plane of symmetry. Segregation in thesetumblers likewise share commonalities, with asymptotic segregated patterns becomingestablished quite rapidly and persisting indefinitely. The patterns themselves look ratherdifferent, yet effectively serve the function of separating large and small particles into zoneswith substantially different velocity fields, each compatible with a particular subspecies.

- 16 -

Page 17: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Several avenues are indicated for future research. First, a lesson from mixing of fluids isthat wherever possible symmetries should be broken. This lesson carries over directly togranular research, and implies that new tumbler designs incorporating broken symmetries-- for example through judicious baffle design [Brone 1997], temporal perturbations[Wightman 1997; Brone 1997, 1998a], or reshaping the geometry of the tumbler [Chang,1992] -- are merited. Second, tumblers such as the V-blender appear to function verydifferently from tumblers like the double-cone. In particular, the V-blender operatesintermittently, combining splitting and merging, while the double-cone operates nearlycontinuously, with a nearly constant flow of particles in a more uniform surface layer.This results in significantly more rapid mixing in the V-blender than in the double-cone;nevertheless both tumblers exhibit reproducible and rapidly occurring segregation patterns.Initial inroads have been made into modeling of constant [Khakhar 1997] or near constant[Shinbrot 1998] tumbler flows, but strongly intermittent granular flow is much less wellunderstood. Since a large class of practical tumblers appear from this work to operateintermittently, research in this field is also warranted. Likewise, this intermittency indicatesthat hybrid and continuum approaches, while justifiable for some blenders, must be viewedwith caution for others. Finally, from a fundamental standpoint, the apparent facts thatsimilar particles mix during mechanical separation and dissimilar ones mix duringrecombination are unexpected and deserve more detailed study.

7. Acknowledgements:

We gratefully acknowledge support from the International Fine Powder Research Institute,the New Jersey Commission on Science and Technology and Pfizer Inc., and technicalassistance from Adela Abad and Jason O'Leary.

- 17 -

Page 18: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

References:Allen, M.P. & Tildesley, D.J. 1987 Computer simulation of liquids (Oxford University Press, Oxford).

Alvarez, M.M., Muzzio, F.J., Cerbelli, S., Adrover, A. Giona, M. 1998 "Self-Similar SpatiotemporalStructure of Intermaterial Boundaries in Chaotic Flows" Phys. Rev. Lett. 81 3395-3398.

Aranson, I. & Tsimring, L.S. 1998 "Dynamics of axial separation in long rotating drums" (preprint,Argonne National Laboratories)

Aref, H. 1984 "Stirring by Chaotic Advection" J. Fluid Mech. 143 1—21.

Baxter, G.W. & Behringer, R.P. 1991 "Cellular automata models for the flow of granular materials,"PhysicaD 51 , 465-471.

Bizon, C., Shattuck, M.D., Swift, J.B., McCormick, W.D.& Swinney, H.L. 1998 "Patterns in 3DVertically Oscillated Granular Layers: Simulation and Experiment" Phys. Rev. Lett. 80 , 57-60.

Brone, D., Wightman, C., Connor, K., Alexander, A., Muzzio, F.J. & Robinson, P. 1997 "Using flowperturbations to enhance mixing of dry powders in V-blenders," Powder Technol. 91 , 165-172.

Brone, D. & Muzzio, F.J. 1998a "Enhanced mixing in double cone blenders," (preprint, RutgersUniversity).

Brone, D., Alexander, A. & Muzzio, F.J. 1998b "Quantitative characterization of mixing of dry powders inV-blenders," AIChE J. 44 , 271-8.

Brone, D.,Wightman, C., Conner, K., Alexander, A., Muzzio, F.J. & Robinson, P. 1997 "Using flowperturbations to enhance mixing of dry powders in V-blenders," Powder Technol. 91 165-72;

Cantelaube, F. Bideau, D. & Roux, S. 1997 "Kinetics of segregation of granular media in a two-dimensional rotating drum," Powder Technol. 93 1-11.

Carley-Macauly, K.W. & Donald, M.B. 1962 "The mixing of solids in tumbling mixers--I," Chem. Eng.Sci. 17 , 493-506.

Carley-Macauly, K.W. & Donald, M.B. 1964 "The mixing of solids in tumbling mixers--II," Chem. Eng.Sci. 19 , 191-199.

Chang, R.-K., Chang, S.-I. & Robinson, J.R. 1992 "A study of the performance of a modified V-shapedsolids mixer using segregating materials," Int. J. Pharm. 80 , 171-178.

Chang, R.-R., Badawy, S., & Buehler, J.D. 1995 "A comparison of free-flowing, segregating and non-free-flowing, cohesive mixing systems in assessing the performance of a modified V-shaped solid mixer,"Drug Development Ind. Pharm. 21 , 361-368.

Chapman, S. & Cowling, T.G. 1970 The Mathematical Theory of Non-Uniform Gases (Cambridge Univ.Press).

Choo, K., Molteno, T.C.A & Morris, S.W. 1997 "Traveling granular segregation patterns in a long drummixer" Phys. Rev. Letters 79 2975-8.

Constantin, P., Grossman, E.L. & Mungan, M. 1995 "Inelastic Collisions of Three Particles on a Line asa Two-Dimensional Billiard" Physica D 83 409-•••

Cundall, P.A. 1971 "A computer model for simulating progressive, large-scale movements in blocky rocksystems," In Proc. Symp. Int. Soc. Rock Mech., Volume2, Nancy, France, 129-136.

Cundall, P.A. & Strack, O. D.L. 1979 "A discrete numerical model for granular assemblies" Geotechnique29 , 47-65.

Dippel, S., Batrouni, G.G. & Wolf, D.E. 1996 "Collision-induced friction in the motion of a singleparticle on a bumpy inclined line" Phys. Rev. E 54 6845-•••.

Dippel, S. 1998 "Microscopic dynamics of granular materials" dissertation from Univ. Jülich.

Duke, T. A.J., Barker, G.C., & Mehta, A. 1990 "A Monte Carlo study of granular relaxation," Europhys.Lett. 13 , 19-24.

Esipov, S.E. & T. Pöschel, T. 1997 "The Granular Phase Diagram" J. Stat. Phys. 86 1385-•••.

Fan, L.T., Chen, S.J., & Watson, C.A. 1972 "Solids mixing," In Annual Review of Industrial andEngineering Chemistry, ed. V.M. Weekman (Amer. Chem. Soc.) 22-56.

- 18 -

Page 19: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Fan, L.T., Chen, Y.-M., & Lai, F.S. 1990 "Recent developments in solids mixing," Powder Technol.61 , 255-287.

Gallas, J.A.C. Herrmann, H.J. Pöschel, T. & Sokolowski, S. 1996 "Molecular dynamics simulation ofsize segregation in three dimensions" J. Stat. Phys 82 , 443-•••.

Gallas, J.A.C., Hermann, H.J. & Blumen, A. 1992 "Convection Cells in Vibrating Granular Media,"Physica A 189 437-•••.

Goldhirsch, I. & Zanetti, G. 1993 "Clustering Instability in Dissipative Gases" Phys. Rev. Lett. 70 1619-•••.

Goldsmith, W. 1964 Impact, the Theory and Physical Behavior of Colliding Solids. (Edward Arnold-•••)

Goodman, M.A. & Cowin, S.C. 1972 "A continuum theory for granular materials," Arch. RationalMech. Anal. 44, 249-266.

Grest, G.S. Dünweg, B. & Kremer, K. 1989 "Vectorized link cell Fortran code for molecular dynamicssimulations for a large number of particles," Comput. Phys. Commun. 55 269-85.

Hobbs, D.M. & Muzzio, F.J. 1997 "The Kenics static mixer: A three-dimensional flow," Chem. Eng. J.67 , 153-166.

Hogg, R. & Fuerstenau, D.W. 1972 "Transverse mixing in rotating cylinders," Powder Technol. 6 139-•••.

Hopkins, M.A. & Louge, M.Y. 1991 "Inelastic microstructure in rapid granular flows of smooth disks,"Phys. Fluids A 3 47-•••

Hopkins, M.A. & Shen, H.H. 1992 "A Monte Carlo solution for rapidly shearing granular flows based onthe kinetic theory of dense gases," J. Fluid Mech. 244 , 477-491.

Jenkins, J.T. & Richman, M.W. 1985 "Kinetic theory for plane flows of a dense gas of identical, rough,inelastic, circular disks," Phys. Fluids 28 , 3485-3494.

Johnson, P.C. & Jackson, R. 1987 "Frictional-collisional constitutive relations for granular materials, withapplication to plane shearing" J. Fluid. Mech. 176 , 67-•••.

Khakhar, D.V., McCarthy, J.J., Shinbrot, T. & Ottino, J.M. 1997 "Transverse flow and mixing ofgranular materials in a rotating cylinder," Phys. Fluids 9 31-43.

Killmeyer, R.P. 1993 National Coal Preparation Process Research Facility report.

Lacey, P.M.C.1943 "The mixing of solid particles" Trans. Inst. Chem. Eng. 21 , 53.

Leighton, D. & Acrivos, A. (1987) "Measurement of shear-induced self-diffusion in concentratedsuspensions of spheres," J. Fluid Mech. 177 , 109-31.

Louge, M.Y. 1994 "Computer simulations of rapid granular flows of spheres interacting with a flat,frictional boundary" Phys. Fluids A 6 2253-•••.

Luding, S. Clément, E. Blumen, A. Rajchenbach, J. & Duran, J. 1994a "The onset of convection inmolecular dynamics simulations of grains" Phys. Rev. E 50 R1762-•••.

Luding, S., Clément, E., Blumen, A., Rajchembach, J. & Duran, J. 1994b "Anomalous energy dissipationin molecular-dynamics simulations of grains: the 'detachment' effect" Phys. Rev. E 50 4113-22.

Luding, S., Clément, E., Rajchenbach, J. & Duran, J. 1996 "Simulations of pattern formation in vibratedgranular media," Europhys. Lett. 36 247-52.

Luding, S. 1997 "Stress Distribution in static two dimensional granular model media in the absence offriction" Phys. Rev. E 55 4720-•••.

Lun, C. C. K., Savage, S. B. Jeffery, D. J. & Chepurniy, N. 1984 "Kinetic theories for granular flow:inelastic particles in Couette flow and slightly inelastic particles in a general flowfield," J. FluidMech. 140 , 223-256.

McCarthy, J.J. & Ottino, J.M. 1998 "Particle dynamics simulation: a hybrid technique applied to granularmixing", Powder Technol. 97 91-99.

McNamara S. & Young, W.R. 1992 "Inelastic collapse and clumping in a one-dimensional granularmedium" Phys. Fluids A 4 496-•••.

McTigue, D.F. 1987 "A nonlinear continuum model for flowing granular materials," PhD thesis, StanfordUniv.

Mehta, A. (Editor) 1994 Granular Matter, an interdisciplinary approach, (Springer-Verlag, NY)

- 19 -

Page 20: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Mehta, A. & Edwards, S.F. 1990 "A phenomenological approach to relaxation in powders" Physica A 168714-•••.

Mindlin, R.D. 1949 "Compliance of elastic bodies in contact," J.Appl. Mech. 16 , 259-268.

Mindlin, R.D. & Deresiewicz, H. 1953 "Elastic spheres in contact under varying oblique forces," J.Appl.Mech. 20 , 327-344.

Muzzio, F.J., Robinson, P., Wightman, C. & Brone, D. 1997 "Sampling practices in powder blending,"Int. J. Pharmaceutics 155 153-78.

Nienow, A.W., Harnby, N. & Edwards, M.F. 1992 "Introduction to mixing problems," in Mixing in theProcess Industries (Butterworth-Heinemann, Oxford) 1-24

Ristow, G. 1992 "Simulating granular flow with molecular dynamics," J. Phys. I France 2 , 649-669.

Ristow, G. & Herrmann, H. 1994a "Density patterns in two-dimensional hoppers," Phys. Rev.E 50 ,R5-R8.

Ristow, G.H. 1994b "Granular dynamics: a review about recent molecular dynamics simulations of granularmaterials," in Annual reviews of computational physics ed. D. Stauffer (World Scientific, Singapore)275-308.

Ristow, G.H. 1996 "Dynamics of granular material in a rotating drum" Europhys. Lett. 34 263-•••.

Rosato, A., Prinz, F., Standburg, K.J. & Swendsen, R. 1986 "Monte Carlo simulation of particulatematter segregation," Powder Technol. 49 , 59-69.

Savage, S.B. & Jeffrey, D.J. 1981 "The stress tensor in a granular flow at high shear rates," J.FluidMech. 110 , 255-272.

Savage, S.B. & Lun, C.K.K. 1988 "Particle size segregation in inclined chute flow of dry cohesionlessgranular solids," J.Fluid Mech. 189 , 311-335.

Sayed, M. 1981 "Theoretical and experimental studies of the flow of cohesionless granular materials,"(PhD dissertation, McGill Univ.)

Schäfer, J., Dippel, S. & Wolf, D.E. 1996 "Force schemes in simulations of granular materials," J. Phys. IFrance 6 5-20.

Shinbrot, T., Khakhar, D.V. & Ottino, J.M. 1997a "A simple model for granular convection," Phys. Rev.Lett. 79 829-32

Shinbrot, T. 1997b "Competition between randomizing shakes and inelastic collisions in granular patternformation" Nature 389 574-6.

Shinbrot, T., Alexander, A. & Muzzio, FJ 1998 "Spontaneous granular chaotic mixing" (preprint, RutgersUniversity).

Spencer, A. J.M. 1982 "Deformation of an ideal granular material," In Mechanics of Solids (PergamonPress) , eds. H.G. Hopkins & M.J. Sewell), 610-652. .

Thompson, P.S. & Grest, GS 1991 "Granular flow: friction and the dilatancy transition," Phys. Rev. Lett.67 1751-5.

Walton, O.R. 1993 "Numerical simulation of inclined chute flows of monodisperse, inelastic, frictionalspheres," Mechanics of Materials 16 , 239-247.

Walton, O.R. 1993 "Numerical simulation of inelastic, frictional particle-particle interactions," InParticulate two-phase flow (Butterworth-Heinemann), ed. M.C. Roco, 884-911. .

Walton, O.R. & Braun, R.L. 1986 "Viscosity, granular-temperature, and stress calculations for shearingassemblies of inelastic, frictional disks," J.Rheol. 30 , 949-980.

Wightman, C., Moakher, M., Muzzio, F.J. & Walton, O.R. 1998 "Use of the discrete element methodsto characterize particulate flow and mixing in a rotating and rocking cylinder," AIChE J. 44 , 1266--1276.

Zhang, Y. & Campbell, C.S. 1992 "The interface between fluid-like and solid-like behaviour in two-dimensional granular flows," J. Fluid Mech. 237 541-•••.

Zheng, X.M. & Hill, J.M. 1996 "Molecular dynamics modelling of granular chute flow: density andvelocity profiles," Powder Technol. 86 , 219-227.

- 20 -

Page 21: Experimentally validated computations of flow, …sol.rutgers.edu/~shinbrot/OnlinePapers/MDPaper.pdfExperimentally validated computations of flow, mixing and segregation of non-cohesive

Zik, O., Levine, D., Lipson, S.G., Shtrikman, S., & Stavans, J. 1994 "Rotationally induced segregationof granular materials" Phys. Rev. Letters 73 644-7.

- 21 -


Recommended