8
Discrete characterization of cohesion in gas–solid flows Kunal Jain, Deliang Shi, J.J. McCarthy * Department of Chemical and Petroleum Engineering, University of Pittsburgh, 1249 Benedum Hall, Pittsburgh, Pennsylvania 15261, United States Received 31 December 2002; received in revised form 2 August 2004; accepted 3 August 2004 Available online 23 September 2004 Abstract Despite the fact that a sizable portion of gas–solid flows are cohesive in nature, the mechanics of cohesive flowing gas–particle systems is still poorly understood and manipulation/control of the flow variables is still largely done on a trial-and-error basis. While recent advances have been made in our understanding of liquid-induced cohesion at the macroscopic level, in general, it is still not possible to directly connect this macroscopic understanding of cohesion with a microscopic picture of the particle properties and interaction forces. In fact, conventional theories make no attempt to distinguish between these modes of cohesion, despite clear qualitative differences (lubrication forces in wet systems and electrostatic repulsion are two good examples). In this work, we introduce a discrete characterization tool for gas–solid flow of wet (cohesive) granular material—the Granular Capillary Number (Ca g ). The utility of this tool, which is a ratio of the capillary force to the drag force, is computationally tested over a range of cohesive strengths in two prototypical applications of gas–solid flows. It is shown that rescaling our results in this way yields a collapse of data for varying surface tensions and fluidization velocities, and that a clear transition from free-flowing to cohesive behavior occurs at a distinct value of Ca g . D 2004 Elsevier B.V. All rights reserved. Keywords: Particle dynamics; Cohesion; Fluidization; Mixing 1. Introduction Gas–solid transport is crucial in a variety of industrially important applications. In particular, fluidization and the transport of solid particles either by gravity or by pneumatic means are used in fluid catalytic cracking, fluid hydro- forming and solid fuel processes such as coal gasification and liquefaction [1]. In many instances, the cohesive nature of a powder sample is a prime factor contributing to difficulties in powder flowability causing, for example, channeling and defluidization in combustion/feeder systems. Despite recent advances [2,3], an understanding of the flow and characterization of cohesive gas–solid flows remains poor. The most widely used classification system for particles in gas–solid flows was introduced by Geldart [4] based on the density difference between the particles and the gas (q p q g ) and the average particle diameter d p —where particles are generally considered cohesive (C), aeratable (A), sand-like (B), and spoutable (D) as particle diameter increases. However, a first principles basis for the group boundaries and/or how they change as a function of particle interaction forces is elusive. In particular, the group A/C demarcation is by no means absolute and in addition to the particle size and density also shows a dependence on a multiplicity of other factors such as the width of the particle size distribution, the particle shape and the surface texture and composition [5]. Recently, computa- tional studies of gas-fluidized beds [2,3,6,7] based on the combined approach of Particle Dynamics (PD) modeling and computational fluid dynamics have been used to investigate the impact of liquid-induced cohesion in these systems. In this paper, we first review several discrete character- ization tools [8] for cohesive particle systems then introduce a novel characterization tool specifically for gas–solid flows. Our goal is to use this tool, combined 0032-5910/$ - see front matter D 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.powtec.2004.08.001 * Corresponding author. Tel.: +1 4126247362; fax: +1 4126249639. E-mail address: [email protected] (J.J. McCarthy). Powder Technology 146 (2004) 160 – 167 www.elsevier.com/locate/powtec

Discrete characterization of cohesion in gas–solid flows

Embed Size (px)

Citation preview

Page 1: Discrete characterization of cohesion in gas–solid flows

www.elsevier.com/locate/powtec

Powder Technology 14

Discrete characterization of cohesion in gas–solid flows

Kunal Jain, Deliang Shi, J.J. McCarthy*

Department of Chemical and Petroleum Engineering, University of Pittsburgh, 1249 Benedum Hall, Pittsburgh, Pennsylvania 15261, United States

Received 31 December 2002; received in revised form 2 August 2004; accepted 3 August 2004

Available online 23 September 2004

Abstract

Despite the fact that a sizable portion of gas–solid flows are cohesive in nature, the mechanics of cohesive flowing gas–particle systems is

still poorly understood and manipulation/control of the flow variables is still largely done on a trial-and-error basis. While recent advances

have been made in our understanding of liquid-induced cohesion at the macroscopic level, in general, it is still not possible to directly connect

this macroscopic understanding of cohesion with a microscopic picture of the particle properties and interaction forces. In fact, conventional

theories make no attempt to distinguish between these modes of cohesion, despite clear qualitative differences (lubrication forces in wet

systems and electrostatic repulsion are two good examples). In this work, we introduce a discrete characterization tool for gas–solid flow of

wet (cohesive) granular material—the Granular Capillary Number (Cag). The utility of this tool, which is a ratio of the capillary force to the

drag force, is computationally tested over a range of cohesive strengths in two prototypical applications of gas–solid flows. It is shown that

rescaling our results in this way yields a collapse of data for varying surface tensions and fluidization velocities, and that a clear transition

from free-flowing to cohesive behavior occurs at a distinct value of Cag.

D 2004 Elsevier B.V. All rights reserved.

Keywords: Particle dynamics; Cohesion; Fluidization; Mixing

1. Introduction

Gas–solid transport is crucial in a variety of industrially

important applications. In particular, fluidization and the

transport of solid particles either by gravity or by pneumatic

means are used in fluid catalytic cracking, fluid hydro-

forming and solid fuel processes such as coal gasification

and liquefaction [1]. In many instances, the cohesive nature

of a powder sample is a prime factor contributing to

difficulties in powder flowability causing, for example,

channeling and defluidization in combustion/feeder systems.

Despite recent advances [2,3], an understanding of the flow

and characterization of cohesive gas–solid flows remains

poor.

The most widely used classification system for particles

in gas–solid flows was introduced by Geldart [4] based on

the density difference between the particles and the gas

0032-5910/$ - see front matter D 2004 Elsevier B.V. All rights reserved.

doi:10.1016/j.powtec.2004.08.001

* Corresponding author. Tel.: +1 4126247362; fax: +1 4126249639.

E-mail address: [email protected] (J.J. McCarthy).

(qp�qg) and the average particle diameter dp—where

particles are generally considered cohesive (C), aeratable

(A), sand-like (B), and spoutable (D) as particle diameter

increases. However, a first principles basis for the group

boundaries and/or how they change as a function of

particle interaction forces is elusive. In particular, the

group A/C demarcation is by no means absolute and in

addition to the particle size and density also shows a

dependence on a multiplicity of other factors such as the

width of the particle size distribution, the particle shape and

the surface texture and composition [5]. Recently, computa-

tional studies of gas-fluidized beds [2,3,6,7] based on the

combined approach of Particle Dynamics (PD) modeling

and computational fluid dynamics have been used to

investigate the impact of liquid-induced cohesion in these

systems.

In this paper, we first review several discrete character-

ization tools [8] for cohesive particle systems then

introduce a novel characterization tool specifically for

gas–solid flows. Our goal is to use this tool, combined

6 (2004) 160–167

Page 2: Discrete characterization of cohesion in gas–solid flows

Table 1

Simulation parameters

Parameter Soda-lime glass

Number of particles 1020

Density 2700 kg/m3

Poisson ratio 0.33

Young’s modulus 68.95 GPa

Particle radius 4.7�10�4 m

Friction coefficient 0.3

Parameter Air

Fluid density 1.23 kg/m3

Viscosity 1.80�10�5 kg/m s

K. Jain et al. / Powder Technology 146 (2004) 160–167 161

with computations of gas–solid flows, to examine the

transition from noncohesive to cohesive behavior in gas–

solid flows. Specifically, we vary gas velocities and

bridging liquid surface tension in order to explore a range

of the possible fluidization parameter space. The paper is

organized as follows. First, we describe the particle

dynamics method, including the implementation of the

various interparticle forces necessary to model liquid-

induced cohesion, and briefly explain the numerical

schemes used in our CFD model as well as how the

discrete/continuum coupling is handled. Next, character-

ization criterion based on the physical picture of liquid-

induced particle-level cohesion is reviewed for static and

sheared beds and is developed for gas–solid flows. We test

the characterization tools in the following section by

measuring both the minimum fluidization velocity as well

as the mixing rate in fluidized systems. Finally, we close by

examining the limitations and potential extensions of the

current work.

2. Discrete/continuum modeling

2.1. Particle dynamics

Particle dynamics, a discrete method of simulation, has

emerged as one of the most important tools in probing

granular flows [2,6–11]. The method is extremely general in

that Newton’s second law of motion is used to determine the

trajectories of individual particles and the time evolution of

these trajectories then determines the global flow of the

granular material. The equations that describe the particle

motion, therefore, are:

Linear motion:

mp

dvp

dt¼ � mpg þ Fn þ Ft ð1Þ

Angular motion:

Ipdxp

dt¼ Ft � R ð2Þ

where Fn and Ft are the interparticle forces—normal and

tangential, respectively—acting on the particle and are

functions of contact, drag, pressure and capillary interac-

tions. The interparticle forces for cohesionless systems are

typically determined from contact mechanics considerations,

so that in their simplest form they include normal (often,

Hertzian) [12] repulsion and some approximation of

tangential friction [13]. In the present work, normal

interactions are modeled as elastoplastic contacts after the

work of Thornton [14], while a single-parameter history-

dependent friction is used in the tangential direction [15]—

details can be found in Ref. [8]. Physical parameters of both

the particles and fluids used in the simulations can be found

in Table 1.

2.1.1. Capillary forces

Moisture is a common cause of cohesion in particle

flows. Several models based on the solution of the Young–

Laplace equation are available in the literature [6,7,16] for

the case when the degree of saturation is low enough that

discrete bridges are present at the points of solid contact

(pendular regime, see Fig. 1). The capillary force, Fc, due to

both the surface tension of the bridge fluid as well as the

pressure difference arising from neck curvature may be

expressed as

Fc ¼ 2pRcsinbsin b þ hð Þ þ pR2DPsin2b ð3Þ

where b is the half filling angle, h is the contact angle, c is

the fluid’s surface tension and DP is the pressure difference

across the air–liquid interface.

Mikami et al. [7] provide an empirical fit to the

numerical solution of the Laplace–Young equations

expressed as

Fc ¼ exp Ahþ BÞ þ C:�

A ¼ � 1:1V�0:53 ð4Þ

B ¼ � 0:34lnV� 0:96Þh2 � 0:019lnVþ 0:48�

C ¼ 0:0042lnVþ 0:0078 ð5Þ

where Fc is the normalized capillary force (Fc/2pRc); V is

the bridge volume made dimensionless by the particle

radius (R); 2h is the separation distance between the

particle made dimensionless by the particle radius (R); and

A, B and C are constants. In our simulations, the moisture

content is assumed to be sufficiently low that bridges only

form upon contact of the solid surfaces. These bridges

Page 3: Discrete characterization of cohesion in gas–solid flows

Fig. 1. Schematic of a symmetric liquid bridge.

K. Jain et al. / Powder Technology 146 (2004) 160–167162

remain in place, however, after solid contact has ceased,

until the particles reach a critical separation (rupture)

distance (hc) given by:

hc ¼ 0:62h þ 0:99ð ÞV 0:34: ð6Þ

2.1.2. Viscous forces

Dynamic formation/breakage of liquid bridges results in

a viscous force resisting motion, derived from lubrication

theory. It is essential that any liquid-induced cohesion

simulation include these effects as they may become large

relative to the capillary force as the particle velocity

increases [17]. In the limit of rigid spheres, Adams and

Perchard [18] derive the viscous force in the normal

direction (Fvn) to be

Fvn ¼ 6plRvnR

2hð7Þ

where l is the bridge fluid’s viscosity, and vn is the relative

normal velocity of the spheres. In the tangential direction

(Fv t), Lian et al. [6] suggest the use of the solution due to

Goldman et al. [19] for the viscous force between a sphere

and a planar surface

Fvt ¼8

15ln

R

2hþ 0:9588

�6plRvt

�ð8Þ

where vt is the relative tangential velocity of the spheres.

2.1.3. Drag force

Drag between the fluidizing medium (gas) and the

particle(s) couples the discrete simulation to the (contin-

uum) fluid flow and represents the primary mode of

interphase momentum transfer. The drag force not only

depends on the local fluid flow field but also on the

presence of the neighboring particles [20].

In this work, we use the drag force (Fd) suggested by Di

Felice [21]. In the formulation

Fd ¼1

2CdqgpR

2ju� vpj u� vp�e�vþ1

�ð9Þ

where u is the local gas velocity, vp is the particle

velocity, and Cd and v are functions of the particle

Reynolds Number Re ¼ 2Rqgej u�vpð Þjlg

� �

Cd ¼ 0:63þ 4:8

Re0:5

� �2ð10Þ

v ¼ 3:7� 0:65exp � 1:5� log10Reð Þ2

2

#:

"ð11Þ

2.2. Fluid dynamics

Anderson and Jackson [22] formulated continuum

equations representing mass and momentum balances from

the point Navier–Stokes and continuity equations using the

concept of local mean variables. The point variables are

averaged over regions that are large compared to the particle

diameter but small with respect to the characteristic

dimension of the complete system.

Continuity equation:

B eqg

� �Bt

þ 5d eqgu�¼ 0

�ð12Þ

Momentum equation:

B eqgu� �Bt

þ 5d eqguu� �

¼ � e5p�X

eFdð Þ

� 5d esg þ eqgg ð13Þwhere

P(eFd) is a summation of all particles’ drag force in

the control volume.

The flow domain is divided into cells of width three

times the particle diameter and discretized using a finite

volume method. An upwind scheme for convective, a

central differencing scheme for diffusive, and a fully

implicit scheme for temporal terms is used. The equations

are then solved on a staggered grid using the SIMPLE

algorithm suggested by Patankar [23]. Periodic boundary

conditions are employed on the left and right sides of the

fluidized bed and a Dirichlet boundary condition is used at

the bottom with a uniform gas inlet velocity. At the top,

Neumann boundary conditions (zero gradient) are applied

assuming the flow to be fully developed. The pressure is

fixed to a reference value at the bottom.

3. Characterization tools

The relevant variables that need to be considered in

studying gas–solid flows include (R, qs, qg, g, c, Dv, g, d,lg), where d is a characteristic length of any shearing

regions and Dv is alternatively the magnitude of the relative

velocity of the particle with respect to the fluid velocity

(Dv=u�vp) or the relative velocity between neighboring

particles (Dv=vpi�vpj). It should be noted that, in this study,

Page 4: Discrete characterization of cohesion in gas–solid flows

Fig. 2. Pressure drop versus time for different gas velocities. It is clear that

fluctuations in the calculated pressure are dramatically smaller prior to the

onset of fluidization than in fluidized systems (see inset).

Fig. 3. Standard deviation of the pressure drop versus gas velocity for dry

and wet systems. A step-change in the standard deviation of the calculated

pressure drop is seen for both wet and dry systems. This change coincides

with the onset of fluidization and is used as a fluidization criterion in this

work.

K. Jain et al. / Powder Technology 146 (2004) 160–167 163

the viscosity of the liquid bridge fluid, l, is maintained

constant as the effects of dynamic viscous forces in wet

media has been aptly explored elsewhere [17,24].

By a Buckingham Pi analysis, five dimensionless groups

are determined:

/1 ¼dR;/2 ¼

qs

qg

;/3 ¼c

qsgR2;/4 ¼

Dv2

gR;/5 ¼

clgDv

ð14Þ

The trivial dimensionless groups arising from the

density, length-scale, and relative velocity ratios do not

directly factor into studying cohesion and are ignored.

The remaining three may be thought of as combinations

of forces acting in the system: the cohesive force, the

force due to particle collisions, the weight of a particle,

and the drag force. Previous work [8] detailed the

significance of the third and fourth (using the relative

interparticle velocity) group of variables for character-

ization of wet granular systems, so they are only briefly

reviewed below.

The third group (/3), the Granular Bond Number (Bog)

[8], represents the ratio of the maximum capillary force and

the weight of a particle. This group has been shown to be

dominant in characterizing the effects of cohesion in static or

near-static systems [8]. The fourth group (/4) can be

combined with Bog to yield the Collision Number (Co)

[8,25], which represents the ratio of maximum cohesive

force and the collisional force due to Bagnold [26]. This

number has been shown to be dominant in highly sheared or

collisional granular flows where BogNCoN1.

It is the fifth group (/5) and its derivatives (i.e., a

combination of groups 4 and 5) which are examined in this

context for the first time and are of primary importance here.

This group may be easily interpreted as a ratio of the

magnitude of the drag force of the particles (Eq. (9)) to the

maximum capillary force

Fcmax¼ 2pRc: ð15Þ

This yields the Granular Capillary Number (Cag) as

Cag ¼Fc

Fd

¼ 4c

CdqgRju� vpj u� vp�e�vþ1:

� ð16Þ

4. Results and discussion

Computer simulations are run using soda lime glass

particles and air with properties as given in Table 1.

Dimensions of the bed are 30 mm (width)�510 mm

(height)�1 mm (thickness). All simulations have the same

initial condition in order to get values for comparison to

differing degrees of cohesion. The initial condition is

created by loading the particles in a randomly perturbed,

rectangular lattice. Cag is varied by changing the surface

tension of the liquid bridge fluid (as might be done with

surfactants, experimentally), while maintaining its viscosity

as a constant.

4.1. Fluidization

As a first test of the utility of the Granular Capillary

Number, we examine the onset of fluidization—the

Page 5: Discrete characterization of cohesion in gas–solid flows

Fig. 4. Increased fluidization velocity versus Granular Capillary Number.

Plotting the values of the minimum fluidization velocity versus Cag shows a

clear transition point where this value increases exponentially.

K. Jain et al. / Powder Technology 146 (2004) 160–167164

minimum fluidization velocity—in cohesive gas–solid

systems. The minimum fluidization velocity is typically

defined as the velocity at which the bed pressure drop

goes through a maximum value. A critical component of

this definition is that, while the pressure drop is ultimately

determined solely by the weight of the fluidized particles,

the value of the pressure drop can exceed this limit prior

to fluidization. In the small fluidization systems examined

here, a simpler, but equivalent, definition of the minimum

fluidization velocity is used. The approach used for

Fig. 5. Mixing progress for noncohesive/cohesive materia

determining the minimum fluidization velocity is similar

in spirit to that followed by Kafui et al. [27] which is

based on monitoring the state of the particle connectivity

network. Fig. 2 shows a plot of pressure drop versus time

for different gas velocities. For ubumf, the pressure drop

essentially remains constant and for uzumf, the pressure

drop varies with time. The amplitude and standard

deviation of the pressure disturbance also increase sharply

as the fluidization velocity becomes greater than minimum

fluidization velocity (Fig. 3). Hence, the minimum fluid-

ization velocity can be defined as the velocity at which

the fluctuations (standard deviation) of the pressure drop

go through a step change. This technique provides

reproducible results and avoids difficulties in averaging

for small systems.

We find that, using this definition of the minimum

fluidization velocity, an increase in the Cag (surface

tension) increases the velocity necessary to achieve a

fluidized system relative to that of the completely dry

(noncohesive) case. Fig. 4 shows a plot of the percentage

increase in the minimum fluidization velocity as a function

of the Cag. For values of surface tension where the Cagb1,

changes in the fluidization velocity from that of a

completely dry granular material are essentially unmeasu-

rable; however, for larger surface tensions, where the values

of CagN1, the fluidization velocities increases markedly

requiring as much as a 30% increase in umf at the highest

Cag examined.

4.2. Mixing

The mixing in gas–solid systems is often extremely rapid

compared to mixing in surface-dominated flows [9]. In this

ls. Particle clusters exist in the cohesive materials.

Page 6: Discrete characterization of cohesion in gas–solid flows

Fig. 7. Mixing rates at different surface tensions. The slope of the intensity

of segregation curves (or mixing rate) decreases markedly with increasing

liquid bridge surface tension.

K. Jain et al. / Powder Technology 146 (2004) 160–167 165

section, we examine changes of the mixing rate of

mechanically identical particles with changes in gas velocity

as well as liquid bridge surface tension.

In all cases, the system is initially completely segregated

with the right-half of the bed consisting of one type (color)

of particle and the left-half another. Typical snapshots of a

dry and wet simulation can be seen in Fig. 5. We measure

the intensity of segregation, essentially the standard devia-

tion of the local concentration, where we define the local

concentration to consist of each particle’s 10 nearest

neighbors. If the intensity of segregation is plotted as a

function of time, the value—initially at 0.5 for completely

segregated—will decrease as the system proceeds toward a

completely mixed state. Fig. 6 shows the evolution of the

intensity of segregation at several different gas velocities,

while Fig. 7 shows similar results for a fixed gas velocity

and varying liquid bridge surface tension. By fitting these

data to an exponential function, a mixing rate constant can

be extracted.

Examining both Figs. 6 and 7, it is clear that, for steady

mixing progress, higher velocities and/or lower surface

tensions result in larger (faster) mixing rates. Our definition

of the Granular Capillary Number (Cag) then suggests that

the importance of cohesion to mixing is determined by an

interplay between the capillary force and the fluid drag.

Plotting the resultant mixing rate constants as a function of

Cag, in Fig. 8, shows that this assertion is valid. That is,

mixing rates are high for small Cag and drop dramatically as

Cag increases. It is interesting to note, however, that under

certain conditions the noncohesive case (where Cag=0)

mixing rates may be lower than for small, but finite values

Fig. 6. Mixing rates at different gas velocities. As might be expected, for

steady mixing progress, the slope of the intensity of segregation curves

(which may be interpreted as the mixing rate) decreases with increasing

fluidization velocity.

of Cag. While this is most likely due to the fact that Cag=0

may be satisfied at essentially any fluidization velocity, a

detailed explanation of this observation is beyond the scope

of the current work.

5. Outlook

Here we examine the transition from free-flowing to

cohesive behavior in gas–solid flows. We have shown in

Fig. 8. Mixing rate versus Granular Capillary Number. Plotting the mixing

rate as a function of Cag causes the data from the previous trials to collapse

on one curve.

Page 7: Discrete characterization of cohesion in gas–solid flows

K. Jain et al. / Powder Technology 146 (2004) 160–167166

previous work [8] that discrete characterization tools are

extremely useful for studying the practical applications of

both static and sheared granular materials. Again, with the

introduction of the Cag, we note that changes in the

fluidization onset are aptly captured by these types of tools.

While this simple approach works surprisingly well for

predicting the transition point for the minimum fluidization

velocity, at present it does not address the nature of the

change—although there is clearly an exponential change in

its value. Moreover, the mixing behavior is surprisingly

sensitive to these changes as well and, in fact, the transition

seems dramatically sharper for this unit operation. Ulti-

mately, this simple characterization tool may serve as a

useful a priori test of the fluidization character to be

expected in a wet gas–solid system.

6. Nomenclature

Symbols

A, B, C constants in the empirical model of the capillary

rce

fo

Bog Granular Bond Number

Cd drag coefficient

Cag Granular Capillary Number

F interparticle force, N

Fc capillary force, N

Fcmaxmaximum capillary force, N

Fd drag force, N

Fv viscous force, N

Fc normalized capillary force

g gravity acceleration, m/s2

h half separation distance between particles, m

h dimensionless half separation distance between

rticles

pa

hc dimensionless critical rupture distance

Ip particle moment of inertia, kg m2/s

mp particle mass, kg

DP pressure difference across the air–liquid interface, Pa

r1 bridge meridional radius of curvature, m

r2 bridge neck radius, m

R particle radius, m

Re Reynolds number

u gas velocity, m/s

Dv relative velocity of the spheres, m/s

vp particle velocity, m/s

V dimensionless bridge volume

Greek letters

b half filling angle

c surface tension, kg/s2

d characteristic length, m

e void fraction

h contact angle

l liquid viscosity, kg/m s

lg gas viscosity, kg/m s

qg gas density, kg/m3

qs particle density, kg/m3

sg gas viscous stress tensor, Pa

/1, /2, /3, /4, /5 dimensionless groups

xp particle angular velocity, 1/s

v a function of Reynolds number

Subscripts

n normal component

t tangential component

Acknowledgments

The authors would like to acknowledge the support of the

Petroleum Research Fund as administered by the American

Chemical Society, the Department of Energy (DOE-NETL),

and the National Science Foundation (CTS-0105688).

References

[1] C.Y. Wen, L.H. Chen, Flow modeling concepts of fluidized beds, in:

N.P. Cheremisinoff, R. Gupta (Eds.), The Handbook of Fluids in

Motion, 1983, pp. 665–714.

[2] M.J. Rhodes, X.S. Wang, M. Nguyen, P. Stewart, K. Liffman, Onset

of cohesive behavior in gas fluidized beds: a numerical study using

DEM simulation, Chem. Eng. Sci. 56 (2001) 4433–4438.

[3] L.J. McLaughlin, M.J. Rhodes, Prediction of fluidized bed behavior in

the presence of liquid bridges, Powder Technol. 114 (2001) 213–223.

[4] D. Geldart, Types of gas fluidization, Powder Technol. 7 (1973)

285–292.

[5] J.L.R. Orband, D. Geldart, Direct measurement of powder cohesion

using a torsional device, Powder Technol. 92 (1997) 25–33.

[6] G. Lian, C. Thornton, M.J. Adams, Discrete particle simulation of

agglomerate impact coalescence, Chem. Eng. Sci. 53 (1998)

3381–3391.

[7] T. Mikami, H. Kamiya, M. Horio, Numerical simulation of cohesive

powder in a fluidized bed, Chem. Eng. Sci. 53 (1998) 1927–1940.

[8] S.T. Nase, W.L. Vargas, A.A. Abatan, J.J. McCarthy, Discrete

characterization tools for cohesive granular material, Powder Technol.

116 (2001) 214–223.

[9] J.J. McCarthy, D.V. Khakhar, J.M. Ottino, Computational studies of

granular mixing, Powder Technol. 109 (2000) 72–82.

[10] B.P.B. Hoomans, J.A.M. Kuipers, W.P.M. van Swaaij, Granular

dynamics simulation of segregation phenomena in bubbling gas-

fluidized beds, Powder Technol. 109 (2000) 41–48.

[11] J.A.M. Kuipers, K.J. van Duin, F.P.H. van Beckum, W.P.M. van

Swaaij, A numerical model for gas-fluidized beds, Chem. Eng. Sci. 47

(1992) 1913.

[12] H. Hertz, Ueber die berhrungfester elastischer korper, J. Reine Angew.

Math. 92 (1881) 1–15.

[13] R.D. Mindlin, Compliance of elastic bodies in contact, J. Appl. Mech.

16 (1949) 256–270.

[14] C. Thornton, Coefficient of restitution for collinear collisions of elasti-

perfectly plastic spheres, J. Appl. Mech. 64 (1997) 383–386.

[15] P.A. Cundall, O.D.L. Strack, A discrete numerical model for granular

assemblies, Geotechnique 29 (1979) 47–65.

[16] R.A. Fisher, On the capillary forces in an ideal soil, J. Agric. Sci. 16

(1926) 491–505.

[17] B.J. Ennis, J. Li, G.I. Tardos, R. Pfeffer, The influence of viscosity on

the strength of an axially strained pendular liquid bridge, Chem. Eng.

Sci. 45 (1990) 3071–3088.

Page 8: Discrete characterization of cohesion in gas–solid flows

K. Jain et al. / Powder Technology 146 (2004) 160–167 167

[18] M.J. Adams, V. Perchard, The cohesive forces between particles with

interstitial fluid, Inst. Chem. Eng. Symp. 91 (1985) 147–160.

[19] A.J. Goldman, R.G. Cox, H. Brenner, Slow viscous motion of a

sphere parallel to a plan wall: I. Motion through a quiescent fluid,

Chem. Eng. Sci. 22 (1967) 637–651.

[20] L.S. Fan, C. Zhu, Principles of Gas–Solid Flows, Cambridge Univ.

Press, New York, 1998.

[21] R. Di Felice, The voidage function for fluid–particle interaction

systems, Int. J. Multiph. Flow 20 (1994) 153–159.

[22] T.B. Anderson, R. Jackson, A fluid mechanical description of

fluidized beds, Ind. Eng. Chem. Fundam. 6 (1967) 527–539.

[23] S.V. Patankar, Numerical Heat and Transfer and Fluid Flow, Hemi-

sphere, New York, 1980.

[24] B.J. Ennis, G.I. Tardos, R. Pfeffer, A microlevel-based character-

ization of granulation phenomena, Powder Technol. 65 (1991)

257–272.

[25] H. Li, J.J McCarthy, Controlling cohesive particle mixing and

segregation, Phys. Rev. Lett. 90 (2003) 184301.

[26] R.A. Bagnold, Experiments on a gravity-free dispersion of large solid

spheres in a Newtonian fluid under shear, Proc. R. Soc. 225 (1954)

4–63.

[27] K.D. Kafui, C. Thornton, M.J. Adams, Discrete particle-continuum

fluid modelling for gas–solid fluidised beds, Chem. Eng. Sci. 57

(2002) 2395–2410.