165
Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity Measurements Compared with Discrete Element and Continuum Models by Kamyar Hashemnia A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Mechanical and Industrial Engineering University of Toronto © Copyright by Kamyar Hashemnia 2015

Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

Vibrationally-Fluidized Granular Flows: Impact and Bulk

Velocity Measurements Compared with Discrete Element

and Continuum Models

by

Kamyar Hashemnia

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Mechanical and Industrial Engineering University of Toronto

© Copyright by Kamyar Hashemnia 2015

Page 2: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

ii

Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity

Measurements Compared with Discrete Element and Continuum

Models

Kamyar Hashemnia

Doctor of Philosophy

Mechanical and Industrial Engineering

University of Toronto

2015

Abstract

A new laser displacement probe was developed to measure the impact velocities of particles

within vibrationally-fluidized beds. The sensor output was also used to measure bulk flow

velocity along the probe window and to provide a measure of the media packing. The

displacement signals from the laser sensors were analyzed to obtain the probability distribution

functions of the impact velocity of the particles. The impact velocity was affected by the

orientation of the laser probe relative to the bulk flow velocity, and the density and elastic

properties of the granular media. The impact velocities of the particles were largely independent

of their bulk flow speed and packing density.

Both the local impact and bulk flow velocities within a tub vibratory finisher were predicted

using discrete element modelling (DEM) and compared to the measured values for spherical steel

media. It was observed that the impact and bulk flow velocities were relatively insensitive to

uncertainties in the contact coefficients of friction and restitution. It was concluded that the

Page 3: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

iii

predicted impact and bulk flow velocities were dependent on the number of layers in the model.

Consequently, the final DE model mimicked the key aspects of the experimental setup, including

the submerged laser sensor. The DE method predictions of both impact velocity and bulk flow

velocity were in reasonable agreement with the experimental measurements, with maximum

differences of 20% and 30%, respectively.

Discrete element modeling of granular flows is effective, but requires large numerical models.

In an effort to reduce computational effort, this work presents a finite element (FE) continuum

model of a vibrationally-fluidized granular flow. The constitutive equations governing the

continuum model were calibrated using the discrete element method (DEM). The bulk flow

behavior of the equivalent continuum media was then studied using both Lagrangian and

Eulerian FE formulations. The bulk flow velocities predicted by the Lagrangian approach were

in good agreement with those obtained using DEM simulations over a wide range of tub wall

amplitudes. The local impact velocity distribution predicted by the DEM was also compared to

the continuum model using the shear rate as a measure of the granular temperature.

Page 4: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

iv

Acknowledgments

I would like to express my great appreciation to my supervisor, Professor Jan K. Spelt, for

providing me with his continuous guidance, enthusiasm and encouragement to assist me in

conducting a successful research.

I would like to thank my Ph.D. committee, Professor Markus Bussmann and Professor Nasser

Ashgriz for their valuable comments and suggestions offered during my annual exams and

reports. I also appreciate the thesis reviews and helpful comments provided by Professors Carl

Wassgren and Murray Grabinsky. I would like to acknowledge the Department of Mechanical

and Industrial Engineering of the University of Toronto for preparing a wonderful atmosphere to

pursue my Ph.D. I am also grateful of the financial support from Natural Sciences and

Engineering Research Council (NSERC) of Canada for providing the necessary means to fund

this project.

I would like to extend my acknowledge to the other members of the Materials and Process

Mechanics Laboratory, Amirhossein Mohajerani, Reza Haj Mohammad Jafar, Hooman Nouraei,

Amir Nourani, Saeid Akbari, Kavin Kowsari, Neda Tamannaee for not only their technical

assistance but also for creating a very good atmosphere to work and share ideas. I would like to

appreciate Amir Nourani for his support in learning Abaqus during the final year of my PhD. I

should also appreciate all my valuable friends who supported me during my Ph.D. experience,

especially: Mohammad Hossein Amini, Mohammad Ali Amini, Roshanak Banan, Ali Ebrahimi,

Javad Esmaeilpanah, Pakeeza Hafeez, Mohammad Reza Kholghy, Masih Mahmoodi, Esmaeil

Safaei, Ashkan Aryaee, Saber Jafarpur and Azadeh Hushmandi.

I would also like to thank my father and mother for their help and support as well as my sister

and brother. Their unconditional support and love has been extremely helpful during each stage

of my life. If they had not supported me emotionally and spiritually, I know that I could not have

reached any of the achievements in my life.

Page 5: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

v

Table of Contents

Acknowledgments .......................................................................................................................... iv

Table of Contents ............................................................................................................................ v

List of Tables ................................................................................................................................. ix

List of Figures ................................................................................................................................ xi

List of Appendices ..................................................................................................................... xviii

Chapter 1 ......................................................................................................................................... 1

1 Introduction ................................................................................................................................ 1

1.1 Overview ............................................................................................................................. 1

1.2 Literature review ................................................................................................................. 2

1.3 Objectives ........................................................................................................................... 5

1.3.1 Objective 1 .............................................................................................................. 5

1.3.2 Objective 2 .............................................................................................................. 6

1.3.3 Objective 3 .............................................................................................................. 6

1.4 Thesis outline ...................................................................................................................... 6

1.5 References ........................................................................................................................... 7

Chapter 2 ....................................................................................................................................... 10

2 Development of a Laser Displacement Probe to Measure Particle Impact Velocities in

Vibrationally-Fluidized Granular Flows .................................................................................. 10

2.1 Introduction ....................................................................................................................... 10

2.2 Experiments ...................................................................................................................... 12

2.2.1 Measurement approach ......................................................................................... 12

2.2.2 Tub motion characterization ................................................................................. 14

2.2.3 Displacement sensor characterization ................................................................... 17

Page 6: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

vi

2.2.4 Test procedures in the vibratory finisher .............................................................. 18

2.3 Signal analysis .................................................................................................................. 24

2.3.1 Ball motion in the vibratory finisher and the laser sensor displacement signals .. 24

2.3.2 Impact velocity calculation ................................................................................... 30

2.3.3 Double impact hypothesis ..................................................................................... 31

2.3.4 Error due to tangential bulk flow velocity ............................................................ 33

2.4 Probe evaluation in the tub finisher .................................................................................. 33

2.5 Conclusions ....................................................................................................................... 43

2.6 References ......................................................................................................................... 44

Chapter 3 ....................................................................................................................................... 47

3 Particle Impact Velocities in a Vibrationally-Fluidized Granular Flow: Measurements and

Discrete Element Predictions ................................................................................................... 47

3.1 Introduction ....................................................................................................................... 47

3.2 Discrete element modeling ............................................................................................... 49

3.2.1 Contact models and properties .............................................................................. 49

3.2.2 Reduced shear modulus ........................................................................................ 54

3.3 Sensitivity of DEM impact velocities to contact properties ............................................. 58

3.4 Measurement of contact properties ................................................................................... 61

3.5 Results and discussion ...................................................................................................... 66

3.5.1 Effect of model width ........................................................................................... 66

3.5.2 Comparisons with immersed laser-probe velocity measurements ........................ 70

3.6 Conclusions ....................................................................................................................... 75

3.7 References ......................................................................................................................... 76

Chapter 4 ....................................................................................................................................... 79

4 Finite Element Continuum Modeling of Vibrationally-Fluidized Granular Flows ................. 79

4.1 Introduction ....................................................................................................................... 80

Page 7: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

vii

4.2 Continuum constitutive equations ..................................................................................... 84

4.3 Determining the continuum model parameters ................................................................. 88

4.3.1 Coefficient of internal friction and volume fraction ............................................. 88

4.3.2 Media equivalent Young's modulus and media-wall effective contact stiffness .. 92

4.3.3 Media-wall equivalent coefficient of friction ....................................................... 97

4.4 Finite element implementation of the continuum model ................................................ 102

4.4.1 Lagrangian FE continuum model ........................................................................ 102

4.4.2 Eulerian FE continuum model ............................................................................ 107

4.5 Sensitivity to uncertainty in continuum model parameters ............................................. 110

4.6 Comparison of FEA and DEM bulk flow velocities ....................................................... 112

4.7 Comparison of FEA and DEM local impact velocities .................................................. 115

4.8 Conclusions ..................................................................................................................... 118

4.9 References ....................................................................................................................... 118

Chapter 5 ..................................................................................................................................... 122

5 Conclusions ............................................................................................................................ 122

5.1 Experimental measurements ........................................................................................... 122

5.2 Discrete element modeling ............................................................................................. 123

5.3 Continuum modeling ...................................................................................................... 124

6 Future work ............................................................................................................................ 126

Appendix A: Analytical sensitivity study in Chapter 3 .............................................................. 127

Appendix B: MATLAB Codes of Chapter 2 .............................................................................. 132

Appendix B.1. Fast Fourier transformation (FFT) of the tub motion based on the

measurements .................................................................................................................. 132

Appendix B.2. Acceleration, velocity and position of the vibratory finisher center of

gravity ............................................................................................................................. 133

Appendix B.3. Path of the top points of the vibratory finisher (A and B in Fig. 2.2a) .......... 136

Page 8: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

viii

Appendix B.4. The impact velocity calculation based on the laser displacement sensor

measurements .................................................................................................................. 137

Appendix C: MATLAB Codes of Chapter 3 .............................................................................. 139

Appendix C.1. The calculation of the in-plane components of the impact velocity based on

DEM simulation results .................................................................................................. 139

Appendix C.2. The calculation of the out- of-plane component of the impact velocity

based on DEM simulation results ................................................................................... 140

Appendix C.3. The calculation of coefficient of friction between spherical steel balls

based on the linear tribometer measurements ................................................................. 141

Appendix D: MATLAB Codes of Chapter 4 .............................................................................. 143

Appendix D.1. The calculation of local stress tensor, pressure and the coefficient of

internal friction based on the DEM data to calibrate the equivalent properties needed

in the continuum model .................................................................................................. 143

Appendix D.2. The calculation of local shear rate based on the DEM data to calibrate the

equivalent properties needed in the continuum model ................................................... 145

Page 9: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

ix

List of Tables

Table 2.1: Physical and mechanical properties of steel and porcelain media (±

indicates standard deviation, 10 measurements) 15

Table 2.2: Bulk flow velocity parallel to window, particle passage frequency, packing

parameter and impact velocity of the steel and porcelain media for locations H, M and

L in different directions (number of data points in brackets). 95% confidence

intervals on the mean values based on student t distribution. 35

Table 3.1: Material properties used in the DEM simulations. 52

Table 3.2: Measured tub center of gravity vibration components used in DEM. 52

Table 3.3: Change in the x and z components of the impact velocity (denoted IV) and

bulk flow velocity (denoted BFV) at the points H and RU (Fig. 3.2) due to a shear

modulus reduction of 1, 2 and 3 orders of magnitude (denoted G-1, G-2, G-3,

respectively) and their corresponding errors (Er). Velocities in mm/s. Root mean

square error (RMS Er) is for all components and locations of either IV or BFV. Time

to run DEM simulation given in last row. 57

Table 3.4: Percentage uncertainty in the linear and angular rebound velocities due to

10% uncertainty in each of the contact parameters. 59

Table 3.5: Percentage uncertainty (U) in the single-layer DEM predictions of the

impact and bulk flow velocities due 10% uncertainty in the contact parameters

normalized by the following mean velocities from the DEM: IVx =75 mm/s, IVz =95

mm/s, BFVx =14 mm/s, BFVz =20 mm/s. p-p particle-particle, p-w particle-wall.

Location H of Fig. 3.2. 60

Table 3.6: The coefficient of friction (µ) between the steel ball of finishing media and

the sample of the polyurethane tub wall as a function of sliding speed. 61

Page 10: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

x

Table 3.7: The measured coefficients of friction and restitution for different pairs of

materials (mean ±95% confidence interval). 65

Table 3.8: The average slope and ±95% confidence intervals of the free surface in

models with various numbers of particle layers in the width direction, y, between the

glass partitions. 67

Table 3.9: Comparison of the predicted and measured impact velocities (±95%

confidence intervals based on at least 200 data points from the DEM and at least 1,000

laser-probe velocity measurements; mm/s) and bulk flow velocities (mean value;

mm/s) at locations H and M in different directions (Figs. 3.2 and 3.8). The bracketed

bulk flow velocity at location M-rightwards refers to the DEM absolute oscillating

velocity. 73

Table 4.1: Material properties used in the DEM simulations ‎[19]. 89

Table 4.2: Measured tub center of gravity vibration components used in DEM ‎[19]. 89

Table 4.3: Material and contact properties used in the FE simulations. The media

effective‎ Young’s‎ modulus,‎ and‎ the contact parameters were evaluated over the

indicated ranges. The values of the media Young's modulus and the media-wall

contact stiffness in parentheses were calculated at the mean volume fraction φmean; i.e.

the best estimate. 105

Table 4.4: Sensitivity of the Lagrangian FE model mean bulk flow velocity (MFV)

and free-surface angle to changes in the effective media-wall contact stiffness, media

equivalent Young's modulus, and media-wall coefficient of friction. 112

Page 11: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xi

List of Figures

Fig. 1.1: Photo of the tub vibratory finisher 2

Fig. 2.1: Schematic of laser displacement sensor triangulation principle. 14

Fig. 2.2: Schematic of the tub vibratory finisher showing the bulk flow circulation

direction: (a) side view, (b) top view. The probe is shown at location H. The curved

arrows show the flow streamlines around the elliptical tube. Dimensions are in

millimeters and are drawn to scale. 16

Fig. 2.3: Displacement of a steel ball falling onto a glass plate acquired by the laser

displacement sensor (LK-G 82). The time step of data acquisition was 50 s (20 kHz

sampling rate). The arrows show the first, second and third impact points. 18

Fig. 2.4: (a) Schematic of the test apparatus used to measure the impact velocity of the

balls in the vertical direction in the vibratory finisher. Sensor, tub and balls are drawn

to scale. Note that the steel rod supporting the laser sensor was mounted directly to a

brick wall and was thus isolated from vibration. (b) enlarged portion of (a). (c)

enlarged view of the probe used to measure the impact velocity in the horizontal

(upstream or downstream) direction. 21

Fig. 2.5: The vibration of the window in Fig. 2.4c and the simultaneous vibration of

the shielding tube. The solid line represents the window motion while the dotted line

represents the shielding tube motion. 22

Fig. 2.6: Schematic of the vibrating tub showing the bulk flow pattern and the three

locations of the probe as it was when measuring toward the right. The x-y distances

correspond to the fixed locations in the flow used in all measurement directions.

Dimensions are in millimeters. 23

Fig. 2.7: (a) Inadequate laser beam reflection from the ball surfaces with excessively

steep slopes; (b) Laser beam reflecting from a non-diametral plane; (c) The range for

the laser sensor to accurately detect the position of the ball in a diametral plane (θ1 ≠ 25

Page 12: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xii

θ2).

Fig. 2.8: (a) Displacement signal (normal to the glass window) generated by the

passage of five balls parallel to the glass window in the rightward direction at location

H. Arrows show readings when laser passed between balls. (b) a magnified portion of

(a) corresponding to the passage of a single ball; (c) a magnified portion of (b); (d) a

magnified portion of Fig. 2.3 which shows the drop test results. The vertical arrows in

(c) and (d) show the points corresponding to contact between the window and ball. 27

Fig. 2.9: Chord length a used to calculate the bulk flow velocity of the balls passing

the sensor in a diametral plane. 28

Fig. 2.10: Predicted positions of the window (dashed line) and ball showing first and

second collisions giving rise to the apparent rounding of the impact peaks. Modeled

using data from location H in the steel media in the rightward direction (Fig. 2.6). 32

Fig. 2.11: Schematic of a ball having both normal (impact) and tangential (bulk flow)

velocity components relative to the window. 33

Fig. 2.12: Raw displacement signals of balls moving in the vibratory finisher at

location H: (a) rightward, (b) downward, (c) leftward (Fig. 2.6). The signal from the

window is shown in the narrow band at the top of each graph. 38

Fig. 2.13: The probability density distributions of the impact velocity of steel balls in

the vibratory finisher (a) at location H in the three orthogonal directions and (b) at

locations M and L in the directions indicated (Fig. 2.6). The types of distributions are

shown in the figure. 39

Fig. 2.14: The probability density distributions of the impact velocity of porcelain

balls in the vibratory finisher at locations H and M in the directions shown (Fig. 2.6).

All the distributions are Gaussian. 40

Fig. 2.15: The mean values of bulk flow velocity, impact velocity and packing

parameter (normalized by the largest values for each media) in different locations and 42

Page 13: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xiii

directions (Fig. 2.6) with: (a) steel, and (b) porcelain media.

Fig. 3.1: (a) Photo of the tub vibratory finisher including the laser probe. Schematic

of the tub vibratory finisher showing the bulk flow circulation direction (drawn to

scale): (b) side view, (c) plan view. The probe is shown at location H. The curved

arrows show the flow streamlines around the elliptical outer tube of the laser velocity

probe. Dimensions in mm ‎[9]. 53

Fig. 3.2: Snapshot of a single particle-layer simulation and the three measurement

points located at (x,z): RU (200, 300), H (100, 300), M (0, 280) (mm). At each

location, the small and large boxes were for impact velocity and bulk flow velocity

calculation, respectively. Figure drawn to scale. Straight line approximation to free-

surface slope established at steady state. Arrow shows average bulk flow direction. 56

Fig. 3.3: Schematic showing the lower steel ball sliding against the upper steel ball

that was constrained in the y-direction to measure vertical displacement depth only. T

is the measured horizontal force. y=0 taken to be the minimum measured depth at the

apex of the upper ball. 62

Fig. 3.4: Tribometer measurements of the horizontal sliding force vs. (a) time and (b)

angle (θ) for steel balls moving against each other. The solid straight lines in (a) and

(b) show the analytical prediction. 63

Fig. 3.5: The drop test apparatus used to measure the coefficient of restitution of: (a)

the steel ball colliding with the polyurethane surface, (b) two colliding balls (side and

plan views). 65

Fig. 3.6: Predicted impact velocities in: (a) horizontal (x) direction, (b) vertical (z)

direction, and (c) transverse (y) direction, in the DEM simulations with different

numbers of layers. 69

Fig. 3.7: Predicted bulk flow velocities in (a) horizontal (x) direction and (b) vertical

(z) direction, in the DEM simulations with different numbers of layers. 70

Page 14: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xiv

Fig. 3.8: Snapshot of the simulation constructed using the plane of symmetry through

the elliptical laser velocity probe used in [9] located at point M (as if measuring

velocities of particles approaching from the right). (a) side view, (b) plan view, (c)

side view of probe and the measurement bins on the probe surface oriented with the

laser window to the right. The large bin with solid boundaries and the small bin with

dotted boundaries were used to obtain the impact and bulk flow velocities,

respectively. (d) Side view of a portion of the probe surface showing the particle bulk

flow with the measurement bin in the downward directions. In (c) and (d), large

arrows show the bulk flow direction, dashed smaller ones represent the laser beam. 71

Fig. 3.9: Probability density distributions of impact velocities predicted by the DEM

and measured in the experiments of ref. [9]. All the distributions were fitted with log-

normal functions. Locations defined in Figs. 3.2 and 3.8. 74

Fig. 4.1: (a) Photograph of the tub vibratory finisher including the laser probe used to

measure the media impact velocities in ref. [9], and two glass partitions within the tub.

(b) Schematic of the tub vibratory finisher showing the bulk flow circulation direction

from the side view. Dimensions in mm ‎[19]. 83

Fig. 4.2: Variation of (a) the equivalent coefficient of internal friction ‎[29] and (b) the

volume fraction with the inertial number seen in simple shear test ‎[28]. 87

Fig. 4.3: The average streamlines of particles moving in a counter-clockwise flow in

the first layer of the 4-layer DE model during a 1 s interval. 89

Fig. 4.4: Measurement bin locations, H1, H2, L1 and L2 used to calculate the shear

rates and stress tensors from the DEM. Three in-plane bins at each location and a forth

bin behind the main bin. 91

Fig. 4.5: DEM predictions of the coefficient of internal friction as a function of: (a) the

inertial number, (b) the local pressure and (c) the shear rate; (d) Local volume fraction

vs. the inertial number. The solid and dashed lines are the average and the 95%

confidence bounds. 92

Page 15: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xv

Fig. 4.6: The equivalent Young's modulus of the vibrationally-fludized bulk flow, Eb,

vs. the solid fraction, φ, based on Eqs. (4.9-4.11) and the DEM. 95

Fig. 4.7: Seven measurement bins used to study the media solid fraction variation with

distance from the wall when: (a) wall moves toward media leading to maximum

compression, (b) wall moves away from media leading to maximum decompression of

the media and some separation between the media and the wall. 96

Fig. 4.8: Average solid fraction, φ, variation with distance from the wall using data

from the measurement bins of Fig. 4.7. 96

Fig. 4.9: Eight measurement locations used to determine average shear and normal

forces acting on the media along the tub wall. 97

Fig. 4.10: Probability density distributions of the ratio of the shear to normal forces

applied to the media by the wall (Ft/Fn) at the 8 locations of Fig. 4.9 for: (a) all data

(columns ordered from left to right as positions 1-8), and (b) for Ft/Fn <0.8. Based on

all particle-wall impact events recorded by the DEM during 47 vibration cycles 99

Fig. 4.11: Shear force vs. normal force at location 2 (Fig. 4.9) for all wall impacts over

47 vibration cycles. The best-fit lines represent the mean effective coefficient of

friction (µmw=0.25) with the correlation coefficient, and the measured COF for sliding

of a single steel ball against the polyurethane wall material (µmw=1.8). The line for

µmw=0.7 is also shown since it will be used in the continuum modelling. 100

Fig. 4.12: The average effective coefficient of friction between the fluidized media

and the tub wall at the 8 locations shown in Fig. 4.9 based on all impacts over 47

vibration cycles for the 4-layer and 8-layer DEM. The horizontal lines show the

average and 95% confidence bounds. 101

Fig. 4.13: Grand average shear force acting on the media by the wall over each 1/12 of

a cycle over 235 vibration cycles. Shear forces recorded over the high wall (HW;

locations 1-3), low wall (LW; location 4-6), and bottom wall (BW; locations 7-8) as

shown in Fig. 4.9. Positive shear force produced a counter-clockwise torque on the 102

Page 16: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xvi

media.

Fig. 4.14: The granular media domain meshed using Lagrangian elements: Free

triangular elements of sizes (a) 2 cm and (b) 1 cm; free quadrilateral elements of sizes

(c) 2 cm and (d) 1 cm; tub wall meshed using free triangular elements of size 2 cm. 104

Fig. 4.15: The bulk flow velocity distribution obtained from the Lagrangian analysis:

Free triangular elements of sizes (a) 2 cm and (b) 1 cm; free quad elements of sizes (c)

2 cm and (d) 1 cm. The simulation parameters: K=30.2 MPa/m, Eb=257 kPa and

µmw=0.3. Nodal velocity scale shown in lower right. Some arrows omitted for clarity.

Mean bulk flow velocity (MFV) across the dark horizontal lines. 106

Fig. 4.16: Material deformations in the Lagrangian and Eulerian analyses ‎[45]. 108

Fig. 4.17: The media meshed by Eulerian elements of sizes (a) 1 cm and (b) 0.5 cm.

The tub wall was meshed with Lagrangian elements. 109

Fig. 4.18: Bulk flow velocity distributions obtained from the Eulerian analysis for

rectangular elements of sizes: (a) 1 cm and (b) 0.5 cm. Nodal velocity scale shown in

upper left. Some arrows omitted for clarity. 110

Fig. 4.19: (a) mean bulk flow velocity and (b) inclination of free surface as a function

of tub amplitude expressed as the fraction of the actual tub amplitude, A. Predictions

of the DEM and of the Lagrangian FE continuum model that used the optimized

media-wall effective coefficient of friction, µmw =0.7. FE predictions with µmw=0.5

included for comparison. 114

Fig. 4.20: Distribution of bulk flow velocity perpendicular to the horizontal line

defined in Fig. 4.15 in the 4-layer DE and the Lagrangian FE using the optimized

µmw=0.7 (K=30.2 MPa/m, Eb=295 kPa). Zero distance corresponds to the center of

circulation and 160 mm is at the tub wall. 115

Fig. 4.21: (a) Average particle impact velocity, Vimp, and its components vs. the shear

rate, , at points H1, L1, L2, H2 (ordered from left to right) in the 4-layer DEM (Fig. 117

Page 17: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xvii

4.4). (b) Shear rate in FE and DE simulations at the same points.

Fig. 3-A.1: Collision of two disks having arbitrary initial linear and angular velocities:

(a) velocities after collision, and (b) impulses during impact. 127

Page 18: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

xviii

List of Appendices

Appendix A: Analytical sensitivity study in chapter 3 127

Appendix B: MATLAB Codes of Chapter 2 132

Appendix C: MATLAB Codes of Chapter 3 139

Appendix D: MATLAB Codes of Chapter 4 143

Page 19: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

1

Chapter 1

1 Introduction

1.1 Overview

Vibratory finishing is widely used to deburr, polish, burnish, harden and clean metal, ceramic

and plastic parts. In a tub vibratory finisher, a container filled with granular media is oscillated

by an eccentric rotating shaft so that it develops a vibrationally-fluidized circulatory bulk flow of

the media that is largely two-dimensional (Fig. 1.1). The media have both a large-scale bulk

flow velocity and a much smaller-scale local impact velocity. Therefore, work-pieces that are

entrained in the flowing media are subjected to the repetitive, high-frequency impacts of the

surrounding particles. The granular media can be made of many materials of various shapes

such as spheres, cylinders or pyramids, with either a smooth or abrasive texture.

Within a vibratory finisher, the erosive wear and plastic deformation of workpieces are largely

affected by the velocity, frequency, and direction of the impacts with particles. High impact

velocities can cause fracture and fragmentation while low impact velocities can make the process

less efficient. The impact velocity is also closely related to the breakage of granular materials in

many processes such as drug tablet coating within rotating drums, bulk materials handling, food

processing, and particle attrition in vibratory finishing. Large particle impact velocities may also

result in erosion of machine components in processes such as vibratory sieving and mixing.

Page 20: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

2

Fig. 1.1: Photo of a tub vibratory finisher

1.2 Literature review

The flow of granular materials has been studied in non-fluidized beds such as in hopper

discharge [1, 2], rotating drums, conveyers, chutes and mixers [3, 4], and in fluidized beds such

as in vibratory finishers and vibrating sieves [5-8]. A granular media becomes fluidized as the

vertical acceleration amplitude exceeds gravity, resulting in a marked decrease in the contact

pressure [8].

Measurement of local quantities like the impact velocity between vibrating particles in a

vibrationally-fluidized granular media is challenging, because of the relatively small scale of the

motion and difficulties in designing probes capable of measuring local quantities without

disturbing the media. A number of purely experimental studies have been done on the local

behavior of granular media [9-11]. Yabuki et al. [11] used a 3 mm diameter “floating”‎ pad‎

Page 21: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

3

connected to four resistive force sensors to measure both normal and tangential impact forces.

They found that forces normal to the sensor were approximately 10 times greater than the

tangential forces in both the wet and dry conditions. In the wet condition, a small percentage of

the media had a tangential to normal force ratio approximately equal to the measured dynamic

coefficient of friction between media, indicating that sliding had occurred. The force signals also

showed that the packing of the media in the vibrating bowl finisher was relatively loose, with

many gaps that were akin to bubbles.

The variation of the impact forces in steel and porcelain media in a tub vibratory finisher was

investigated in [10, 11] using a unidirectional high-speed impact force sensor. Local impact

velocities were inferred from impact forces, but this indirect measurement of impact velocity was

dependent on a system calibration and required a relatively large sensor that disrupted the bulk

flow in the finisher [10]. These impact force measurements were sensitive to the stiffness of the

sensor, and hence were specific to the experimental setup. This points to a key advantage of

measuring impact velocities directly; the impact velocities of particles will be unaffected by the

sensor compliance. Comparing the results of investigations in tub and bowl-type finishers, and

in finishers operating with different amplitudes [10-12], it can be concluded that the major

characteristics of the local behavior of the particles are largely independent of the type of

vibratory finisher producing the granular flow. For example, the vibrationally-fluidized flows

tend to be loosely packed and collisions with workpiece surfaces are dominated by the normal

component of impact.

Most research in the field of flowing granular media has focused on the bulk flow of granular

media inside different machines such as hopper discharge [1, 2], rotating drums [3, 4], and

vibratory finishers [7, 8]. The discrete element method (DEM) is a numerical technique that is

often used to solve problems involving transient dynamics of systems comprising a large number

of moving bodies that interact with each other. Generally, discrete element modeling (DEM)

simulations [13, 14] give reasonable predictions of bulk flow velocity and volume fraction in

both fluidized and non-fluidized flows [15-17]. For instance, the behavior of the convection

cells produced by a spherical media within a vertically vibrated container has been studied using

DEM and the results were validated by the experimental data [16]. As another example, the bulk

flow velocity distribution of glass beads in a vertical axis mixer with rotating flat blades was

Page 22: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

4

obtained in DEM simulations and validated by the positron emission particle tracking (PEPT)

measurements [17].

The local behavior of non-fluidized granular flows has also been investigated experimentally and

numerically [2, 18-20], while some studies have examined vibrationally-fluidized flows [10-12,

21-23]. For instance, DEM was used to obtain the planar distributions of the collision velocity

and frequency of the granular media inside a horizontal rotating drum [2]. In the same geometry,

DEM was used to predict the effects of particle stiffness, size and coefficient of restitution on the

solid fraction, collision frequency, and impact velocity inside the granular media [18]. A

vibrated bubbling fluidized bed was simulated using DEM to give the instantaneous particle

velocity and local volume fraction [23]. However, only in ref. [19] were DEM predictions of

local behavior compared with experimental measurements. In that case, the local volume

fraction, granular temperature and bulk flow velocity in the surface layer of a shear flow were

measured using digital particle tracking velocimetry and compared with DEM predictions [19].

Previous efforts to model granular flows using a continuum approach have focused on the bulk

flow behavior of the non-fluidized granular media such as in the flow down inclined chutes [24-

27], plane shear flow [28, 29], flow in rotating drums [27, 28], flow in annular shear cells [24,

27, 28], and silo discharge flows [25, 27, 28]. In these cases of non-fluidized flows, constitutive

equations were defined to describe the bulk flow of the granular materials under quasi-static and

moderate flow (liquid-like) regions [24, 26, 30]. To‎ the‎ author’s‎ knowledge,‎ no‎ papers‎ have‎

been published on continuum modelling of vibrationally-fluidized granular flows. The

continuum models that have been used for quasi-static and moderate flow include different

elasto-plastic formulations of the equivalent continuum media [24, 30, 32], and visco-plastic

formulations considering only the plastic behavior of the equivalent media under time-varying

shear deformation [26-27]. In many cases, DEM was used to obtain the equivalent properties

needed to model the granular flows as a continuum [25, 27-29, 31, 32]. For example, the

equivalent stress tensor, pressure and shear rate at different points of a flowing granular bed were

obtained using DEM, and were then used to determine the equivalent continuum media elastic

and plastic properties [25, 30, 31].

Generally, the average streamlines, and hence the bulk flow behavior, determined through these

continuum simulations have been in fairly good agreement with the predictions of discrete

Page 23: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

5

element modeling [24, 27, 30]. For example, Kamrin [24] proposed an elasto-plastic constitutive

law for use in a Lagrangian finite element model of granular flows in inclined chutes, rectangular

silos and annular Couette cells. The predicted flow fields were compared with those calculated

using DEM. Andrade et al. [30, 31] used the same material law to model the static 3D

compression of sand particles. Forterre and Pouliquen [27] used a visco-plastic constitutive law

with the fixed-grid finite difference method, which is equivalent to an Eulerian mesh to simulate

granular flows in the geometries of [24] and in rotating drums. They compared their results with

experimental measurements and DEM predictions of bulk flow [27]. The Lagrangian

formulation has been used in most of the papers that modelled granular flows using the finite

element method [24, 32].

1.3 Objectives

The first two objectives of the research were to measure the impact velocities of particles in the

granular flow produced by a tub vibratory finisher, and to compare these with the impact

velocities and flow characteristics predicted by numerical simulations using the discrete element

method (DEM). It was expected that many of these research results would be useful in modeling

the local impact conditions found in other applications of granular flow. The third objective was

to investigate whether continuum models might offer a simpler means of predicting bulk flow

and impact velocities.

1.3.1 Objective 1

It was of interest to develop a new probe based on a high-speed laser displacement sensor to

measure directly the surface-normal impact velocities of particles in vibrationally-fluidized

granular flows. The probe was demonstrated in a tub vibratory finisher where it was used to

make measurements in various locations and directions within flows of steel and porcelain

media.

Page 24: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

6

1.3.2 Objective 2

The main goal was to compare the bulk flow and impact velocities obtained from a three-

dimensional discrete element model with those measured by using the setup including the

submerged laser velocity probe used in the earlier experiments.

1.3.3 Objective 3

The goal was to compare the bulk flow velocities in a vibrationally-fluidized granular flow

modeled as a visco-plastic media using different finite element formulations with those obtained

in DEM simulations and experimental measurements. The equivalent material properties of the

granular media had to be estimated using grain-scale 3D DEM simulations prior to modeling the

media as a continuum.

1.4 Thesis outline

The first Chapter of this thesis gives an introduction to vibrationally-fluidized granular flows and

how to study their bulk and local behavior experimentally and numerically. Chapter 2 covers the

experimental procedure using an impact velocity probe, the corresponding measurements and the

analysis of the results. This work has been published as:

K. Hashemnia, A. Mohajerani, J. K. Spelt, Development of a laser displacement probe to

measure particle impact velocities in vibrationally fluidized granular flows, Powder Technology,

235 (2013) 940-952

Page 25: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

7

In Chapter 3, the vibrationally-fluidized bed of a tub vibratory finisher was modeled using the

discrete element method and the resulting bulk flow and impact velocities of the particles at

different points in the granular flow were compared with the experimental measurements. This

study has been published as:

K. Hashemnia, J. K. Spelt, Particle impact velocities in a vibrationally fluidized granular flow:

measurements and discrete element predictions, Chemical Engineering Science, 109 (2014) 123–

135.

Chapter 4 presents a study of how the vibrationally-fluidized bed can be modeled as a continuum

using different finite element methods, and compares the results with DEM simulations. The

following article has been submitted on this part of the project:

K. Hashemnia, J. K. Spelt, Finite Element Continuum Modeling of Vibrationally-Fluidized

Granular Flows, Chemical Engineering Science, 2014, submitted.

Finally, Chapter 5 reviews the overall conclusions of the Ph.D. research.

1.5 References

[1] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, A. Kong, P. J. Narayan, B. C. Hancock,

Granular segregation in discharging cylindrical hoppers: A discrete element and experimental

study, Chem. Eng. Sci. 62 (2007) 6423 – 6439.

[2] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, B. C. Hancock, Predicting the flow mode

from hoppers using the discrete element method, Powder Technol. 195 (2009) 1–10.

[3] R.Y. Yang, R.P. Zou, A.B. Yu, Microdynamic analysis of particle flow in a horizontal

rotating drum, Powder Technol. 130 (2003) 138–146.

[4] V. Jasti, C. F. Higgs, Experimental study of granular flows in a rough annular shear cell,

Phys. Rev. E 78 (2008) 041306-1-8.

[5] N.G. Deen, M. Van Sint Annaland, M.A. Van der Hoef, J. A. M. Kuipers, Review of discrete

particle modeling of fluidized beds, Chem. Eng. Sci. 62 (2007) 28-44.

Page 26: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

8

[6] X. Z. An, C.X. Li, R.Y. Yang, R.P. Zou, A.B. Yu, Experimental study of the packing of

mono-sized spheres subjected to one-dimensional vibration, Powder Technol. 196 (2009) 50–

55.

[7] S.E. Naeini, J.K. Spelt, Two-dimensional discrete element modeling of spherical steel media

in a vibrating bed, Powder Technol. 195 (2009) 83-90.

[8] S.E. Naeini, J.K. Spelt, Development of single-cell bulk circulation in granular media in a

vibrating bed, Powder Technol. 211 (2011) 176-186.

[9] C. Fan, X. T. Bi, J. R. Grace, A. Goto, Grid zone performance of a fluidized bed through

analysis of local solids holdup signals, Powder Technol. 219 (2012) 37–44.

[10] D. Ciampini, M. Papini, J. K. Spelt, Impact velocity measurement of media in a vibratory

finisher, J. Mater. Process Tech. 183(2007) 347-357.

[11] A. Yabuki, M. R. Baghbanan, J. K. Spelt, Contact forces and mechanisms in a vibratory

finisher, Wear 252 (2002) 635-643.

[12] S. Wang, R. S. Timsit, and J. K. Spelt, Experimental Investigation of Vibratory Finishing of

Aluminum, Wear 243 (2000) 147-156.

[13] A. Munjiza, The combined finite discrete element method, John Wiley & Sons, Ltd,

Chichester, West Sussex, England, 2004.

[14] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of particulate

systems: A review of major applications and findings, Chem. Eng, Sci. 63 (2008) 5728-

5770.

[15] C.H. Tai, S.S. Hsiau, C.A. Kruelle, Density segregation in a vertically vibrated granular bed,

Powder Technol. 204 (2010) 255-262.

[16] M. Majid, P. Walzel, Convection and segregation in vertically vibrated granular beds,

Powder Technol. 192 (2009) 311-317.

[17] R. L. Stewart, J. Bridgwater, Y. C. Zhou, A.B.Yu, Simulated and measured flow of granules

in a bladed mixer - A detailed comparison, Chem. Eng. Sci56 (2001) 5457-5471.

[18] B. Freireich, J. Litster, C. Wassgren, Using the discrete element method to predict collision-

scale behavior: A sensitivity analysis, Chem. Eng. Sci. 64 (2009) 3407- 3416.

[19] J.J. McCarthy, V. Jasti, M. Marinack, C.F. Higgs, Quantitative validation of the discrete

element method using an annular shear cell, Powder Technol. 203 (2010) 70-77.

Page 27: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

9

[20] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, Stress results from two-dimensional

granular shear flow simulations using various collision models, Phys. Rev. E 71 (2005)

061307-1-11.

[21] K. Hashemnia, A. Mohajerani, J. K. Spelt, Development of a laser displacement probe to

measure particle impact velocities in vibrationally fluidized granular flows, Powder

Technol., 235 (2013) 940-952.

[22] K. Hashemnia, J. K. Spelt, Particle impact velocities in a vibrationally fluidized granular

flow: Measurements and discrete element predictions, Chem. Eng. Sci., 109 (2014) 123–

135.

[23] L. Xiang, W. Shuyan, L. Huilin, L. Goudong, C. Juhui, L. Yikun, Numerical simulation of

particle motion in vibrated fluidized beds, Powder Technol. 197 (2010) 25–35.

[24] K. Kamrin, Nonlinear elasto-plastic model for dense granular flow, Int. J. Plasticity, 26

(2010) 167–188.

[25] C. H. Rycroft, K. Kamrin, M. Z. Bazant, Assessing continuum postulates in simulations of

granular flow, J. Mech. Phys. Solids, 57 (2009) 828–839.

[26] P. Jop, Y. Forterre, O. Pouliquen, A constitutive law for dense granular flows, Nature, 441

(2006) 727-730.

[27] Y. Forterre, O. Pouliquen, Flows of dense granular media, Annu. Rev. Fluid Mech. (2008)

1-24.

[28] G. D. R Midi, On dense granular flows, Eur. Phys. J. E (2004) 341-365.

[29] F. Cruz, S. Emam, M. Prochnow, J. N. Roux, and F. Chevoir, Rheophysics of dense

granular materials: Discrete simulation of plane shear flows, Phys. Rev. E 72 (2005)

021309-1-17.

[30] J. E. Andrade, Q. Chen, P. H. Le, C. F. Avila, T. M. Evans, On the rheology of dilative

granular media: Bridging solid- and fluid-like behavior, J. Mech. Phys. Solids 60 (2012)

1122–1136.

[31] J. E. Andrade, C. F. Avila, S. A. Hall, N. Lenoir, G. Viggiani, Multiscale modeling and

characterization of granular matter: From grain kinematics to continuum mechanics, J.

Mech. Phys. Solids 59 (2011) 237–250.

[32] J. E. Andrade, X.Tu, Multiscale framework for behavior prediction in granular media,

Mech. Mater. 41 (2009) 652–669.

Page 28: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

10

Chapter 2

2 Development of a Laser Displacement Probe to Measure Particle Impact Velocities in Vibrationally-Fluidized Granular Flows

2.1 Introduction

Vibratory finishing is widely used to deburr, polish, burnish, harden and clean metal, ceramic

and plastic parts. In a tub vibratory finisher, a container filled with granular media is oscillated

by an eccentric rotating shaft so that it develops a vibrationally-fluidized circulatory bulk flow of

the media that is largely two-dimensional. The media have both a large-scale bulk flow velocity

and a much smaller-scale local impact velocity. Therefore, work-pieces that are entrained in the

flowing media are subjected to the repetitive, high-frequency impacts of the surrounding

particles. The granular media can be made of many materials of various shapes such as spheres,

cylinders or pyramids, with either a smooth or abrasive texture.

Granular flow has been studied in non-fluidized beds such as in hopper discharge, conveyers,

chutes and mixers [1-8], and in fluidized beds such as in vibratory finishers, vibrating sieves and

rotating drums [3, 4, 9-14]. A granular media becomes fluidized as the vertical acceleration

amplitude exceeds gravity, resulting in a marked decrease in the contact pressure and constraint

provided by surrounding media [12, 14]. In addition to bed vibration, fluidization can result

from interstitial gas flow and from particle-wall interactions resulting from high-speed chute

flow [15].

Within a vibratory finisher, the erosive wear and plastic deformation of workpieces are largely

affected by the velocity, frequency, and direction of the impacts with particles. High impact

velocities can cause fracture and fragmentation while low impact velocities can make the process

less efficient. The impact velocity is also closely related to the breakage of granular materials in

Page 29: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

11

many processes such as drug tablet coating within rotating drums, bulk materials handling, food

processing, and particle attrition in vibratory finishing [16]. Large particle impact velocities may

also result in erosion of machine components in processes such as vibratory sieving and mixing

[16].

Measurement of local quantities like the impact velocity between vibrating particles in a

vibrationally-fluidized granular media is challenging, because of the relatively small scale of the

motion and difficulties in designing probes capable of measuring local quantities without

disturbing the media. Therefore, most existing research in the field of flowing granular media

has focused on the bulk flow of granular media inside different machines [1, 2, 5-9, 11-14, 17-

19]. The discrete element method (DEM) is a numerical technique that is often used to solve

problems involving transient dynamics of systems comprising a large number of moving bodies

that interact with each other. DEM simulations have been used to study the bulk flow behavior

of granular media and the numerical results have been verified experimentally [7, 12].

Generally, these investigations have shown that discrete element simulations give reasonable

predictions of bulk flow velocity and volume fraction in both fluidized and non-fluidized flows.

Only a few studies have focused on the local behavior of the media in granular flows [7, 14, 20-

22, 23, 24] and some of these have investigated vibrationally-fluidized flows [14, 23]. For

example, DEM was used to obtain the planar distributions of the collision velocity and the

collision frequency of the granular media inside a horizontal rotating drum [7]. This

configuration was also examined in [20], using DEM to predict the effects of particle stiffness,

size and coefficient of restitution on the solid fraction, collision frequency, and impact velocity

inside the granular media. In most of these studies, the predicted collision scale quantities such

as impact velocity, collision frequency or impact energy, were not validated experimentally [7,

20, 22, 23]. Therefore, it is not known if the commonly used approaches and parameters in DEM

simulations yield correct predictions of the impact velocity. This remains as a significant

limitation since the impact forces that govern erosion, wear and fracture in granular flows are

fundamentally linked to the particle impact velocity and kinetic energy [20].

A number of purely experimental studies have been done on the local behavior of granular media

[24- 26]. Yabuki et al. [26] used a 3 mm diameter “floating”‎pad‎connected‎ to‎ four‎ resistive‎

force sensors to measure both normal and tangential impact forces. They found that forces

Page 30: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

12

normal to the sensor were approximately 10 times greater than the tangential forces in both the

wet and dry conditions. In the wet condition, a small percentage of the media had a tangential to

normal force ratio approximately equal to the measured dynamic coefficient of friction between

media, indicating that sliding had occurred. The force signals also showed that the packing of the

media in the vibrating bowl finisher was relatively loose, with many gaps that were akin to

bubbles. The variation of the impact forces in steel and porcelain media in a tub vibratory

finisher was investigated in [25, 26] using a unidirectional high-speed impact force sensor.

Local impact velocities were inferred from impact forces, but this indirect measurement of

impact velocity was dependent on a system calibration and required a relatively large sensor that

disrupted the bulk flow in the finisher [25]. These impact force measurements were sensitive to

the stiffness of the sensor, and hence were specific to the experimental setup. This points to a

key advantage of measuring impact velocities directly; the impact velocities of particles will be

unaffected by the sensor compliance.

Comparing the results of investigations in tub and bowl-type finishers, and in finishers operating

with different amplitudes [25-27], it can be concluded that the major characteristics of the local

behavior of the particles are largely independent of the type of vibratory finisher producing the

granular flow. For example, the vibrationally-fluidized flows tend to be loosely packed and

collisions with workpiece surfaces are dominated by the normal component of impact.

The main objective of the present investigation was to develop a new probe based on a high-

speed laser displacement sensor to measure directly the surface-normal impact velocities of

particles in vibrationally-fluidized granular flows. The probe was demonstrated in a tub

vibratory finisher where it was used to make measurements in various locations and directions

within flows of steel and porcelain media.

2.2 Experiments

2.2.1 Measurement approach

Page 31: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

13

There are different methods to measure the particles velocity in the fluidized beds, such as: laser-

Doppler velocimetry, photographic and video techniques, optical fiber probes and laser

displacement sensors. Photographic methods have been used to measure bulk flow properties,

but the limited spatial resolution of these approaches makes them unsuited to the measurement of

local impact velocities [11, 27]. Laser-Doppler velocimetry is restricted to the low solid

concentrations (loose media). Optical fiber probes have been used to measure bulk flow

velocities and the void fraction of media passing transversely across the end of the sensor [28].

Some designs can measure displacements along the optical probe axis using a correlation with

the amount of reflected light, but such devices become inaccurate when the light reflected by the

particle varies because of a significant transverse velocity across the probe tip. Therefore, laser

displacement sensors were chosen to measure the impact velocity of particles inside a relatively

packed granular flow produced in the tub vibratory finisher.

Laser displacement sensors measure the distance to an object using triangulation as illustrated in

Fig. 2.1. Laser light reflected from the object is concentrated on a linearized charge-coupled

device (LI-CCD) such that its position on the LI-CCD depends on the distance to the object.

Two laser sensors with different working distances were used: LK-G 82 with a working distance

of 80±15 mm and a resolution of 0.2 µm, and LK-G 157 with a working distance of 150±40 mm

and a resolution of 0.5 µm (Keyence Inc.). Both sensors had a red semiconductor laser light

source, but the LK-G 82 had a round beam diameter of 70 µm while the LK-G 157 had an

elliptical beam with minor diameter of 120 µm and a major diameter of 1700 µm. The data

acquisition rate of these laser sensors was adjustable from 1-20 kHz [29].

Page 32: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

14

Fig. 2.1: Schematic of laser displacement sensor triangulation principle.

2.2.2 Tub motion characterization

The laser probe developed in the present work was tested in a tub vibratory finisher (Burr-Bench

2016, Brandon Industries, TX) shown in Fig. 2.2a. Two types of granular media were used: 91

kg of steel balls (ABCO, Abbott Ball Company Inc., West Hartford, CT, USA) and 41 kg of

porcelain balls (MicrobriteTM

, Abrasive Finishing Inc., Chelsea, MI, USA). Additional mass was

added in the latter case to maintain the same tub dynamics. Table 2.1 shows the density,

Young's modulus, mass and diameter (mean ± standard deviation) of each particle in the steel

and porcelain media. The Young’s‎ modulus‎ and‎ Poisson’s‎ ratio‎ of the porcelain balls were

measured ultrasonically on a disk (3 mm thick, 5 mm diameter) ground from a single ball. An

ultrasonic pulser-receiver (5072PR, Olympus Corp.) was used to measure the velocities of

longitudinal and shear waves to determine these elastic properties [30].

Page 33: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

15

Table 2.1: Physical and mechanical properties of steel and porcelain media (± indicates standard

deviation, 10 measurements)

Material Young's

Modulus (GPa)

Poisson's

Ratio

Density

(g/cm3)

Mass (g)

[25]

Diameter

(mm) [25]

Steel 207 0.27 7.8 1.03±0.002 6.3±0.005

Porcelain 71.4 0.23 2.4 0.29±0.020 6.1±0.180

The granular media was confined to a compartment made of two 12 mm thick glass plates, 21 cm

apart, installed rigidly in the central portion of the tub (Fig. 2.2b). A tri-axial accelerometer and

two uni-axial accelerometers were attached to the top of the tub at points A and B (Fig. 2.2a) in

order to measure the horizontal and vertical acceleration components. The accelerometer signals

were filtered to eliminate high-frequency noise, and the dominant mode of vibration at 47 Hz

was determined using a fast Fourier transform. The velocity and displacement of points A and B

were then calculated using symbolic integration of the equations of motion in MATLAB. The

angular velocity, Ω, and the angular acceleration, α, that characterized the tub rocking motion

were then determined using:

B A AB ABa a r r (2.1)

B A ABv v r (2.2)

where rAB is the position vector from A to B. Equations (2.3) and (2.4) were then used to find

the amplitudes of displacement, velocity and acceleration at other points (P) on the tub wall.

P A AP APa a r r (2.3)

P A APv v r (2.4)

The accuracy of these procedures was assessed by comparing the measured vertical displacement

at point A and the calculated vertical displacement at point C (middle of tub floor) in an empty

tub with direct laser measurements at these two points. It was found that the tub displacement

amplitudes calculated from the tub accelerations measured by the accelerometers (0.86 mm at A

and 0.64 mm at C) were 10% deviated from the laser measurements, probably due to the noise in

the accelerometer readings since the errors did not display a consistent bias.

Page 34: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

16

Fig. 2.2: Schematic of the tub vibratory finisher showing the bulk flow circulation direction: (a)

side view, (b) top view. The probe is shown at location H. The curved arrows show the flow

streamlines around the elliptical tube. Dimensions are in millimeters and are drawn to scale.

(a)

(b)

Page 35: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

17

2.2.3 Displacement sensor characterization

The accuracy of the impact velocity measurement using the laser displacement sensor was

assessed by dropping a spherical steel ball of the type used in the finisher from an electromagnet

held various heights above a glass plate. The laser displacement sensor was positioned vertically

under the glass plate such that the laser beam reflected from the midpoint of‎ the‎ball’s‎bottom‎

surface, and‎the‎ball‎remained‎within‎the‎sensor’s‎working‎distance during its fall. To simulate

the oblique impacts of the balls in the vibratory finisher as they vibrate against a target while

moving with the bulk flow velocity, the glass plate was tilted between 0º and 16º (the maximum

impact angle corresponding to the largest bulk flow velocity), to examine impacts of different

incidence angles while keeping the incident laser beam normal to the glass.

Figure 2.3 shows the experimental results for the displacement and velocity of a 6.3 mm

diameter smooth steel ball falling onto the glass plate from a height of 12.7 mm and bouncing

five times. A quadratic relation was fitted to the approach portion of the first impact of Fig. 1.3

in order to determine the impact velocity from the first derivative. For a predicted impact

velocity of 500.0 mm/s (maximum uncertainty of 2.0 mm/s or 0.4%), the average measured

value was 496.9 mm/s (standard deviation of ±0.3 mm/s based on 6 measurements); i.e. a 0.6%

error. The coefficient of restitution corresponding to the ratio of the measured rebound and

incident velocities was 0.755±0.004. The surface-normal component of the impact velocity for

an oblique impact with a small inclination angle of the glass plate (less than 16) was predicted

with less than 10% error using the laser sensor.

The apparatus was also used to measure the coefficient of restitution of the porcelain balls

colliding with a glass plate. In this case, tweezers were used to release a porcelain ball from a

platform at a prescribed height. The coefficient of restitution was then calculated as the ratio of

the measured rebound and incident velocities to be 0.84±0.01 (mean ± standard deviation based

on 6 measurements).

Page 36: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

18

Fig. 2.3: Displacement of a steel ball falling onto a glass plate acquired by the laser displacement

sensor (LK-G 82). The time step of data acquisition was 50 s (20 kHz sampling rate). The

arrows show the first, second and third impact points.

2.2.4 Test procedures in the vibratory finisher

In order to minimize the effects of vibration, the laser sensor was attached to an adjacent brick

wall using a horizontal steel beam connected to a vertical 30 mm diameter steel rod (Fig. 2.4a).

The laser was then directed down an elliptical aluminum shielding tube (70 mm major axis, 25

mm minor axis) that was immersed in the flowing media and supported by a 20×20 mm steel box

section that was also attached to the brick wall. The major axis was parallel to the bulk flow to

minimize disturbance to the flow. The laser sensor had no contact with the elliptical tube or the

media, and thus it was effectively stationary during the operation of the vibratory finisher. A

very small vibration of the sensor was due to acoustic wave propagation, but this found to be

negligible; i.e. less than 0.2 m. A glass window at the bottom of the tube was used to measure

impact velocities against the glass of upward flowing media. Two 12×40 mm glass windows at

opposite ends of the major axis of the cross-section were used to measure impact velocities in the

horizontal direction by redirecting the laser with a 45º polished steel mirror (Fig. 2.4c). To

Page 37: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

19

minimize its vibration, the mirror was attached directly to the laser sensor using a rod passing

through the elliptical pipe without contact so that it was isolated from the vibration of the media.

As with the laser sensor, a negligible vibration was due to acoustic waves (measured to be less

than 0.5 m.)

Although the shielding elliptical pipe was supported by a steel bar connected to the brick wall,

some vibration of the three glass windows embedded in its wall was inevitable due to its

exposure to the granular flow. Although this window vibration did not influence the incident and

reflected laser light from the media, it was of interest to quantify the amplitude of this vibration

relative to the motion of the vibrating media. The vibration of the glass window could not be

recorded directly using the second sensor, since the window was covered by the media in the tub.

Therefore, a correlation was established between the window vibrations and the vibration of the

shielding tube to which they were bonded. To measure the vibration of each of the sideways

facing windows, the second laser sensor (LK-G 82) was positioned outside the media and

reflected from a point on the shielding tube in the horizontal direction (Fig. 2.4c). To measure

the vibration of the bottom window used to measure vertical impact velocities, the second laser

was reflected from a steel beam clamped to the shielding tube (Fig. 2.4b). A piece of tape was

put on each window to acquire the window vibration when the tub was operating using the

primary laser, and simultaneously the vibration of the shielding tube was recorded using the

second laser. A correlation was then established between the vibration amplitudes of the

windows and the output of the second laser. Figure 1.5 illustrates this for the window of Fig.

2.4c used to measure the particle horizontal impact velocities.

Page 38: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

20

(b)

(a)

Page 39: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

21

Fig. 2.4: (a) Schematic of the test apparatus used to measure the impact velocity of the balls in

the vertical direction in the vibratory finisher. Sensor, tub and balls are drawn to scale. Note that

the steel rod supporting the laser sensor was mounted directly to a brick wall and was thus

isolated from vibration. (b) enlarged portion of (a). (c) enlarged view of the probe used to

measure the impact velocity in the horizontal (upstream or downstream) direction.

(c)

Page 40: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

22

10 10.02 10.04 10.06 10.08 10.1 10.12 10.14 10.16 10.18 10.2 10.22 10.24

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

0.2

0.25

Time (s)

Positio

n (

mm

)

Fig. 2.5: The vibration of the window in Fig. 2.4c and the simultaneous vibration of the shielding

tube. The solid line represents the window motion while the dotted line represents the shielding

tube motion.

The impact velocity in different orientations was measured at three locations in the tub. Each

location was 5 cm below the free surface of the flowing media which had the curved shape

shown in Fig. 2.6. The position of the shielding tube was adjusted slightly each time so that the

measurement locations in the flow were fixed regardless of the window being used. Location H

corresponded to one of the locations used in the same tub in ref. [25], allowing direct comparison

with the data obtained with the impact load cell. The tub as well as the type and amount of steel

and porcelain media were the same as in ref. [25], ensuring that the flow conditions were

identical to those in [25].

The vibrating tub finisher produced a single-cell circulation as illustrated in Fig. 2.6 because of a

resultant upward shear force between the media and the right wall of the tub [12]. This also

produced a free surface that had‎a‎pronounced‎slope‎from‎the‎“high‎wall”‎to‎the‎“low‎wall”,‎and‎

a bulk flow near the free surface that accelerated as the media cascaded downward to the low

wall.

Page 41: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

23

The impact velocities of the steel balls were measured with the laser sensor pointing in various

directions depending on the location: At location H, measurements were made in three

directions: horizontal upstream facing the bulk flow, horizontal downstream, and vertically

facing downward. At location M, the impact velocity was measured in the horizontal direction

facing upstream and in the downward vertical direction. At location L, the impact velocity was

measured in the horizontal direction facing upstream. In the porcelain media, impact velocities

were measured at locations H and M in the horizontal direction facing upstream toward the

oncoming flow and in the vertical direction facing downward. Two additional sets of

experiments were performed in the porcelain media at locations 2 cm above and 2 cm below

location M in order to assess the sensitivity of the impact velocity to depth.

Prior to each experiment, the relevant window was cleaned with acetone to minimize laser

reflection from the glass. The steady-state tub vibration was then measured using the

accelerometers described previously, and the laser sensor signals were acquired for 33 s.

Fig. 2.6: Schematic of the vibrating tub showing the bulk flow pattern and the three locations of

the probe as it was when measuring toward the right. The x-y distances correspond to the fixed

locations in the flow used in all measurement directions. Dimensions are in millimeters.

Page 42: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

24

The porcelain particles were much lighter than the steel particles. Therefore, in the experiments

with the porcelain media, in order to maintain the same dynamics of the tub while preserving the

bed depth used with the steel media, additional 50 kg steel media was added equally to the

empty compartments located at either end of the tub on the outside of the glass partitions.

Displacement measurements confirmed that the vibration amplitude of the tub differed by less

than 5% between the experiments done with the steel and porcelain media.

2.3 Signal analysis

2.3.1 Ball motion in the vibratory finisher and the laser sensor displacement signals

Two components of particle motion were evident: a cyclic impact against the glass window at

the tub vibration frequency of 47 Hz, and bulk flow parallel to the glass window (Fig. 2.7a). The

angles shown in the Figs. 2.7b and 2.7c, illustrate the range through which the laser sensor can

accurately read the position of the points on the surface of a ball in the plane of the laser beams.

This range was greatest when the nominal plane of the sensor was on a diametral plane of the

ball (i.e. plane containing the ball axis) and became smaller as the path of the ball shifts further

from the plane of the sensor. In any event, Fig. 2-7c illustrates that the laser beam reflected

adequately over an asymmetric range of angles in the diametral plane (i.e. θ1 and θ2 were

approximately 80º and 65º, respectively).

The vibrations of the balls and the shielding tube were recorded simultaneously using two laser

sensors as described previously. Figures 2.8a-c shows the typical displacement signal obtained

from the motion of the steel balls in the finisher with a data acquisition rate of 2 kHz. Using the

correlation mentioned in the Section 2.4, the window vibration was calculated from the measured

shielding tube vibration. For example, Fig. 2.8a shows the window vibrations and the

simultaneous signals from 5 balls passing before the sensor. Figure 2.8b shows these signals for

the passage of a single ball, and illustrates that the window and the passing balls vibrated in

Page 43: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

25

phase at 47 Hz. The displacement signal contained some erroneous readings of very small

displacements (indicated by arrows in Fig. 2.8a) generated when the laser beam passed through

the empty space between two adjacent balls, or when the laser was reflected from ball surfaces

with excessively steep slopes. In the former case, the sensor recorded the laser beam reflecting

from either the mirror surface or the glass window.

Laser

sensor

Emitter

Receiver

Direction of the

bulk flow

Balls path

Inappropriate

reflection

Window

Fig. 2.7: (a) Inadequate laser beam reflection from the ball surfaces with excessively steep

slopes; (b) Laser beam reflecting from a non-diametral plane; (c) The range for the laser sensor

to accurately detect the position of the ball in a diametral plane (θ1 ≠ θ2).

(a) (b)

(c)

Page 44: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

26

29.4 29.5 29.6 29.7 29.8 29.9

-1.6

-1.4

-1.2

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

Time (s)

Positio

n (

mm

)

(b)

(a)

Page 45: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

27

29.48 29.485 29.49 29.495 29.5 29.505 29.51 29.515 29.52 29.525 29.53-0.9

-0.7

-0.5

-0.3

-0.1

0.1

0.3

0.5

Time (s)

Positio

n (

mm

)

0.04 0.06 0.08 0.1 0.12 0.14 0.16

-0.9

-0.8

-0.7

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0

Time (s)

Positio

n (

mm

)

Fig. 2.8: (a) Displacement signal (normal to the glass window) generated by the passage of five

balls parallel to the glass window in the rightward direction at location H shown with dotted

lines. Arrows show readings when laser passed between balls. (b) a magnified portion of (a)

corresponding to the passage of a single ball; (c) a magnified portion of (b); (d) a magnified

portion of Fig. 2.3 which shows the drop test results. The vertical arrows in (c) and (d) show the

points corresponding to contact between the window and ball. The solid lines show the window

position.

(c)

(d)

Δt

0.15Δt

0.3Δt

Page 46: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

28

The large-scale sinusoidal variation of the average displacement reflects the curvature of the

balls and their bulk flow past the sensor (Fig. 2.8a). If the nearest approach of the ball was

coincident with the displacement of the window, the ball was passing before the sensor on a

diametral plane as shown in Fig. 2.8b. Figure 2.8c shows a portion of the displacement signal

corresponding to a portion of the ball surface that is not itself making contact with the window

(e.g. location B in Fig. 2.7b or a point between 1 and 2 in the diametral plane of Fig. 2.7c).

The bulk flow velocity was then be estimated by dividing the chord length a by the

corresponding time interval while the ball passed in front of the sensor along a diametral plane

(Fig. 2.9). The chord length a was calculated from

2 22 ( ( ) )a R R h (2.5)

where R is the ball radius, and h was chosen to be 0.5 mm in the calculations (i.e. h was the

maximum difference between the ball and window positions); therefore, a would be 3.1 mm for

the steel balls and 3.35 mm for the porcelain balls.

Fig. 2.9: Chord length a used to calculate the bulk flow velocity of the balls passing the sensor in

a diametral plane.

The bulk flow velocity at location H in the horizontal direction facing upstream determined from

this method was 13±1.0 mm/s (average ± 95% confidence interval based on 20 measurements).

Although it was not possible to visually record the bulk flow exactly at location H, video images

Page 47: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

29

of the media flowing past one of the glass partitions in the finisher at the same x-y location (Fig.

2.2) were used to estimate the value. A video image with a 40×40 mm field of view was

replayed frame-by-frame to track individual balls over a distance of 20 mm. This gave a bulk

flow velocity adjacent to the glass partition of 15±1.0 mm/s (average ± 95% confidence interval,

20 measurements). While viewing the free surface (Fig. 2.2b), the average bulk flow velocity

midway between the glass partitions was measured similarly to be 85% of that at the glass

partitions (64 mm/s in the middle and 75 mm/s at the glass partitions). If it is assumed that this

correlation between bulk flow velocities in the middle of the flow and at the glass partitions was

also valid at location H, the actual bulk flow velocity at location H was 13 mm/s, in very good

agreement with the value from the laser sensor.

Another property of the granular flow that was obtained from the displacement signals was the

“particle passage frequency” defined as the number of balls passing in front of the sensor per unit

time. In this case, all balls passing within 0.5 mm of the diametral plane were considered (as

judged by the measured distance of nearest approach to the window). This provided a measure

that was a function of both the bulk flow velocity and the degree of packing in the granular flow.

The ascending line in each oscillation of Fig. 2.8b corresponded to the approach of a ball to the

surface of the window (i.e. the impact velocity). As can be seen in Fig. 2.8c, the displacement

curves were not as sharp near the impact points as they were in the drop tests (Fig. 2.3). The

main reason for this difference was that the window vibrated in-phase with the balls (Fig. 2.8b)

and with an amplitude that was 30-50% of the ball amplitude; therefore, after a ball collided with

the moving window, it rebounded from the window, but continued to move toward the sensor

with a reduced speed. The ball was then accelerated in the opposite direction by the window

moving outward once again (Fig. 2.11). This is discussed further in Section 3.3. Another factor

that can contribute to the rounding of the impact peaks in Fig. 2.8c was the tangential velocity of

the ball parallel to the window, which was about 17% of the average normal impact velocity.

This caused the laser reflection point to shift on the ball surface creating an error in the actual

displacement attributed to the surface-normal impact velocity; however, as explained below, this

error was negligible (less than 1%) in the present experiments.

Page 48: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

30

2.3.2 Impact velocity calculation

Impact velocities normal to the sensor window were determined by calculating the slopes of the

straight line segments of the ascending (approach) portions of the curves shown in Fig. 2.8c prior

to contact. The following procedure was used to consistently identify the data points that were

fitted to a straight line to determine the impact velocity: 1. The minimum and maximum points

were used to calculate the half-period, Δt and the midpoint (Fig. 2.8c). 2. The selected data points

were in the range from 0.30Δt to the left of the midpoint to 0.15Δt to the right of the midpoint.

This criterion ensured that the ball had not yet contacted the window as discussed further below.

A MATLAB code was written to identify the correct displacement signals and perform these

calculations.

The calculated impact velocities were relatively insensitive to the exact range of the data used on

the approach portion of the ball displacement signal. For example, if the data points were

selected to be those in the interval 0.25Δt extending from the point 0.05Δt to the right of the

minimum, the difference in the calculated impact velocity was less than 3%. Although impact

velocities could also be calculated for balls passing a certain distance off the diametral plane, in

the present work it was convenient to use only the signals used in bulk flow velocity calculations

for impact velocity calculation; i.e. the signals in which the difference between the point of

closest approach and the window was less than 0.1 mm (10% of the vibration amplitude of the

tub center of gravity (1.1 mm)).

The impact velocities calculated as described above were always greater than the maximum

window velocity which was determined at each location and direction using the dominant

frequency of 47 Hz and the recorded amplitude of the sinusoidal motion. The raw displacement

data were smoothed to remove erroneous readings using the 5-point central moving average

method prior to calculating the impact velocities (Fig. 2.8a).

Page 49: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

31

2.3.3 Double impact hypothesis

The double impact hypothesis was used above to explain the relatively blunt impact peaks in the

displacement signals. The displacement of the window (y1) and ball (y2) can be modeled as (Fig.

2.8c)

1 0 1 sin( )y y A wt

2 2 sin( )y A wt

(2.6)

(2.7)

where A1 and A2 are the corresponding vibration amplitudes and y0 is the initial distance between

the ball and the target. In Fig. 2.10 the ball and the window follow the solid and dashed lines,

respectively. Solving Eqs. (2.6) and (2.7) simultaneously, the time corresponding to the first

collision is

1 01

2 1

1sin ( )imp

yt

w A A

(2.8)

where w is the vibration frequency in rad/s. Then the target and the ball velocities immediately

prior to the first impact are, respectively

arg 1 1cos( )t et impV A w wt

2 1cos( )ball impV A w wt

(2.9)

(2.10)

The coefficient of restitution e is defined as

arg

arg

ball t et

t et ball

V Ve

V V

(2.11)

where the primed parameters correspond to the velocities after the impact and unprimed ones are

those before the collision. Since the mass of the window/shielding tube is much larger than that

of the ball, conservation of momentum indicates that the window velocity remains approximately

unchanged, so that the ball velocity after impact is

Page 50: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

32

arg 1 2 1( ) cos( )ball t et impV V e A A w wt (2.12)

As shown in Fig. 2.10, the ball can be assumed to move with a new constant speed after the first

impact until the second impact occurs when the window moves outward again. The time of

impact, and the target and ball velocities prior to and after the second collision can be determined

using Eqs. (2.8) - (2.12). Figure 2.10 was drawn assuming e=0.75, A1 =0.15 mm, A2 =0.30 mm,

y0 =0.12 mm, and w=295.1 rad/s, and it is found that the second impact occurs 3 ms after the

first one. This is a relatively short time compared with the period of oscillation of the ball

(τ=1/47‎s=21‎ms).

0 1 2 3 4 5 6 7 8 9 100

0.05

0.1

0.15

0.2

0.25

Time (ms)

Ball

and targ

et positio

ns (

mm

)

Fig. 2.10: Predicted positions of the window (dashed line) and ball showing first and second

collisions giving rise to the apparent rounding of the impact peaks. Modeled using data from

location H in the steel media in the rightward direction (Fig. 2.6).

Page 51: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

33

2.3.4 Error due to tangential bulk flow velocity

The bulk flow of the balls past the window changed the point of reflection of the laser and

caused an apparent change in the distance between the ball and the window, as illustrated in Fig.

2.11. Assuming an average value of the maximum bulk flow velocity parallel to the window of

νt=20.5 mm/s, the largest error this produced in the normal (impact) velocity measurement was

less than 1% and hence could be neglected.

Fig. 2.11: Schematic of a ball having both normal (impact) and tangential (bulk flow) velocity

components relative to the window.

2.4 Probe evaluation in the tub finisher

The probe was tested in the tub vibratory finisher using both the steel and porcelain spherical

media while maintaining the same tub vibration amplitude. Table 2.2 shows the component of

the bulk flow velocity parallel to the sensor window and the particle passage frequency of the

steel and porcelain media at locations H, M and L and in the different directions. The average

bulk flow velocity in the steel media was approximately 40% larger than that in the porcelain

Page 52: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

34

media. Within each media, the bulk flow velocity along the window was approximately the

same in all cases, ranging from 16 to 20 mm/s in the steel media, except for location H rightward

where it was only 13 mm/s, and from 8.7 to 13 mm/s in the porcelain media.

The particle passage frequency showed a much greater variation, from 0.22 to 0.73 s-1

in the steel

media, and from 0.16 to 0.49 s-1

in the porcelain media. Since the bulk flow velocity was

approximately the same at all locations in the steel media except H rightward, and in all locations

in the porcelain media except M downward, the variation in particle passage frequency was

mostly attributable to differences in the packing density of the media. This was examined by

defining‎a‎dimensionless‎“packing‎parameter”‎as‎ the‎product‎of‎ the‎particle‎passage‎ frequency‎

and the ball diameter divided by the bulk flow velocity, in order to eliminate the effect of the

bulk flow velocity. Table 2.2 shows that the packing parameter was maximum at location M

downward in both media, indicating that the media had the largest packing density at this

location and direction. In locations such as H leftward, this quantity was a minimum, probably

as a consequence of the wake formed behind the shielding tube. However, there was no

consistent relationship between the magnitude of the packing parameter and the flow direction.

For example, at location H it was essentially the same in the rightward and downward directions,

but at location M it was significantly greater in the downward direction. There was also no

consistent correlation between the relative packing and the type of media; i.e. in some locations

and directions the porcelain media were more highly packed, but not in others.

Page 53: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

35

Table 2.2: Bulk flow velocity parallel to window, particle passage frequency, packing parameter

and impact velocity of the steel and porcelain media for locations H, M and L in different

directions (number of data points in brackets). 95% confidence intervals on the mean values

based on student t distribution.

Location and

orientation of laser

Bulk flow

velocity (mm/s)

Particle passage

frequency (s-1

)

Packing

parameter

Impact velocity

(mm/s)

Steel balls

H rightward 13±1.0 (20) 0.28±0.05 (12) 0.139±0.0179 74±1.1 (1,032)

H downward 16±1.3 (20) 0.33±0.04 (14) 0.132±0.0136 80±1.3 (1,068)

H leftward 20±1.5 (20) 0.22±0.05 (15) 0.0693±0.0116 61±1.0 (621)

M rightward 19±2.0 (20) 0.35±0.06 (29) 0.113±0.0177 91±1.2 (1,254)

M downward 20±2.1 (20) 0.73±0.05 (32) 0.233±0.0200 110±1.6 (1,618)

L rightward 19±1.7 (20) 0.63±0.07 (19) 0.208±0.0219 72±1.8 (939)

Porcelain balls

H rightward 10±1.4 (17) 0.16±0.02 (25) 0.0976±0.0133 85±0.8 (2,786)

H downward 8.7±1.2 (17) 0.22±0.04 (20) 0.146±0.0262 140±1.3 (1,665)

M rightward 10±1.1 (17) 0.22±0.02 (24) 0.134±0.0138 67±1.3 (1,525)

M rightward

(2 cm above)

32±1.5 (559)

M rightward

(2 cm below)

94±1.4 (1,849)

M downward 13±1.0 (17) 0.49±0.03 (25) 0.232±0.0162 120±1.2 (1,946)

Figures 2.12a and 2.13a show the displacement signal and the probability density distribution of

the normal impact velocity of the balls in the vibratory finisher at location H in the rightward

direction. Figure 2.13a also shows the best-fit log-normal probability distribution. Figure 2.13a

was calculated from the slopes of more than 1,032 displacement signals corresponding to the

passage of 120 balls using 15 recordings of 33 s duration. The impact velocity distribution had a

mean value of 74 mm/s with more than 90% of the impact velocities in the range of 50-100

mm/s. In comparison, more than 80% of the impact velocities measured at the same location and

in the same direction in ref. [25] using an impact force sensor were in the range of 0-20 mm/s.

There are several reasons for this difference, although the most significant is probably the linear

correlation that was assumed to exist between measured impact force and the impact velocity

over the range 0-1,000 mm/s [25]. Since the impact force varied nonlinearly with the impact

velocity [4], the slope of a line fitted to the lower impact velocity range 0-100 mm/s was

Page 54: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

36

observed to be twice the slope of the line corresponding to the impact velocity over the range 0-

1,000 mm/s [4, 25]. Therefore, the correlation in [25] underestimated the impact velocities since

most of them were below 100 mm/s. Another reason for the difference in impact velocities was

that “non-impact”‎contact‎events‎(e.g.‎a‎ball‎that‎dwelled‎over‎the‎window‎without‎vibrating)‎that‎

were excluded in the present measurements were assigned an effective impact velocity in [25]

based on their quasi-static contact force. Thirdly, although the location and measurement

direction were identical, the impact probe used in ref. [25] was significantly larger than the

present probe (twice the volume and twice the projected area normal to the average bulk flow).

This may have altered the local conditions at the probe surface.

Figures 2.12b and 2.12c show the displacement signals of the steel balls moving in the vibratory

finisher at location H in the downward and leftward and directions, respectively. The best-fit

impact velocity probability distributions corresponding to the different directions at location H in

the steel media were quite different in both their breadth and their mean, as seen in Fig. 2.13a.

Figure 2.13b shows that this was also true for the impact velocity distributions for the rightward

and downward directions at point M and the rightward direction at point L in the steel media.

As at location H rightward, a log-normal probability density distribution of impact velocities was

fitted to the steel media data at location H downward and location M rightward (Fig. 2.13). In

contrast, the distributions of impact velocities at locations M downward and L rightward were

found to fit better to a Gaussian curve (Fig. 2.13).

The probability density distribution of steel media impact velocities at location H leftward was

skewed more than could be fitted using a log normal distribution, similar to the distributions

found in [25]; therefore, a generalized extreme value distribution [31] was used to fit this impact

velocity distribution (Fig. 2.13a).

Freireich et al. [20] used the discrete element method to study the effects of different parameters

on the impact velocity of 3 mm diameter glass beads moving in a rotating drum. They also

found that the probability density distribution of impact velocity was skewed with an elongated

tail to the rights, similar to the log-normal and generalized distributions used with the present

steel media.

Page 55: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

37

In contrast to the steel results, a Gaussian probability density distribution was found to fit best to

the impact velocities at locations H and M in the porcelain media, as shown in Fig. 2.14.

(a)

(b)

Page 56: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

38

Fig. 2.12: Raw displacement signals of balls moving in the vibratory finisher at location H: (a)

rightward, (b) downward, (c) leftward (Fig. 2.6). The signal from the window is shown in the

narrow band at the top of each graph.

(c)

Page 57: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

39

0 20 40 60 80 100 120 140 160 180 200 2200

0.05

0.1

0.15

0.2

0.25

Impact velocity (mm/s)

Pro

babili

ty d

ensity d

istr

ibution

0 20 40 60 80 100 120 140 160 180 200 2200

0.02

0.04

0.06

0.08

0.1

0.12

Impact velocity (mm/s)

Pro

babili

ty d

ensity d

istr

ibution

Fig. 2.13: The probability density distributions of the impact velocity of steel balls in the

vibratory finisher (a) at location H in the three orthogonal directions and (b) at locations M and L

in the directions indicated (Fig. 2.6). The types of distributions are shown in the figure.

(b)

(a)

Rightward-L-Steel

(Gaussian)

Downward-M-steel

(Gaussian)

Rightward-M-Steel

(Lognormal)

Rightward-H-steel (Lognormal)

Leftward-H-steel (Generalized extreme value)

Downward-H-steel (Lognormal)

Page 58: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

40

0 20 40 60 80 100 120 140 160 180 200 2200

0.02

0.04

0.06

0.08

0.1

Impact velocity (mm/s)

Pro

babili

ty d

ensity d

istr

ibution

Fig. 2.14: The probability density distributions of the impact velocity of porcelain balls in the

vibratory finisher at locations H and M in the directions shown (Fig. 2.6). All the distributions

are Gaussian.

The mean values of impact velocities in the steel and porcelain media along with the 95%

confidence intervals are listed in Table 2.2. As the sample sizes were sufficiently large (typically

N>30), the mean values of the impact velocities were normally distributed, regardless of the

underlying population distribution [32]; therefore, the 95% confidence intervals for the means

were calculated using the t distribution.

Table 1.2 shows that the packing parameter and the average impact velocity at location H were

smaller in the leftward direction than in the rightward direction. This can be interpreted as a

result of the wake produced behind the shielding tube.

In the porcelain media, it was observed that the impact velocity in the downward direction (Fig.

2.14 and Table 2.2) and the corresponding packing parameter (Table 2.2) were higher than those

in the rightward direction at location H and M, similar to what was observed in the steel media,

except at location H where the steel packing parameter was approximately the same in the

downward and rightward directions.

Rightward-H-Porcelain

Rightward-M-Porcelain

Downward-M-Porcelain

Downward-H-Porcelain

Page 59: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

41

Comparing the values of the rightward impact velocities at location M in the porcelain media

with those at points 2 cm above and below it (Table 2.2), it was concluded that impact velocity

of the particles is strongly dependent on depth, increasing sharply with distance from the free

surface.

Figures 2.13, 2.14, and Table 2.2 show that the average impact velocities in the porcelain media

were almost 15% greater than those in the steel media. There are two possible reasons for this

observation: Firstly, the density and mass of the porcelain balls was lower than that of steel balls;

therefore, for the same vibrational kinetic energy transferred from the wall, the porcelain balls

had a higher velocity than the steel balls. Secondly, based on the measurements of Section 2.3, it

may be assumed that the coefficients of restitution of impacts of porcelain balls with each other

and with the tub walls were larger than those corresponding to the steel balls; therefore, less

kinetic energy was dissipated during the collisions, thereby increasing the impact velocities of

the porcelain balls compared with the steel balls.

Freireich et al. [20] reported that the predicted mean impact velocity increased and the

probability density distribution of the impact velocity becomes less skewed as the coefficient of

restitution increased. This is consistent with the present experimental results where it was

observed that the impact velocity of the porcelain balls (with larger coefficient of restitution)

fitted better to a normal distribution (symmetric) rather than the lognormal distribution that was

used to fit the impact velocity of the steel balls.

Figure 2.15 shows the normalized mean values of bulk flow velocity, impact velocity and

packing parameter in the different locations and directions for the steel and porcelain media. It is

evident that all three parameters varied considerably in the various locations and directions, and

that there was no simple correlation between them. Moreover, the relative magnitudes of the

impact velocity, packing density and bulk flow velocity, and their dependence on location and

direction, were often different in the steel and porcelain media.

Page 60: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

42

Fig. 2.15: The mean values of bulk flow velocity, impact velocity and packing parameter

(normalized by the largest values for each media) in different locations and directions (Fig. 2.6)

with: (a) steel, and (b) porcelain media.

(b)

(a)

Page 61: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

43

2.5 Conclusions

A laser displacement sensor was used to construct a probe that can be immersed in a granular

flow to measure local impact velocities, bulk flow velocities and a measure of particle packing.

The accuracy of impact velocity measurements was verified using drop tests with the same

granular media. The sensor output was modeled to interpret laser reflections from media passing

before the sensor in different trajectories and to distinguish the media vibrational impact velocity

from the oscillations of the sensor window.

The sensor was demonstrated in a tub vibratory finisher with two types of media – steel and

porcelain spheres. In this case, the average bulk flow velocity parallel to the sensor window in

the steel media was almost 40% larger than that in the porcelain media, and both varied

appreciably over the different measurement locations within a single media. Similarly, the

impact velocity and particle packing varied considerably among the various locations and in

different directions. Although it was concluded that there was no direct proportionality between

the impact velocity of particles, the bulk flow velocity, and the packing density of the media, the

smallest packing density and impact velocity occurred in the wake formed behind the sensor

shielding tube. These two parameters were larger when the sensor was facing the average

moving bulk flow. It was also observed that the impact velocity of the particles was strongly

dependent on depth in the flow, increasing sharply with distance from the free surface. The

average impact velocities in the steel media were approximately 15% smaller than those in the

porcelain media due to differences in the density and coefficient of restitution. In addition to

providing a means of quantifying the impact energy distribution within a vibratory bed, the

experimental results can be used to validate the predictions of numerical simulations such as

discrete element modeling (DEM). Such models can help to understand the complex patterns of

behavior that have been observed within these vibrationally fluidized beds. The laser

displacement probe and the current procedures are directly applicable to the measurement of the

local impact velocities of non-spherical particles. The measurement of the bulk flow velocity in

the case of irregular particle shapes would be more difficult, since the variation in the average

displacement signal would be more irregular.

Page 62: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

44

2.6 References

[1] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, B. C. Hancock, Predicting the flow mode

from hoppers using the discrete element method, Powder Technol. 195 (2009) 1–10.

[2] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, A. Kong, P. J. Narayan, B. C. Hancock,

Granular segregation in discharging cylindrical hoppers: A discrete element and experimental

study, Chem. Eng. Sci. 62 (2007) 6423 – 6439.

[3] N.G. Deen, M. Van Sint Annaland, M.A. Van der Hoef, J. A. M. Kuipers, Review of discrete

particle modeling of fluidized beds, Chem. Eng. Sci. 62 (2007) 28-44.

[4] W.J. Stronge, Impact Mechanics, first ed., Cambridge University Press, New York, 2000.

[5] T. Kawaguchi, MRI measurement of granular flows and fluid-particle flows, Adv. Powder

Technol. 21 (2010) 235-241.

[6] X. Z. An, C.X. Li, R.Y. Yang, R.P. Zou, A.B. Yu, Experimental study of the packing of

mono-sized spheres subjected to one-dimensional vibration, Powder Technol. 196 (2009) 50–55.

[7] R.Y. Yang, R.P. Zou, A.B. Yu, Microdynamic analysis of particle flow in a horizontal

rotating drum, Powder Technol. 130 (2003) 138–146.

[8] V. Jasti, C. F. Higgs, Experimental study of granular flows in a rough annular shear cell,

Phys. Rev. E 78 (2008) 041306-1-8.

[9] C.H. Tai, S.S. Hsiau, C.A. Kruelle, Density segregation in a vertically vibrated granular bed,

Powder Technol. 204 (2010) 255-262.

[11] S.E. Naeini, J.K. Spelt, Two-dimensional discrete element modeling of spherical steel media

in a vibrating bed, Powder Technol. 195 (2009) 83-90.

[12] S.E. Naeini, J.K. Spelt, Development of single-cell bulk circulation in granular media in a

vibrating bed, Powder Technol. 211 (2011) 176-186.

[13] L. Xiang, W. Shuyan, L. Huilin, L. Goudong, C. Juhui, L. Yikun, Numerical simulation of

particle motion in vibrated fluidized beds, Powder Technol. 197 (2010) 25–35.

[14] C. H. Tai, S.S. Hsiau, Dynamic behaviors of powders in a vibrating bed, Powder Technol.

139 (2004) 221– 232.

[15] J. K. Spelt, C. E. Brennen, and R. H. Sabersky, Heat Transfer to Flowing Granular Material,

Int. J. Heat Mass Transf., 25 (6) (1982) 791-796.

[16]‎EDEM‎software’s‎website,‎http://www.dem-solutions.com/academic, 2011.

Page 63: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

45

[17] C. S. Campbell, Granular material flows–An overview, Powder Technol. 162 (2006) 208–

229.

[18] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of particulate

systems: A review of major applications and findings, Chem. Eng, Sci. 63 (2008) 5728-5770.

[19] A. Wu, Y. Sun, Granular Dynamic Theory and Its Applications, first ed., Metallurgical

Industry Press, Beijing and Springer-Verlag GmbH Berlin Heidelberg, 2008.

[20] B. Freireich, J. Litster, C. Wassgren, Using the discrete element method to predict collision-

scale behavior: A sensitivity analysis, Chem. Eng. Sci. 64 (2009) 3407- 3416.

[21] J.J. McCarthy, V. Jasti, M. Marinack, C.F. Higgs, Quantitative validation of the discrete

element method using an annular shear cell, Powder Technol. 203 (2010) 70–77.

[22] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, Stress results from two-dimensional

granular shear flow simulations using various collision models, Phys. Rev. E 71 (2005) 061307-

1-11.

[23] L. McElroy, J. Bao, C.T. Jayasundara , R.Y. Yang, A.B. Yu, A soft-sensor approach to

impact intensity prediction in stirred mills guided by DEM models, Powder Technol. 219 (2012)

151–157.

[24] C. Fan, X. T. Bi, J. R. Grace, A. Goto, Grid zone performance of a fluidized bed through

analysis of local solids holdup signals, Powder Technol. 219 (2012) 37–44.

[25] D. Ciampini, M. Papini, J. K. Spelt, Impact velocity measurement of media in a vibratory

finisher, J. Mater. Process Tech. 183(2007) 347-357.

[26] A. Yabuki, M. R. Baghbanan, J. K. Spelt, Contact forces and mechanisms in a vibratory

finisher, Wear 252 (2002) 635-643.

[27] S. Wang, R. S. Timsit, and J. K. Spelt, Experimental Investigation of Vibratory Finishing of

Aluminum, Wear 243 (2000) 147-156.

[28] J. Liu, J. R. Grace, and X. Bi, Novel Multifunctional Optical-Fiber Probe: I. Development

and Validation, AIChE Journal 49 (2003) 1405-1420.

[29] Laser displacement sensors manual, LK-G series, KEYENCE Corporation, Osaka, Japan,

2008.

[30] J. Krautkramer, H. Krautkramer, Ultrasonic Testing of Materials, fourth ed., Springer-

Verlag, Berlin, 1990, pp. 13-14, 533-534.

[31] T. G. Bali, The generalized extreme value distribution, Econ. Lett., 79 (2003) 423–427.

Page 64: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

46

[32] D. C. Montgomery, G.C. Runger, Applied Statistics and Probability for Engineers, fourth

ed., John Wiley and Sons Inc., United States, 2007.

Page 65: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

47

Chapter 3

3 Particle Impact Velocities in a Vibrationally-Fluidized Granular Flow: Measurements and Discrete Element Predictions

3.1 Introduction

The flow of granular materials has been studied in non-fluidized beds such as in hopper

discharge [1,2], rotating drums, conveyers, chutes and mixers [3,4], and in fluidized beds such

as in vibratory finishers and vibrating sieves [5-8]. A granular media becomes fluidized as the

vertical acceleration amplitude exceeds gravity, resulting in a marked decrease in the contact

pressure [8]. Vibratory finishing is widely used to polish, burnish, harden, and clean metal,

ceramic and plastic parts. In a tub vibratory finisher, the two-dimensional vibration of the walls

produces a vibrationally-fluidized circulatory bulk flow of the media. The media have both a

large-scale bulk flow velocity resulting from the resultant shear force with the walls, and a local

impact velocity during each vibration cycle [9,10].

The erosive wear and plastic deformation of a workpiece within a vibratory finisher are

determined mainly by the velocity, frequency, and direction of the impacts of the granular

finishing media [11,12]. Moreover, these impacts can lead to breakage of the finishing media.

In processes such as drug tablet coating within rotating drums, and bulk materials handling, the

impact of the granular products can contribute to their fragmentation. The machine components

may also be eroded in processes such as vibratory sieving and mixing because of large particle

impact velocities ‎[12].

Most research in the field of flowing granular media has focused on the bulk flow of granular

media inside different machines such as hopper discharge [1,2], rotating drums [3,4], and

vibratory finishers [7,8] . Generally, discrete element modeling (DEM) simulations [14, 15] give

Page 66: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

48

reasonable predictions of bulk flow velocity and volume fraction in both fluidized and non-

fluidized flows [16-18]. For instance, the behavior of the convection cells produced by a

spherical media within a vertically vibrated container has been studied using DEM and the

results were validated by the experimental data ‎[16]. As another example, the bulk flow velocity

distribution of glass beads in a vertical axis mixer with rotating flat blades was obtained in DEM

simulations and validated by the positron emission particle tracking (PEPT) measurements ‎[17].

Relatively few studies have dealt with the local behavior of the media in granular flows [16, 19-

23], and some of these studies have investigated vibrationally-fluidized flows [9, 11, 16, 23].

For instance, DEM was used to obtain the planar distributions of the collision velocity and

frequency of the granular media inside a horizontal rotating drum ‎[3]. In the same geometry,

DEM was used to predict the effects of particle stiffness, size and coefficient of restitution on the

solid fraction, collision frequency, and impact velocity inside the granular media ‎[18]. A

vibrated bubbling fluidized bed was simulated using DEM to give the instantaneous particle

velocity and local volume fraction ‎[11]. However, only in ref. ‎[19] were DEM predictions of

local behavior compared with experimental measurements. In that case, the local volume

fraction, granular temperature and bulk flow velocity in the surface layer of a shear flow were

measured using digital particle tracking velocimetry and compared with DEM predictions ‎[19].

Among the three different contact models used in these DEM simulations, only the large plastic

dissipation model predicted both the measured local volume fraction and bulk flow velocity with

reasonable accuracy (maximum error of 35%); however, the predictions of granular temperature

had an error of approximately 120% ‎[19]. A key objective of the present study was to compare

DEM predictions with local impact velocity measurements made previously in the tub vibratory

finisher using the submerged high-speed laser displacement probe of Fig. 1b, c ‎[9].

Although it has been shown that collision-level quantities are relatively insensitive to the choice

of the contact model used in DEM ‎[23], it is not known how sensitive the particle impact

velocities are to uncertainties in the contact coefficients used in these models, such as the

coefficients of restitution and friction corresponding to the particle-particle and the particle-wall

interactions. Therefore, it was of interest to begin the present work with an analytical and

numerical study of the sensitivities of the predicted bulk flow and impact velocities to these

contact coefficients used in the DEM. This was then used to guide the accuracy requirements for

the measurement of these coefficients for the spherical steel media used in the urethane-lined tub

Page 67: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

49

finisher with glass end walls. This was then followed by investigations of the relationship

between the DEM accuracy and the number of particle layers modelled between the glass end

walls of the tub finisher, and the use of a reduced shear modulus in order to decrease the model

run time. Finally, the DE model was expanded to mimic the complete experimental setup

including the submerged laser velocity probe used in the earlier experimental study of the bulk

flow and impact velocities [9, 10].

3.2 Discrete element modeling

3.2.1 Contact models and properties

In DEM, the contact forces between interacting particles and between the particles and the

container boundaries are calculated from the magnitude of the overlaps between the bodies in the

normal and tangential directions. The contact force models that can be used in DEM simulations

are defined using different assumptions about the elasticity of the contacting particles ‎[24].

Some of the models are elastic such as the linear-spring and Hertz models, while some are visco-

plastic like the linear-spring/dashpot (LSD), Kuwabara-Kono and Lee-Hermann models, and

some are elasto-plastic such as the Hertz-Mindlin, Thornton and Walton- Braun models [25, 26].

Most of these models are non-linear as they were defined based on the Hertz theory. In the

present simulations, since the amount of plastic deformation of the steel spheres was marginal,

the Hertz-Mindlin contact model was selected. In this model, both elastic and plastic

deformations of particles are calculated in the normal and tangential directions by considering

the coefficients of restitution, friction, and rolling resistance ‎[12]. The normal elastic force, e

nF ,

was calculated using the common Hertz formulas (Eq. 3.1) ‎[24].

3

* * 24

3

e

n nF E R (3.1a)

*

1 1 1

i jR R R (3.1b)

Page 68: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

50

22

*

111 ji

i j

vv

E E E

(3.1c)

where E* and R

* are the equivalent Young's modulus and the equivalent radius, respectively, and

the indices i and j indicate the particles in contact. δn is the normal deformation (overlap)

between the particles. The normal plastic (damping) force, d

nF , is determined from the

knowledge of the normal overlap and the normal coefficient of restitution, e, and the relative

normal velocity between particles before impact, rel

nv (Eq. 3.2) ‎[23]. The coefficient of

restitution is defined as the ratio of the relative normal velocity just after collision to that just

prior to the collision (Eq. 3.2e) ‎[23].

*52

6

d rel

n n nF S m v (3.2a)

*

1 1 1

i jm m m (3.2b)

* *2n nS E R (3.2c)

2 2

ln

e

ln e

(3.2d)

2 1

1 2

n n

n n

v ve

V V

(3.2e)

Here, m*

is the equivalent mass of the contacting particles, β is a function of the normal

coefficient of restitution, and Sn is called the normal stiffness. Vn and vn are the normal velocities

before and after the collision, respectively. The tangential elastic force, e

tF , is calculated using

tangential overlap, δt, and tangential stiffness, St (Eq. 3.3) ‎[23]. The tangential overlap is defined

as the relative displacement of the particles during impact in the tangential direction, having

subtracted the rolling contribution ‎[23]. G*=E

*/2(1+v)

is the equivalent shear modulus of the

contacting bodies.

e

t t tF S (3.3a)

* *2t tS G R (3.3b)

Page 69: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

51

The tangential plastic (damping) force, d

tF , is determined using the tangential overlap, the

tangential coefficient of restitution, et, and the relative tangential velocity between particles

before impact, rel

tv (Eq. 3.4) ‎[23].

*52

6

d rel

t t t tF S m v (3.4a)

2 2

ln

tt

t

e

ln e

(3.4b)

2 1

1 2

t tt

t t

v ve

V V

(3.4c)

The tangential force is constrained by the Coulomb friction force fs= µsFn such that:

e d

t t t sF F F f , where µs is the coefficient of static friction.

The present DEM simulations were made using EDEM 2.5 (DEM Solutions Inc. 2012) and its

standard library of contact models ‎[12]. The shear modulus, Poisson's ratio and density of the

steel balls, polyurethane wall of the tub and the glass end partitions were input to the software,

along with the particle-particle and particle-wall coefficients of restitution, friction and rolling

resistance, which were measured as described in Section 3.3. The glass partition at either end of

the tub was assumed to be frictionless. Table 3.1 gives the material properties used in the DEM.

The tub geometry was defined (Fig. 3.1a) and its motion was determined using the accelerometer

and laser displacement measurements described in ref. ‎[9]. The tub had sinusoidal translations in

the x and z directions of Fig. 3.1b and one sinusoidal rotation in the plane of Fig. 3.1b, all at 47

Hz. Table 3.2 gives the measured amplitudes of these vibration components at the tub centre of

gravity and the phase difference between the vertical and horizontal vibrations. The submerged

velocity probe shown in Fig. 3.1b, c could be oriented to the right, left or downward so that the

laser sensor measured particle impact velocities in these three directions.

Page 70: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

52

Table 3.1: Material properties used in the DEM simulations.

Material Shear modulus (GPa) Poisson's ratio Density (kg/m3)

Steel 76 0.29 7800

Polyurethane 0.0086 0.50 1200

Glass 26 0.23 2500

Table 3.2: Measured tub center of gravity vibration components used in DEM.

Horizontal (x)

displacement

amplitude (mm)

Vertical (z)

displacement

amplitude (mm)

Phase difference

between

translations (deg.)

Angular

displacement

amplitude (deg.)

Vibration

frequency

(Hz)

0.85 0.66 103 0.244 47

(a)

Page 71: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

53

Fig. 3.1: (a) Photo of the tub vibratory finisher including the laser probe. Schematic of the tub

vibratory finisher showing the bulk flow circulation direction (drawn to scale): (b) side view, (c)

plan view. The probe is shown at location H. The curved arrows show the flow streamlines

around the elliptical outer tube of the laser velocity probe. Dimensions in mm ‎[9].

(b)

(c)

Z

Page 72: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

54

The duration of an elastic impact between two bodies, Ti, was calculated based on the Hertz

theory from Eq. (3.5),

2

2

1

* 5

* *2.87i rel

n

mT

R E v

(3.5)

At least six time steps should be included during an impact in order to accurately solve the

equations of motion of the particles in contact ‎[12]. To make the solution stable, the simulation

time step was set to be 20% of the Rayleigh critical time step, ΔtR, defined as the time taken for a

shear wave to propagate through a solid particle. The Rayleigh critical time step is a function of

the particle diameter d, density ρ, shear modulus G and Poisson's ratio ν, and it was calculated

from Eq. (3.6) to be 3.43 µs ‎[12].

/ 2 / / 0.1631 0.8766Rt d G v

2 1

EG

v

(3.6)

The simulation time step was then calculated to be 0.686 µs for the 6.3 mm diameter steel balls

used in the experiments.

3.2.2 Reduced shear modulus

By decreasing the shear modulus used for the steel balls, the simulation time step could be

increased so that the total simulation run time was decreased, as indicated by Eq. (3.6). Previous

studies have used a reduced shear modulus in DEM to determine bulk-flow fields ‎[26], but it was

important in the present case to study the effect of the shear modulus reduction on the impact

velocities and forces. Considering the conservation of momentum of two particles during

impact, the particle velocity change is proportional to the product of the normal impact force (Fn)

and the impact time duration (Δt) (Eq. (3.7)). ΔVn and m are, respectively, the change in the

Page 73: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

55

normal velocity and the mass of a single particle in a binary impact. Fn is calculated using Eq.

(3.8) knowing the relative normal velocity of particles before impact, rel

nv ‎[23].

, Δ 0.2nn n n R

F tV v V t t

m

(3.7)

0.4

*0.6 1.2 0.4 1.220.1431 ( ) ( )

1

rel rel

n n n

G dF m V A G V

v

(3.8)

The coefficient A is constant, since all other quantities except rel

nV and G are constant. Equation

(3.9) gives the particle velocity change after one time step as derived by combining Eqs. (3.6-

3.8):

0.1

0.21 0.1102n n nv V VG

(3.9)

The parameter

0.1

G

was varied from 0.199 to 0.316 as the shear modulus was reduced by one

order of magnitude. Considering the approach velocities to be equal to the maximum values

observed in the experiments (0.5 m/s) [9], the relative error in the normal component of the

particle rebound velocity (vn) introduced by using the reduced modulus instead of the actual one

was less than 2%.

Equations (3.8) and (3.9) give the normal component of the impact force calculated using either

the actual shear modulus (Ga) or the reduced one (Gr), Fna and Fnr, respectively.

1.20.4 1.2 0.4 1.022na a na a nrF A G v A G V

0.4 1.20.4 1.2 1 10 1.044nr r nr a nrF A G v A G V

(3.10)

Fnr was found to be 2.5 times smaller than Fna due to the shear modulus reduction.

Therefore, although, the contact forces were decreased significantly as the shear modulus was

artificially reduced, the impact velocities, which were the main results to be obtained from the

present DEM, remained almost unchanged.

Page 74: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

56

In order to confirm that the impact velocities calculated using a reduced shear modulus for the

steel media were acceptably accurate, DEM simulations were run for the single particle-layer

model of Fig. 3.2 using the actual shear modulus and the shear modulus reduced by 1, 2 and 3

orders of magnitudes. The material properties and the tub motion given in Tables 3.1 and 3.2,

and the contact parameters provided in Table 3.7 (Section 3.4), were used in these simulations.

The model had the same bed depth as in the vibratory tub experiments of [9], with 2,500

particles introduced to the DEM at 100,000 particles/s at the beginning of the simulation. The

tub was then held stationary for 1 s for the particles to settle before the sinusoidal vibrations were

applied to the tub for 19 s.

Fig. 3.2: Snapshot of a single particle-layer simulation and the three measurement points located

at (x,z): RU (200, 300), H (100, 300), M (0, 280) (mm). At each location, the small and large

boxes were for impact velocity and bulk flow velocity calculation, respectively. Figure drawn to

scale. Straight line approximation to free-surface slope established at steady state. Arrow shows

average bulk flow direction.

The average impact and bulk flow velocities predicted by the DEM at points H and RU of Fig.

3.2 along with the corresponding total simulation run time are presented in Table 3.3. It was

observed that the calculated impact velocities changed less than 2% as the shear modulus was

H RU

M

Surface slope line

H2

z x

Page 75: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

57

decreased from its true value by a factor of 10 (Er-1), while the predicted bulk flow velocity

increased by 6.6%; however, the run time was reduced appreciably from 24 h to 10 h. These

errors increased quickly as the shear modulus was further reduced by factors of 100 (Er-2) and

1,000 (Er-3), making these additional decreases in run time impossible. Therefore, it was

concluded that the shear modulus could be safely reduced by one order of magnitude without

excessively compromising the accuracy of the predictions of either the impact or bulk flow

velocities. The simulation time step corresponding to this reduced shear modulus was 2.17 µs,

which was 9 times smaller than the elastic impact duration T, thereby satisfying the stability

condition mentioned in ‎[12]. The DEM data acquisition rate was 94 Hz, which was twice the

frequency of the tub vibration 47 Hz.

Table 3.3: Change in the x and z components of the impact velocity (denoted IV) and bulk flow

velocity (denoted BFV) at the points H and RU (Fig. 2.2) due to a shear modulus reduction of 1,

2 and 3 orders of magnitude (denoted G-1, G-2, G-3, respectively) and their corresponding errors

(Er). Velocities in mm/s. Root mean square error (RMS Er) is for all components and locations

of either IV or BFV. Time to run DEM simulation given in last row.

Actual

G

Reduced

G-1

Er-1

(%)

Reduced

G-2

Er-2

(%)

Reduced

G-3

Er-3

(%)

H-IVx 74.7 74.1 -0.8 62.7 -16 60.4 -19

H-IVz 90.9 91.2 0.3 90.5 -0.4 81.3 -11

RU-IVx 147.8 144.5 -2.2 136.7 -7.5 118.3 -20

RU-IVz 140.6 138.1 -1.8 126 -10 108.8 -23

RMS Er %

1.3

8.6

18

H-BFVx 12.8 12.8 0.0 13.4 4.7 4.2 -67

H-BFVz 18.8 16.2 -14 16.4 -13 6.6 -65

RU-BFVx 4.8 5.4 12 5.1 6.3 0.1 -98

RU-BFVz 19.3 19.3 0.0 16.1 -17 12.1 -37

RMS Er % 6.6 10 67

Run time (h) 24 10 5 1

Page 76: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

58

3.3 Sensitivity of DEM impact velocities to contact properties

The dependence of the bulk flow velocity and particle impact forces on coefficients of friction in

vibrationally-fluidized beds has been demonstrated in several earlier studies. Decreasing the

coefficient of friction between spherical steel and glass media, by applying a thin film of water,

lowered the bulk flow velocity significantly in the same tub finisher as in the present work ‎[8].

The bulk flow velocity also decreased when the coefficient of friction between the particles and

the tub walls was reduced by applying a smooth sheet of polytetrafluoroethylene ‎[8]. This is

consistent with the observation that the velocity of cylindrical work-pieces entrained in a bowl-

type finisher decreased when the media were made slightly wet ‎[27]. Measurements of

individual particle impact forces in a bowl-type vibratory finisher showed that reducing the

coefficient of friction by means of a thin film of water resulted in lower impact forces, implying

that the particle impact velocities were also reduced ‎[28]. Therefore, in order to establish the

accuracy of the DEM in predicting impact and bulk flow velocities, it was necessary to assess the

sensitivity of the predictions to uncertainties in the contact coefficients used in the DEM

simulations. This could then be used to guide the accuracy required in the experimental

measurement of the coefficients.

To begin, a pair of simple two-body collisions was analyzed: (1) the collision of two moving

disks, and (2) the collision of a moving disk with a moving wall. The procedure described in

Appendix 2-A was used to determine the sensitivity of the rebound velocities (normal, vn,

tangential, vt, and angular, ω) to the contact parameters. Based on the measurements described

in Section 3.4, the coefficient of restitution, e, was investigated over the range 0.4-0.95, the

coefficient of friction, µ, between 0.4 and 2.0, and coefficient of rolling resistance, µr, was varied

from 0.01 to 0.09, in both particle-particle (p-p) and particle-wall (p-w) interactions.

Table 3.4 gives the average percentage uncertainties in the rebound velocities due to 10%

uncertainty in each of the contact parameters calculated using the equations in Appendix 3-A.

Each uncertainty was normalized by the average rebound velocity calculated using the mean

values of the three contact parameters (e=0.675, µ=1.2 and µr=0.05). These average velocities

used in each normalization are given in Appendix 3-A.

Page 77: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

59

Table 3.4: Percentage uncertainty in the linear and angular rebound velocities due to 10%

uncertainty in each of the contact parameters.

vn (case 1) vt (cases 1) ω (case 1) vn (case 2) vt (cases 2) ω (case 2)

e 10 4 4 7 4 4

µ 0 10 10 0 9 10

µr 0 0 0.2 0 0.8 0.4

It is concluded that the normal rebound velocity, vn was only dependent on the coefficient of

restitution. The tangential velocity, vt and the angular velocity, ω were affected most by the

coefficient of friction. The angular velocity and the tangential velocity were weakly dependent

on the coefficient of rolling resistance. Overall, Table 3.4 illustrates that the uncertainties in the

rebound velocity components were relatively insensitive to uncertainties in the contact

parameters, being at most directly proportional to such uncertainty. It was also confirmed that

the rebound velocities varied directly with each of the three contact parameters (e.g. they

increased as each of the contact parameters increased), consistent with the observations of ‎[27]

and ‎[28].

These conclusions regarding the sensitivity of the rebound velocities to the coefficients of

restitution and friction were confirmed using the same single-layered DE model of the tub

described in Fig. 3.2. This model used the material properties given in Table 3.1 and the tub

motion defined in Tables 3.2. Design of experiments (DOE; Taguchi method) was used to

reduce the number of required runs of the model to 25 in order to investigate the relative

sensitivity of the velocity predictions to the 6 contact parameters under consideration ‎[29]. Five

values were selected for each coefficient within the same ranges assumed in the analytical study

(i.e. e: 0.4, 0.6, 0.8, 0.9, 0.95; µ: 0.4, 0.8, 1.2, 1.6, 2.0, and µr: 0.01, 0.03, 0.05, 0.07, 0.09). For

any particular combination of contact parameters within the interaction table, the average impact

and bulk flow velocities were calculated at points H and RU (Fig. 3.2) in the horizontal and

vertical directions using the method described in Section 3.2. Then, 25 combinations of the

contact parameters and the values corresponding to one of the velocity components were used to

calculate the sensitivities.

Page 78: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

60

Table 3.5 shows the uncertainties of the predicted impact and bulk flow velocities due to 10%

uncertainty in the contact parameters at location H. The uncertainty of each velocity component

was normalized by its mean value calculated from the DEM results (Table 3.5). In order to

combine the uncertainties in the x and z velocity components, the root mean square (RMS) values

of the uncertainties are also shown in Table 3.5. The coefficient of rolling resistance was not

included in this study because it does not contribute significantly to energy dissipation, and

Table 3.4 suggested its effect would be negligible as it is much smaller than the coefficient of

friction (less than 5%) ‎[17]. It is seen that the impact velocities predicted by the single-layer DE

model were quite insensitive to uncertainty in the contact parameters, being considerably smaller

than the specified 10% uncertainty in each contact parameter. This agrees with the conclusions

of Table 3.4 from the analytical study. The relatively small changes that did occur in the

predicted particle impact and bulk flow velocities varied directly with the coefficients of friction

and the coefficients of restitution of the particle-particle and particle-wall interactions. This is

consistent with the observations of refs. [8, 28, 29]. Similar results were obtained at location RU

of Fig. 3.2. As with the results of Table 3.4, it was confirmed that the impact velocities varied

directly with the contact parameters. This was also true for the bulk flow velocities, in agreement

with the experimental observations of ‎[27].

Table 3.5: Percentage uncertainty (U) in the single-layer DEM predictions of the impact and bulk

flow velocities due 10% uncertainty in the contact parameters normalized by the following mean

velocities from the DEM: IVx =75 mm/s, IVz =95 mm/s, BFVx =14 mm/s, BFVz =20 mm/s. p-p

particle-particle, p-w particle-wall. Location H of Fig. 3.2.

Impact velocity Bulk flow velocity

Contact

parameters Ux (%) Uz (%)

RMS

U(%) Ux (%) Uz (%)

RMS

U(%)

e (p-p) 0.6 0.4 0.5 5 3 4

µ (p-p) 0.2 0.3 0.3 1 4 3

e (p-w) 1 0.2 1 1 2 2

µ (p-w) 2 0.4 1 0.9 1 1

Page 79: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

61

3.4 Measurement of contact properties

A linear tribometer (Nanovea Inc. 2012 ‎[30]) was used to measure the coefficient of friction as a

function of sliding speed between a steel ball and the tub polyurethane wall, and between two

steel balls. In the first test, a 6.3 mm diameter ball of the steel media was fixed in the tribometer

as the stationary element against a flat sheet cut from the polyurethane tub wall that reciprocated

with an amplitude of 1 mm (2 mm peak-to-peak, sinusoidal oscillation). A weight of 1 N was

applied to the ball holder and a variety of reciprocating frequencies were tested to obtain a range

of sliding speeds up to 31 mm/s. The estimated maximum sliding speed at the wall in the

vibratory finisher was previously estimated to be 50 mm/s using data from a submerged high-

speed laser displacement probe [9, 10]. Each measurement involved 15 cycles and each

experiment was repeated 5 times on a different surface each time. Before each trial, the

contacting surfaces were wiped with a tissue soaked in ethanol. Table 3.6 shows the average and

maximum sliding speeds during the sinusoidal oscillation of the friction force corresponding to

each frequency, and the average and maximum µ during the cycle. The variation in µ with

sliding speed was negligible; therefore, the spindle speed was adjusted to be 5 rpm for the rest of

the experiments. This confirmed that a single value of µ could be used to model all impacts with

the tub walls.

Table 3.6: The coefficient of friction (µ) between the steel ball of finishing media and the sample

of the polyurethane tub wall as a function of sliding speed.

Spindle

speed (rpm)

Mean sliding

speed (mm/s)

Maximum sliding

speed (mm/s)

µ mean±95% confidence

interval (maximum value)

5 0.330 0.520 1.73±0.12 (1.88)

20 1.31 2.08 1.63±0.02 (1.65)

50 3.33 5.22 1.86±0.09 (1.94)

80 5.32 8.36 1.76±0.14 (1.90)

100 6.67 10.5 2.01±0.07 (2.07)

300 20.0 31.4 1.84±0.06 (1.91)

Average 1.80±0.06

Page 80: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

62

The measurement of the coefficient of friction between two steel balls was complicated by the

curved contacting surfaces during sliding. It was obtained by fitting a theoretical sliding force

versus time curve to the measured curve. One ball was fixed while the other was glued to the

stage reciprocating at the mean speed of 0.167 mm/s (spindle speed of 5 rpm). As in the

previous tests, 1 N was applied to the fixed ball. The upper ball vertical displacement, y, (Fig.

3.3) was measured using the tribometer depth sensor and the corresponding horizontal

displacement, x, and the inclination angle, θ, between the tangent to the contact point and the X-

axis, were calculated using Eqs. (3.11) and (3.12), respectively. The ball radius, R, was 3.15 mm

and the maximum depth at the end of each cycle was 30 µm giving a total arc length amplitude

of 0.5 mm.

22x R R y (3.11)

arccosR y

R

(3.12)

Fig. 3.3: Schematic showing the lower steel ball sliding against the upper steel ball that was

constrained in the y-direction to measure vertical displacement depth only. T is the measured

horizontal force. y=0 taken to be the minimum measured depth at the apex of the upper ball.

Page 81: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

63

Fig. 3.4: Tribometer measurements of the horizontal sliding force vs. (a) time and (b) angle (θ)

for steel balls moving against each other. The solid straight lines in (a) and (b) show the

analytical prediction.

The surface normal force, N, and the tangential friction force, f, were calculated using Eq. (3.13)

from the applied vertical force, W= 1 N, and the measured horizontal force, T. A horizontal

force signal acquired by the tribometer is shown in Fig. 3.4a. The sloped portions of the force

(a)

(b)

(b)

(a)

Page 82: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

64

signal corresponded to sliding while the vertical lines corresponded to the changes in the force

direction. The small fluctuations observed in the sloped portions were due to the surface

roughness, and were considered to be straight lines. Assuming that the upper ball moves to the

right relative to the lower ball as is shown in Fig. 3.3:

T fcos Nsin

W fsin Ncos (3.13)

considering θ to be positive; therefore, when the upper ball was to the left of the lower ball, θ

was negative value in Eq. (3.16). The coefficient of friction f

N , was expressed as a

function of T, W and θ as in Eq. (3.14). Rather than obtain µ directly from Eq. (3.14) using the

measured T as a function of time or θ, it was determined using a best fit of the predicted function

T(θ) given by Eq. (3.15) to the measured function, treating µ as an adjustable parameter (Fig.

3.4b).

Tcos Wsin

Tsin Wcos

(3.14)

cos sin

cos sinT W

(3.15)

The coefficients of restitution of a steel ball colliding with the polyurethane and of two

impacting steel balls were measured using the apparatus of Fig. 3.5. A solenoid actuator was

used to hold a steel ball above the urethane target (Fig. 2.5a) while a high-speed laser

displacement sensor (Keyence, Osaka, Japan, LK-G157) with an accuracy of 0.5 µm was used to

measure the impact and rebound velocities as in ‎[9]. The ball height was adjusted such that the

incident speed was 0.5 m/s as in ref. ‎[9]. Two laser sensors (LK-G82 and LK-G157 with

accuracies of 0.2 µm and 0.5 µm, respectively) were used to measure the incident and rebound

velocities of colliding steel balls as shown in Fig. 3.5b. One of the balls was left stationary on

the two glass guide rods while the other ball was released from rest through an inclined tube and

hit the stationary ball with an incident speed of 0.5 m/s. Equation (3.2e) and the conservation of

linear momentum were used to calculate the coefficient of restitution of the colliding steel balls.

Since the glass guide rods were very smooth, the falling ball tended mostly to slide into contact

Page 83: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

65

with the second ball. The possible influence of rolling was assessed by repeating the

measurement with the glass rods covered with tape so that the ball rolled. The coefficient of

restitution was unchanged. Table 3.7 shows the mean and 95% confidence intervals of the

coefficients of friction and restitution corresponding to the different material pairs for the 5

repeat experiments in each case. The friction results were consistent with the observation that

the coefficient of friction decreases as the hardness of the materials increases, because of reduced

plowing ‎[31]. As well, since the structural damping ratio (proportional to the collisional energy

dissipation ‎[32]) of polyurethane is about 3 times larger than that of steel ‎[32], steel balls

colliding with each other should indeed have a larger coefficient of restitution than a steel ball

colliding with a polyurethane surface.

Fig. 3.5: The drop test apparatus used to measure the coefficient of restitution of: (a) the steel

ball colliding with the polyurethane surface, (b) two colliding balls (side and plan views).

Table 3.7: The measured coefficients of friction and restitution for different pairs of materials

(mean ±95% confidence interval).

Materials e µ

Steel-Polyurethane (p-w) 0.662±0.004 1.81±0.170

Steel-Steel (p-p) 0.851±0.030 0.616±0.026

(b) (a)

Page 84: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

66

3.5 Results and discussion

3.5.1 Effect of model width

The present tub vibratory finisher produced a mostly two-dimensional oscillation and bulk flow

that was bounded at each end by a glass partition (Fig. 3.1). One of the objectives of the present

study was to assess the relation between the predicted DEM impact velocities and the width of

the model; i.e. the number of particle layers between the glass walls. In a series of models using

the material properties, tub motion and contact parameters provided in Tables 3.1, 3.2 and 3.7,

the glass partitions were separated by 1d, 2d, 4d, 8d and 12d where d is the particle diameter and

the instantaneous particle velocities (impact velocities) were acquired at each data time step in

two measurement bins, 9×9×6.3 mm (1.5d×1.5d×d) in the x, z and y dimensions, respectively),

located half way from the partitions at points H and M (Fig. 3.2). In the 1d simulation, a

clearance of 0.1 mm was added between the glass and the particles. These locations

corresponded to the locations of the laser probe used in ‎[9] to measure the impact velocities in

this same tub finisher. The size of the measurement bin had an insignificant effect on the impact

velocities recorded from the model, with a difference of less than 5% when the bin size was

increased to 3d×3d×d.

As seen in Fig. 3.2, the free surface of the vibrationally-fluidized bed was sloped at steady state

as a result of the bed expansion and bulk flow. The establishment of a constant free surface

slope was used to assess whether the DEM had reached steady-state, and to compare the flows

predicted by the models of different width in the y-direction. Table 3.8 shows the average and

95% confidence interval for slopes calculated every second from 10-20 s after the start of the

simulation. Since the maximum variation in the slope angle over this time interval was just 10%,

it was decided that these simulations had attained steady state, and that DEM data collection

could begin 9 s after the start of tub vibration for each of the models of different width.

Table 3.8 also shows that the mean slopes corresponding to the single layer and double layer

models were significantly greater than those of the 4, 8, 12 and 16-layer models. The main

reason for this was that the particles in the 1d and 2d models were prevented from moving freely

in the y-direction (perpendicular to the simulated glass partitions). Even though the tub

Page 85: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

67

vibrations and the dominant bulk flow were in the x-z plane, this constraint in the y-direction

produced considerable changes in the bulk flow and solid fraction compared with the wider

simulations in which the diffusion of particles between the layers was less restricted. This

constraint effect appeared to vanish for models of 4 layers and greater, as the predicted free-

surface slopes were nearly the same (Table 3.8).

Table 3.8: The average slope and ±95% confidence intervals of the free surface in models with

various numbers of particle layers in the width direction, y, between the glass partitions.

1 layer 2 layers 4 layers 8 layers 12 layers 16 layers

21±1.6 21±2.2 16±0.9 15±0.9 16±0.9 15±0.9

The mean impact velocity was calculated as the average of the absolute values of the

instantaneous velocities of the particles within the measurement bin acquired during 10 s of

simulation, beginning 9 s after the tub started to move. Extremely small particle velocities (less

than 1 mm/s or about 2% of the average impact velocity ‎[9]) were excluded since they were

unreliable and would have no influence on the surface finishing process.

Figure 3.6 shows the mean impact velocities of particles at points H, M and H2 (Fig. 3.2) in the x

(horizontal), z (vertical), and y (transverse) directions for different simulations with various

layers of particles in the transverse (y) direction. As the number of particle layers increased from

1-16, the average impact velocities in the three directions tended to become relatively constant.

This was expected as a consequence of the diminishing relative influence of the container side

walls on the particle behavior in the central part of the flow field. Since the tub motion was in

the x-z plane, the x and z impact velocities were much larger than those in the transverse y-

direction. The average impact velocities in the x-direction (Fig. 3.6a) decreased sharply from 1-4

layers at locations H and M, but not at the deeper location H2, which was 8 cm below point H.

There it was constant for all of the layer models, and significantly greater than at the two

shallower locations. In contrast, the impact velocity in the z-direction (Fig. 3.6b) displayed only

a small decrease with increasing numbers of layers, and the behavior was essentially the same at

all three locations. In the y-direction, the impact velocity increased sharply from 1-4 layers, but

then became approximately constant for the 8, 12 and 16-layer models (Fig. 3.6c). This result

Page 86: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

68

was a logical consequence of the reduced spatial confinement in the y direction as the number of

layers increased. However, the magnitude of the impact velocity in the y-direction depended on

the location in the flow, being appreciably greater at the deeper location H2, just as it was for the

x-component of impact velocity. The reason for this is unclear, and illustrates the variability of

the local impact conditions from point to point in the flow. This variability was also evident in

the experimental measurements of impact velocity that will be compared with DEM predictions

in the next section. These large variations in local impact velocity highlight the need for

accurate DE models capable of predicting the distribution of local impact conditions within a

vibratory finisher in order to understand how the degree of surface finishing changes within the

flow.

The average bulk flow velocity was obtained from the DEM using one of two procedures that

were shown to give equivalent results. The first measured the transit time for 20 particles across

a 20×20×6.3 mm (3d×3d×d) bin located adjacent to one of the glass partitions in either the H or

M locations in order to be consistent with the procedure that was used in [9] for the experimental

measurements of the bulk flow velocity using video recordings through the glass partitions. It is

noted that locating the bin half way between the glass partitions yielded the same results in all of

the simulations. The second procedure used the average of the instantaneous velocities of 60

particles inside the bins after assigning them a positive and negative sense so that the net bulk

flow was calculated. The average bulk flow velocities of the two approaches agreed to within

5%, and so the second procedure was adopted because it did not require the visual tracking of

particles entering and exiting the bins.

Figure 3.7a shows that the bulk flow velocity in the x-direction varied with location as expected

from the overall circulation in the tub, indicated by the arrow in Fig. 3.2; i.e. the largest x-

component at location M where the bulk flow is mostly parallel to the free surface, and the

smallest component at the location H2 where the flow is mostly vertical. The z-direction bulk

flow velocities of Fig. 3.7b show the expected corresponding trends in the vertical direction, with

the largest bulk flow velocity at location H2 and the smallest at location M. In all cases, the bulk

flow velocity became approximately constant beyond about 8 layers. The average bulk flow

velocity in the transverse (y) direction was zero for all of these simulations, consistent with the

symmetry of the model.

Page 87: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

69

Fig. 3.6: Predicted impact velocities in: (a) horizontal (x) direction, (b) vertical (z) direction, and

(c) transverse (y) direction, in the DEM simulations with different numbers of layers.

(a)

(c)

(b)

Page 88: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

70

Fig. 3.7: Predicted bulk flow velocities in (a) horizontal (x) direction and (b) vertical (z)

direction, in the DEM simulations with different numbers of layers.

3.5.2 Comparisons with immersed laser-probe velocity measurements

The dependence of the model results on the number of simulated layers, and the potential

influence of the immersed laser probe on the bulk flow and impact velocities, motivated the DE

modelling of the actual experimental setup of ref. [9], including the immersed laser probe as

shown in Fig. 3.8. In order to decrease the simulation time, only half of the geometry in the y-

direction was modeled (105 mm; i.e. 16 particle diameters), using a smooth, frictionless

symmetry boundary made of the same material as the particles ‎[33]. The model was, however,

insensitive to this friction coefficient, and essentially the same results were obtained if the

symmetry boundary was assigned the coefficient of friction measured for particle-particle

contact (Table 3.6).

(a) (b)

Page 89: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

71

Fig. 3.8: Snapshot of the simulation constructed using the plane of symmetry through the

elliptical laser velocity probe used in [9] located at point M (as if measuring velocities of

particles approaching from the right). (a) side view, (b) plan view, (c) side view of probe and the

measurement bins on the probe surface oriented with the laser window to the right. The large bin

with solid boundaries and the small bin with dotted boundaries were used to obtain the impact

and bulk flow velocities, respectively. (d) Side view of a portion of the probe surface showing

the particle bulk flow with the measurement bin in the downward directions. In (c) and (d), large

arrows show the bulk flow direction, dashed smaller ones represent the laser beam.

Five models were constructed to mimic the experimental measurement locations and orientations

used in ref. [9] where the probe was located at either position H or M (Fig. 3.2) and then

positioned to record particle velocities normal to the plane of the laser window in the leftward,

(b)

(d)

Y

x

z x

(a)

(c)

Page 90: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

72

rightward and downward directions; i.e. H-rightward, leftward, downward; M-rightward and

downward. These models used the material properties, tub motion and contact parameters given

in Tables 3.1, 3.2 and 3.7, respectively. In each simulation, the impact velocities of the particles

colliding with the probe, normal to the laser window of the probe, in a d×3d×d bin adjacent to

the probe with its 3d side parallel to the window (Fig. 3.8c) were acquired in 10 s and then

averaged. The mean bulk flow velocity in the DEM was calculated as the average velocity

(considering both positive and negative signs) of particles moving in a d×d×d bin adjacent to the

probe and parallel to its surface (Fig. 3.8c). The bulk flow velocity measurement bin was

smaller to have only one particle in the bin at each time step. This was consistent with the

experimental measurements where the bulk flow velocity was calculated considering only the

laser reflections from the particle layer immediately adjacent to the probe ‎[9].

In the experiments of ref. [9], although the laser was stationary, the external protective shell and

glass laser window of the submerged probe experienced some vibration with a velocity between

20 to 45 mm/s, depending on the location in the flow ‎[9]. Consequently, the average impact

velocity was determined considering only those particles with velocities normal to the window

were greater than that of the shell of the probe ‎[9]. In the DEM simulations, the probe was fixed,

but the same velocity filter was applied to the model data to be consistent with the experimental

procedure; i.e. only surface-normal particle velocities greater than that of the probe.

Table 2.9 shows that the predicted impact and bulk flow velocities for the five DEM simulations

agreed quite well with the measured values from ref. [9]. At point H, the DEM impact velocities

overestimated the measured values by between 10% and 19%, while at point M the impact

velocities were underestimated by less than 7%. One cause for these differences was the

sensitivity of the impact velocities measured in [9] to the laser probe depth relative to the free

surface of the flow. For example, shifting the probe 1 cm upward produced approximately a

25% change in the impact velocity, while moving it down 1 cm, changed it by about 15%.

Fig. 3.9 shows the log normal probability density distributions of the impact velocities calculated

using the DEM results and the experimental data of [9]. It is seen that the numerical and

experimental distributions were very similar in most of their main features such as their mean

values and skewness. However, the DEM probability distributions were wider than those

determined experimentally (Fig. 3.9).

Page 91: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

73

Table 3.9: Comparison of the predicted and measured impact velocities (±95% confidence

intervals based on at least 200 data points from the DEM and at least 1,000 laser-probe velocity

measurements; mm/s) and bulk flow velocities (mean value; mm/s) at locations H and M in

different directions (Figs. 3.2 and 3.8). The bracketed bulk flow velocity at location M-

rightwards refers to the DEM absolute oscillating velocity.

Measurement

locations and

directions

Impact

velocity

(DEM)

Impact

velocity (Expt

‎[9])

Diff

(%)

Bulk flow

velocity

(DEM)

Bulk flow

velocity (Expt

‎[9])

Diff (%)

H-rightwards Vx =86±4.5 Vx=74±1.1 +16 Vz =12 Vz =13 -7.7

H-downwards Vz =95±4.6 Vz=80±1.3 +19 Vx=15 Vx=16 -6.3

H-leftwards Vx=70±5.4 Vx=61±1.0 +10 Vz =15 Vz =20 -25

M-rightwards Vx=85±3.8 Vx=91±1.2 -6.6 Vz =5 (18) Vz =19 -74%

(-5.5%)

M-downwards Vz =110±7.7 Vz =110±1.6 0 Vx=14 Vx=20 -30%

Page 92: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

74

0 20 40 60 80 100 120 140 160 180 2000

0.05

0.1

0.15

0.2

Impact velocity (mm/s)

Pro

ba

bili

ty d

en

sity d

istr

ibu

tio

n

H-rightward-DEM

H-rightward-exp

H-leftward-DEM

H-leftward-exp

M-rightward-DEM

M-rightward-exp

Fig. 3.9: Probability density distributions of impact velocities predicted by the DEM and

measured in the experiments of ref. ‎[9]. All the distributions were fitted with log-normal

functions. Locations defined in Figs. 3.2 and 3.8.

The predicted mean bulk flow velocities at point H in different directions and at point M-

downwards in the x-direction were reasonably close to the measured values, underestimating

them by between 6% and 30%. However, the bulk flow velocity in the z-direction at point M-

rightwards was apparently underestimated by 74%. In reality, this DEM prediction was probably

more accurate than the measurement due to a limitation of the laser sensor velocity data which

did not distinguish the direction of ball travel parallel to the laser window. In other words, all of

the velocities from the laser displacement signals in ref. ‎[9] were assumed to be in the same bulk

flow direction, which was correct in most cases. However, in locations where the bulk flow was

weak and the balls were moving in more random patterns, this assumption would lead to an

artificially high apparent bulk flow velocity, since the sign of ball velocity was not taken into

account. This was indeed the case at location M-rightward, where the particles oscillated back

and forth in the z-direction as the flow collided with the upstream side of the elliptical shell of

Page 93: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

75

the laser probe. Therefore, the laser probe likely overestimated the bulk flow velocity in this

case. A better simulation of these erroneous laser measurements at location M-rightwards could

be obtained by taking the average of the DEM absolute oscillating velocities. Doing this, the

average absolute oscillating z-direction velocity predicted by the DEM for the M-rightward

experiment was 18 mm/s, as shown in brackets in Table 3.9. This agreed well with the laser

velocity measurement of 19 mm/s [9]. However, the absolute oscillating velocities at point H in

different directions and at point M-downwards in the x-direction were the same as the bulk flow

calculated using signed velocities, and so the differences between the measured and predicted

bulk flow velocities in these cases was due to other causes.

3.6 Conclusions

The impact and bulk flow velocities in a tub vibratory finisher predicted by discrete element

modeling (DEM) were compared to the measured values for steel granular media. This was the

first time measurements of local impact velocities have been compared with DEM predictions in

a vibrationally-fluidized granular flow. The sensitivity of the predicted local impact velocities

and the bulk flow velocities to the DEM contact parameters (coefficients of friction, restitution

and rolling resistance) was investigated both analytically and numerically. It was found that the

predictions of both the bulk flow and impact velocities were relatively insensitive to

uncertainties in these impact coefficients.

The vibration of the tub finisher walls was essentially two-dimensional, with negligible motion

in the transverse direction. The effect of the number of DE particle layers in the transverse

direction was investigated to determine the minimum number required for accurate predictions of

the particle impact and bulk flow velocities. It was concluded that accurate velocities could only

be predicted using at least 4-12 layers. Fewer layers created too much spatial constraint to allow

realistic particle motion. Moreover, a reduced shear modulus could be used to decrease the

model run time without significantly affecting the accuracy of the predicted impact and bulk

flow velocities.

Page 94: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

76

The complete DE model simulated the actual setup used in an earlier experimental study that

measured the bulk flow and local impact velocities in the tub vibratory finisher. The coefficients

of friction and restitution between the steel balls, and between the balls and the tub walls, were

measured in order to obtain the most accurate DE simulations. The predicted impact velocities in

several directions and at two locations in the tub displayed the same log-normal probability

density distributions as did the previously measured impact velocity distributions. Moreover, the

predicted mean impact velocities were in good agreement with the experimental measurements,

with a maximum error of 19%. The DE predictions of the mean bulk flow velocities were also

consistent with the experimental measurements, having a maximum error of 30%.

Therefore, it was concluded that DEM can be used to give reasonably accurate predictions of

both the local impact velocities and the bulk flow of particles in vibrationally-fluidized beds.

This will be useful in predicting the impact energy and force of the media on workpiece surfaces,

and hence the resulting wear and surface deformation during vibratory finishing.

3.7 References

[1] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, A. Kong, P. J. Narayan, B. C. Hancock,

Granular segregation in discharging cylindrical hoppers: A discrete element and experimental

study, Chem. Eng. Sci. 62 (2007) 6423 – 6439.

[2] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, B. C. Hancock, Predicting the flow mode

from hoppers using the discrete element method, Powder Technol. 195 (2009) 1–10.

[3] R.Y. Yang, R.P. Zou, A.B. Yu, Microdynamic analysis of particle flow in a horizontal

rotating drum, Powder Technol. 130 (2003) 138–146.

[4] V. Jasti, C. F. Higgs, Experimental study of granular flows in a rough annular shear cell,

Phys. Rev. E 78 (2008) 041306-1-8.

[5] N.G. Deen, M. Van Sint Annaland, M.A. Van der Hoef, J. A. M. Kuipers, Review of discrete

particle modeling of fluidized beds, Chem. Eng. Sci. 62 (2007) 28-44.

[6] X. Z. An, C.X. Li, R.Y. Yang, R.P. Zou, A.B. Yu, Experimental study of the packing of

mono-sized spheres subjected to one-dimensional vibration, Powder Technol. 196 (2009) 50–

55.

Page 95: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

77

[7] S.E. Naeini, J.K. Spelt, Two-dimensional discrete element modeling of spherical steel media

in a vibrating bed, Powder Technol. 195 (2009) 83-90.

[8] S.E. Naeini, J.K. Spelt, Development of single-cell bulk circulation in granular media in a

vibrating bed, Powder Technol. 211 (2011) 176-186.

[9] K. Hashemnia, A. Mohajerani, J. K. Spelt, Development of a laser displacement probe to

measure particle impact velocities in vibrationally fluidized granular flows, Powder Technol.,

235 (2013) 940-952.

[10] D. Ciampini, M. Papini, J. K. Spelt, Impact velocity measurement of media in a vibratory

finisher, J. Mater. Process Tech. 183 (2007) 347-357.

[11] L. Xiang, W. Shuyan, L. Huilin, L. Goudong, C. Juhui, L. Yikun, Numerical simulation of

particle motion in vibrated fluidized beds, Powder Technol. 197 (2010) 25–35.

[12] EDEM‎software’s‎website,‎http://www.dem-solutions.com/academic, 2012.

[13] A. Munjiza, The combined finite discrete element method, John Wiley & Sons, Ltd,

Chichester, West Sussex, England, 2004.

[14] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of particulate

systems: A review of major applications and findings, Chem. Eng, Sci. 63 (2008) 5728-

5770.

[15] C.H. Tai, S.S. Hsiau, C.A. Kruelle, Density segregation in a vertically vibrated granular bed,

Powder Technol. 204 (2010) 255-262.

[16] M. Majid, P. Walzel, Convection and segregation in vertically vibrated granular beds,

Powder Technol. 192 (2009) 311-317.

[17] R. L. Stewart, J. Bridgwater, Y. C. Zhou, A.B.Yu, Simulated and measured flow of granules

in a bladed mixer - A detailed comparison, Chem. Eng. Sci56 (2001) 5457-5471.

[18] B. Freireich, J. Litster, C. Wassgren, Using the discrete element method to predict collision-

scale behavior: A sensitivity analysis, Chem. Eng. Sci. 64 (2009) 3407- 3416.

[19] J.J. McCarthy, V. Jasti, M. Marinack, C.F. Higgs, Quantitative validation of the discrete

element method using an annular shear cell, Powder Technol. 203 (2010) 70-77.

[20] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, Stress results from two-dimensional

granular shear flow simulations using various collision models, Phys. Rev. E 71 (2005)

061307-1-11.

Page 96: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

78

[21] L. McElroy, J. Bao, C.T. Jayasundara , R.Y. Yang, A.B. Yu, A soft-sensor approach to

impact intensity prediction in stirred mills guided by DEM models, Powder Technol. 219

(2012) 151–157.

[22] C. Fan, X. T. Bi, J. R. Grace, A. Goto, Grid zone performance of a fluidized bed through

analysis of local solids holdup signals, Powder Technol. 219 (2012) 37–44.

[23] A. Di Renzo, F. P. Di Maio, Comparison of contact-force models for the simulation of

collisions in DEM-based granular flow codes, Chem. Eng. Sci. 59 (2004) 525 – 541.

[24] A.B. Stevens, C.M. Hrenya, Comparison of soft-sphere models to measurements of collision

properties during normal impacts, Powder Technol. 154 (2005) 99 –109.

[25] J. Ai, J. F. Chen, J. M. Rotter, J. Y. Ooi, Assessment of rolling resistance models in discrete

element simulations, Powder Technol. 206 (2011) 269–282.

[26] T. N. Tang, Input parameters of discrete element methods, J. Eng. Mech., 132 (2006) 723-

729.

[27] S. Wang, R. S. Timsit, and J. K. Spelt, Experimental investigation of vibratory finishing of

aluminum, Wear 243 (2000) 147-156.

[28] A. Yabuki, M.R. Baghbanan, J.K. Spelt, Contact forces and mechanisms in a vibratory

finisher, Wear 252 (2002) 635–643.

[29] R. K. Roy, A primer on the Taguchi method, 2nd edition, Society of Manufacturing

Engineers, New York, 2010.

[30] Tribometer, Nanovea Inc. website, http://www.nanovea.com/Tribometers, Irvine, California,

US, 2012.

[31] W. Ni, Y. Cheng, M. J. Lukitsch, A. M. Weiner, L. C. Lev, D. S. Grummon, Effects of the

ratio‎of‎hardness‎to‎Young’s‎modulus‎on the friction and wear behavior of bilayer coatings,

Appl. Phys. Lett., 85 (18) (2004) 4028-4030.

[32] H. Bachmann, W. J. Ammann, F. Deischl, J. Eisenmann, I. Floegl, G. H. Hirsch, G. K.

Klein, G. J. Lande, O. Mahrenholtz, H. G. Natke, H. Nussbaumer, A. J. Pretlove, J. H.

Rainer, E. U. Saemann, L. Steinbeisser, Vibration problems in structures: Practical

guidlines, Birkhauser Verlag, Berlin, 1995.

[33] M. H. Sadd, Elasticity: Theory, applications, and numerics, 2nd edition, Academic Press,

Malden, Massachusetts, 2009.

[34] R.M. Brach, Friction, restitution and energy loss in planer impacts, J. Appl. Mech., 51

(1984) 164-170.

Page 97: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

79

Chapter 4

4 Finite Element Continuum Modeling of Vibrationally-Fluidized Granular Flows

Nomenclature

Dij Rate of strain tensor Vs Speed of sound in air

dev

ijD Deviator rate of strain tensor Kb Media bulk modulus

σij Stress tensor ν Poisson's ratio

dev

ij Deviator stress tensor Eb Media equivalent Young's

modulus

σ1 Maximum principal stress Ew Wall Young's modulus

σ3 Minimum principal stress Lb Effective thickness of the media

Shear rate Lw Wall thickness

η Shear viscosity Kmedia Media stiffness

µ Coefficient of internal

friction Kwall Wall stiffness

I Inertial number K Effective contact stiffness

ρs Solid phase density Ft/Fn Shear to normal force ratio

d Particle diameter µmw

Media-wall coefficient of

friction

p Local pressure u'i, v

'i, w

'i Particle velocity fluctuations

Δ Dilatation ui, vi, wi Particle instantaneous velocity

( , )k l

iF Contact force vector ū, ῡ, w Particle average velocity

( , )k l

jx Relative center-to-center

distance of particles l Mean free path

Page 98: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

80

vi Particle velocity vector e Coefficient of restitution

Mk Mass of particle k fi Body force vector

V Measurement bin volume N Number of particles in a bin

Lij Velocity gradient tensor ui Displacement vector

φ Volume fraction T Granular temperature

Ud Sound wave propagation

speed V Average bulk flow velocity

ρb Bulk density Vimp Average local impact velocity

ρg Gas phase density

4.1 Introduction

The flow of granular materials has been studied in non-fluidized beds such as in hopper

discharge [1, 2], rotating drums, conveyers, chutes and mixers [3, 4], and in fluidized beds such

as in vibratory finishers and vibrating sieves [5-8]. Vibratory finishing is widely used to polish,

burnish, harden, and clean metal, ceramic and plastic parts. In a tub vibratory finisher, the two-

dimensional vibration of the walls produces a vibrationally-fluidized circulatory bulk flow of the

media (Fig. 4.1). The granular media becomes fluidized as the vertical acceleration amplitude

exceeds that of gravity, resulting in a marked decrease in the contact pressure ‎[8]. The media

have both a large-scale bulk flow velocity resulting from the resultant shear force with the walls,

and a local impact velocity during each vibration cycle ‎[9].

The erosive wear and plastic deformation of a workpiece within a vibratory finisher are

determined mainly by the velocity, frequency, and direction of the impacts of the granular

finishing media [10-12]. Moreover, these impacts can lead to breakage of the finishing media.

In processes such as drug tablet coating within rotating drums, and bulk materials handling, the

impact of the granular products can contribute to their fragmentation ‎[12].

Generally, discrete element modeling (DEM) simulations [14, 15] give reasonable predictions of

bulk flow velocity and volume fraction in both fluidized and non-fluidized granular flows [16-

18]. The local behavior of the non-fluidized granular flows has also been investigated

Page 99: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

81

experimentally and numerically [2, 21-23], while some studies have examined vibrationally-

fluidized flows [9-12, 19, 20]. Recently, it has been shown that DEM predictions of both bulk

flow and local impact velocities in a tub vibratory finisher agreed well with measurements made

using a high-speed laser displacement probe immersed in the flow ‎[19].

Previous efforts to model granular flows using a continuum approach have focused on the bulk

flow behavior of the non-fluidized granular media such as in flow down inclined chutes [24, 26,

30], plane shear flow [28, 29], flow in rotating drums [27, 28], flow in annular shear cells [24,

27, 28], and silo discharge flows [25, 27, 28]. In these cases of non-fluidized flows, constitutive

equations were defined to describe the bulk flow of the granular materials under quasi-static and

moderate flow (liquid-like) regions [24, 26, 30]. To‎ the‎ author’s‎ knowledge,‎ no‎ papers‎ have‎

been published on continuum modelling of vibrationally-fluidized granular flows. The

continuum models that have been used for quasi-static and moderate flow include different

elasto-plastic formulations of the equivalent continuum media [24, 30, 32], and visco-plastic

formulations considering only the plastic behavior of the equivalent media under time-varying

shear deformation [26-27]. In many cases, DEM was used to obtain the equivalent properties

needed to model the granular flows as a continuum [25, 27-29, 31, 32]. For example, the

equivalent stress tensor, pressure and shear rate at different points of a flowing granular bed were

obtained using DEM, and were then used to determine the equivalent continuum media elastic

and plastic properties [25, 30, 31]. Generally, the average streamlines, and hence the bulk flow

behavior, determined through these continuum simulations have been in fairly good agreement

with the predictions of discrete element modeling [24, 27, 30]. For example, Kamrin [24]

proposed an elasto-plastic constitutive law for use in a Lagrangian finite element model of

granular flows in inclined chutes, rectangular silos and annular Couette cells. The predicted flow

fields were compared with those calculated using DEM. Andrade et. al [30, 31] used the same

material law to model the static 3D compression of sand particles. Forterre and Pouliquen [27]

used a visco-plastic constitutive law with the fixed-grid finite difference method, which is

equivalent to an Eulerian mesh to simulate granular flows in the geometries of [24] and in

rotating drums. They compared their results with experimental measurements and DEM

predictions of bulk flow [27]. The Lagrangian formulation has been used in most of the papers

that modelled granular flows using the finite element method [24, 32].

Page 100: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

82

The present work modelled the tub vibratory finisher shown in Fig. 4.1a filled with 6.3 mm

diameter steel balls. Its motion was determined using the accelerometer and laser displacement

measurements as described in ref. [9]. The tub had sinusoidal translations in x and z directions

and sinusoidal rotation in the plane of Fig. 4.1b, all at 47 Hz. The equivalent material properties

of the granular media needed in modeling the media as a continuum were estimated using grain-

scale 3D DEM simulations.

The vibrationally-fluidized granular flow was then modeled as a continuum visco-plastic media

using both Lagrangian and Eulerian finite element formulations ‎[34]. The bulk flow velocities

were compared with those obtained in DEM simulations and with experimental measurements

made in the same tub finisher [9, 19].

Page 101: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

83

Fig. 4.1: (a) Photograph of the tub vibratory finisher including the laser probe used to measure

the media impact velocities in ref. [9], and two glass partitions within the tub. (b) Schematic of

the tub vibratory finisher showing the bulk flow circulation direction from the side view.

Dimensions in mm ‎[19].

(b)

Z

(a)

Page 102: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

84

4.2 Continuum constitutive equations

Since the flow in the present vibrationally-fluidized bed was in the moderate regime as will be

discussed below, the visco-plastic model of ref. ‎[24] was used, as given by Eq. (4.1). This visco-

plastic constitutive equation expresses the stress tensor, σij, in terms of the rate of strain tensor,

Dij, and the pressure, p, which is equal to the mean value of the diagonal components of the

stress tensor (Eq. (4.2a)) [24-26]. The bulk modulus of elasticity, Kb, was obtained by dividing p

by one third of the dilatation, Δ, which is the trace of the strain tensor (Eq. (4.2b)) ‎[35]. Kb is

also related to the Young's modulus, Eb, and Poisson's ratio, v, of the continuum material

representing the granular media ‎[35]. The shear rate, , is defined as the norm of the deviator

strain rate tensor dev

ijD (Eq. (4.3)), and the shear viscosity, η, defined in Eq. (4.4), is a function

of the pressure, the shear rate, and the internal coefficient of friction, µ, which was defined as the

ratio of the maximum shear stress to the mean normal stress (Eq. (4.5a)), where σ1 and σ3 are the

maximum and minimum principal stresses, respectively ‎[25][24, 25]. It is noted that µ could

also be calculated using the Drucker-Prager criterion (Eq. (4.5b)) where dev

ij is the norm of the

deviator stress tensor.

The inertial number, I, defined in Eq. (4.6) captures the effects of the shear rate and the

hydrostatic pressure, and the media properties ρs and d, the particle density and diameter,

respectively. The inertial number reflects the relative magnitudes of inertial forces (dependant

on the shear rate) and confining forces (dependant on the pressure) [28, 29]. It can also be

interpreted as the ratio of two time scales: (i) a microscopic time scale of particle

rearrangements,/ s

d

p , which represents the time taken for a particle to move a distance d

under the pressure p, and (ii) a macroscopic time scale, 1/ , corresponding to the mean shear

deformation [27, 28]. Small values of I correspond to a quasi-static regime where particle

inertial forces are small compared to confining forces, whereas large values of I correspond to

rapid flows with high shear rates. Therefore, the coefficient of internal friction, µ, is a function

of I as illustrated in Fig. 4.2a [24, 26, 29]. The data presented of Fig. 4.2 were obtained from a

Page 103: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

85

2D shear test in which the particles between two parallel planes are confined by a normal force

and sheared by moving one of the planes ‎[29].

( , ) dev

ij ij ij ij ijp p p D (4.1)

1

3iip (4.2a)

1; 3(1 2 )

3b b bp K K v E (4.2b)

3 3

1 1

1

2

dev dev dev

ij ij ij

i j

D D D

(4.3)

( , )p

p

(4.4)

1 3

1 3

(4.5a)

2

dev

ij

p

,

3 3

1 1

1

2

dev dev dev

ij ij ij

i j

(4.5b)

/ s

dI

p

(4.6)

In the quasi-static and moderate flow regimes up to I=0.01, the coefficient of internal friction, μ,

and the volume fraction are often modeled as being independent of I ‎[24], as shown in Fig. 4.2a.

As will be explained in Section 4.3, DE modelling of the present vibrationally-fluidized granular

media in the tub finisher (Fig. 4.1) revealed that I was smaller than 0.012 throughout the flow.

This also permitted the assumption of a constant volume fraction, since Fig. 4.2b shows that φ is

approximately constant if I is less than about 0.1.

Two stress components contribute to the stress tensor σij in granular flows, as shown by Eq.

(4.7a): the collisional stress due to contact forces and the kinetic (or streaming) stress due to

momentum transfer by particle diffusion [24, 36]. The kinetic stress tensor is produced by the

transfer of momentum by particle diffusion through the bulk material, and is appreciable in

highly agitated flows with relatively large mean free paths [36]. Equation (4.7a) expresses these

stresses in terms of ( , )k l

iF and ( , )k l

jx which are i-component of the contact force vector and j-

Page 104: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

86

component of the center-to-center vector corresponding to the contact of particle k and its

neighbor particle l. vi, Mk and V are, respectively, the velocity i-component, the mass of particle

k, and the volume of the measurement bin. In most quasi-static and moderate granular flows, the

kinetic stress is much smaller than the collisional stress since the momentum of most of the

particles is transferred by particle impact rather than the particle diffusion. Therefore, most

previous studies have ignored the second term of Eq. (4.7a), thereby reducing computational

time [24, 25]. The same assumption was made in the current study leading to (Eq. (4.7b)). For

highly agitated flows, this assumption cannot be made and the kinetic stress becomes dominant.

( , ) ( , ) ( ) ( )

1 1 1

1( ) ( )

n m nk l k l k k

ij i j k i j

k l k

F x M v vV

(4.7a)

( , ) ( , )

1 1

1( )

n mk l k l

ij i j

k l

F xV

(4.7b)

( , ) ( , )

1 1

1( ( ))

n mk l k l

ij i j

k l

sym F xV

(4.7c)

The stress tensors calculated using Eq. (4.7b) were asymmetric because, in granular flows, local

body moments are produced everywhere in the media due to the frictional forces acting between

the contacting particles [36-39]. Therefore, only the symmetric part of the stress tensor was

considered when determining the principal stresses and the coefficient of internal friction µ (Eq.

(4.7c)) in the vibrationally-fluidized granular flow under investigation.

Page 105: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

87

Fig. 4.2: Variation of (a) the equivalent coefficient of internal friction ‎[29] and (b) the volume

fraction with the inertial number seen in simple 2D shear test ‎[28].

The stress tensor from Eq. (4.1) was then used to define the equation of motion of the equivalent

continuum material as shown in Eq. (4.8), where fi and ui are the body force and displacement

vectors, respectively, ρb is the media equivalent bulk density, and xj and t are the position vector

and time, respectively ‎[35]. The media equivalent bulk density is the product of the particle

density and the average volume fraction, φ which was defined as the total volume of particles in

a measurement bin divided by the volume of the bin. As discussed above, φ, is constant except

in the collisional regime where it decreases due to highly energetic impacts (Fig. 4.2b) ‎[25][24,

28]. Equation (4.8) was then solved using the finite element method.

2

2

ij ii b

j

uf

x t

(4.8)

(a) (b)

µ

φ

I

I

Collisional

Quasi-static Moderate

Page 106: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

88

4.3 Determining the continuum model parameters

4.3.1 Coefficient of internal friction and volume fraction

A three-dimensional DE model of the vibratory finisher with the glass partitions separated by a

distance equal to 4 particle diameters was used to obtain the equivalent continuum media

properties used in the constitutive model (Eq. (4.1)).

The present DEM simulations were made using EDEM 2.5 (DEM Solutions Inc. 2014) and its

standard library of contact models ‎[12] following the procedure used in the ref. ‎[19] The DEM

input parameters including the coefficients of friction and restitution, and the tub vibration

amplitudes were the same as those measured and used in ‎[19], and are listed in Tables 4.1 and

4.2. The only departure from the procedures used in ref. [19] was that the actual shear modulus

of the steel balls was used in the present simulations in order to predict accurate impact forces.

In contrast, it was possible to use a reduced ball shear modulus in [19] to decrease the

computational effort since only the impact velocities were of interest. The appropriate simulation

time step was calculated to be 0.686 µs for the 6.3 mm diameter steel balls ‎[19]. An 8-layer DE

model was also simulated to assure that the results were not dependent on the distance between

the glass partitions. Figure 4.3 shows the granular flow streamlines (trajectories) of the particles

moving in the first layer of the 4-layer DE model during a 1 s interval. The shear rate and the

stress tensor were calculated at four points in the flow (H1, H2, L1 and L2 in Fig. 4.4) to

determine the average equivalent continuum properties. At each of these points a 3d×3d×1.5d

measurement bin was defined at the mid-plane between the glass partitions to measure the stress

tensor and consequently, the equivalent coefficient of internal friction (Eq. (4.5)) and the local

pressure (Eq. (4.2)).

Page 107: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

89

Fig. 4.3: The average streamlines of particles moving in a counter-clockwise flow in the first

layer of the 4-layer DE model during a 1 s interval.

Table 4.1: Material properties used in the DEM simulations ‎[19].

Material Shear modulus (GPa) Poisson's ratio Density (kg/m3)

Steel 76 0.29 7800

Polyurethane 0.0086 0.50 1200

Glass 26 0.23 2500

Table 4.2: Measured tub center of gravity vibration components used in DEM ‎[19].

Horizontal (x)

displacement

amplitude (mm)

Vertical (z)

displacement

amplitude (mm)

Phase difference

between x and z

translations (deg.)

Angular

displacement

amplitude (deg.)

Vibration

frequency

(Hz)

0.85 0.66 103 0.244 47

z

x

Page 108: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

90

The stress tensor in the granular media was calculated according to Eq. (4.7b) using a MATLAB

code that processed the normal and tangential impact forces acquired during 10 s of the DEM

simulation. The local pressure, p, and the coefficient of internal friction, μ, were then calculated

using Eqs. (4.2) and (4.5), respectively. The independence of these values from the

measurement bin size was assessed by changing the side dimension to 4d and 5d at the

measurement point L1 (Fig. 4.4). Since the coefficient of internal friction calculated using

different bin sizes remained unchanged, the 3d bin was used in all measurements. Varying the

number of particle layers between the glass partitions from 4 to 8 did not change the stress

tensors, as expected from the results of ref. ‎[19].

In order to calculate the shear rate, three more measurement bins of the same size were defined

adjacent to the original bin and were located on the right side, top side and back side of the

points H1, L1 and L2 and on the left side, bottom side and back side of the point H2 (Fig. 4.4).

The instantaneous particle velocities were acquired in the four bins in three orthogonal directions

(x ,y ,z) and then used as an input to another MATLAB code to calculate the velocity gradient

tensor L. The ij-component of the velocity gradient tensor was defined as the variation in the

mean bulk flow velocity in direction i from the main bin to its adjacent bins divided by the

distance between the bin centers (equal to the size of the bin) along direction j (Eq. (4.9a)). The

mean bulk flow velocity was calculated as the average of the instantaneous velocity (considering

the resultant of velocities with positive and negative signs) of the particles moving in the main

measurement bin during 10 s ‎[19]. The strain rate tensor, D, is the symmetric part of the velocity

gradient tensor, L, (Eq. (4.9b)), and the shear rate was dev

ijD (Eq. (4.3)). The inertial number

was then calculated using Eq. (4.6) taking the steel particle density and diameter as ρs=7800

kg/m3 and d=6.3 mm, respectively.

( ) ( )i j j i j

ij

j

V x x V xL

x

(4.9a)

2

ij ji

ij

L LD

(4.9b)

Figure 4.5a shows that the coefficient of internal friction did not vary significantly with the

inertial number at the four measurement points; i.e. =0.12±0.01 as I varied from 0.005 to 0.012.

The coefficient of internal friction also did not change significantly with either the local pressure

Page 109: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

91

over the range of 50-170 kPa, or with the shear rate over range 2.5-8 1/s (Figs. 4.5b and 4.5c).

Therefore, a single constant value of =0.12 was used in the continuum model.

The volume fraction was calculated in the 4d measurement bins, and was also found to be

independent of bin size. As with , Fig. 4.5d shows that φ was essentially independent of I,

being the same at the four measurement locations (φ=0.57-0.61). Therefore, the volume fraction

was assumed to be constant, equal to 0.6.

Fig. 4.4: Measurement bin locations, H1, H2, L1 and L2 used to calculate the shear rates and

stress tensors from the DEM. Three in-plane bins at each location and a fourth bin behind the

main bin.

H1

L1

L2

H2

z

x

Page 110: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

92

Fig. 4.5: DEM predictions of the coefficient of internal friction as a function of: (a) the inertial

number, (b) the local pressure and (c) the shear rate; (d) Local volume fraction vs. the inertial

number. The solid and dashed lines are the average and the 95% confidence bounds.

4.3.2 Media equivalent Young's modulus and media-wall effective contact stiffness

It was necessary to determine the equivalent Young's modulus of the bulk media, Eb, to use in the

continuum model as shown in Eqs. (4.1) and (4.2b). The equivalent Young's modulus, Eb, of

quasi-static granular flows has been conventionally determined in tri-axial compression tests

either experimentally or numerically using DEM [25, 30-32]. It is also possible to use the

(a) (b)

(c) (d)

Page 111: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

93

Mindlin theory to derive equations for sound wave propagation speed in the granular media

assuming that the particles are always in contact and that force chains remain largely unchanged

‎[30]. Therefore, for the present vibrationally-fluidized beds, these approaches give inaccurate

values of the equivalent Young's modulus since the force chains change at every time step ‎[30].

An alternate approach was therefore employed, using the sound wave propagation speed, Ud, as

calculated from Eq. (4.10) using the density, stiffness and volume fractions of the solid and fluid

phases in the vibrationally-fluidized bed [25, 40-42].

;(1 ) (1 )

g g

d s s g s

b s

U V V

(4.10)

bd

b

KpU

(4.11)

where φ is the solid fraction, ρb is the bulk density calculated from ρb= φ ρs+(1- φ) ρg, and ρs and

ρg are the solid and gas phase densities, respectively. Vs is the speed of sound in air at standard

pressure and temperature; 330 m/s. In deriving this equation, the mixture was assumed to be

compressible, pseudo-homogeneous, and without relative motion between the phases ‎[41]. The

latter assumption was mostly valid in the present vibrationally-fluidized bed since the interstitial

air moved with the particles. As the air density is much smaller than that of the steel particles, ρb

was approximated by φρs in Eq. (4.10). The speed of sound in any media can also be obtained

using the Newton-Laplace equation (Eq. (4.11)) and depends on its bulk modulus of elasticity,

Kb, and the bulk density‎[25] [41, 42]. Knowing φ and hence ρb from the DEM simulations, Kb in

the vibrationally-fluidized state was determined by equating Eqs. (4.10) and (4.11). Finally,

assuming the Poisson's ratio to be ν=0.25, which is a typical value for granular flows ‎[32], the

equivalent Young's modulus was obtained from:

3(1 2 )

bb

KE

v

(4.12)

Figure 4.6 shows these predictions of the equivalent Young's modulus, Eb, for the present

vibrationally-fluidized media as a function of the solid fraction, φ, which varied with distance

from the wall. For the present tub finisher, the maximum Eb=498 kPa corresponded to the

maximum solid fraction φ=0.60 inside the media far from the wall, while the minimum value

Page 112: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

94

Eb=257 kPa corresponded to φ=0.23 obtained near the wall through the procedure described

below.

The effective normal contact stiffness between the vibrationally-fluidized media and the

tub wall, which was needed to apply the appropriate boundary condition to the media-wall

interface in the continuum model, was defined as the slope of the pressure-overlap curve; i.e. the

amount of pressure needed to cause an overlap of unit distance of the contacting surfaces.

Although the relationship might be slightly nonlinear ‎[24], it was approximated as linear so that

the contact stiffness, K, between the media and the wall was given by Eq. (4.13) as

1 1 1

media wallK K K (4.13)

where Kmedia=Eb/Lb and Kwall=Ew/Lw where Ew is the Young's modulus of the polyurethane wall,

and Lb and Lw are, respectively, the effective thickness of the media in the vicinity of the wall,

and the wall thickness. Since the modulus of the polyurethane wall was so much larger than that

of the bulk fluidized media (i.e. Ew=26 MPa, Eb=498 kPa) the effective contact stiffness could be

approximated as K=Eb/Lb. The parameter Lb represents a form of boundary layer thickness, over

which media-wall contact produces significant changes in the solid fraction, and hence the bulk

modulus and contact stiffness. In order to estimate the media effective thickness, Lb, the

variation of the media solid fraction normal to the wall near location H2 (Fig. 4.4) was obtained

using 7 measurement bins of size d×12d×16d (7×76×100 mm) in the 12-layer DEM simulation

(Fig. 4.7). The solid fraction variation was evaluated under both maximum compression (Fig.

4.7a) and maximum decompression (Fig. 7b) corresponding to the wall at its maximum

vibrational displacement to the left and right, respectively, in Fig. 4.7. Figure 4.8 shows how the

solid fraction varied with distance from the wall of Fig. 4.7 using 10 s of DEM simulation to

obtain average solid fraction values over many instances of maximum compression and

decompression. As expected, the largest variation in the solid fraction occurred near the wall

within the first few particle diameters. There was also only a relatively small difference between

the curves as maximum compression and decompression. The media effective thickness was

defined initially as the distance from the wall where the solid fraction changed less than 10%

from bin to bin, which was Lb=24 mm in Fig. 4.8, yielding Kb=15 MPa/m. The sensitivity of the

Page 113: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

95

continuum model to this choice of the effective media thickness at the wall will be examined

below.

0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65250

300

350

400

450

500

Solid fraction

Eq

uiv

ale

nt

Yo

un

g's

mo

du

lus

(k

Pa

)

Fig. 4.6: The equivalent Young's modulus of the vibrationally-fludized bulk flow, Eb, vs. the

solid fraction, φ, based on Eqs. (9-11) and the DEM.

Page 114: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

96

Fig. 4.7: Seven measurement bins used to study the media solid fraction variation with distance

from the wall when: (a) wall moves toward media leading to maximum compression, (b) wall

moves away from media leading to maximum decompression of the media and some separation

between the media and the wall.

Fig. 4.8: Average solid fraction, φ, variation with distance from the wall using data from the

measurement bins of Fig. 4.7.

(a) (b)

HW HW

Page 115: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

97

4.3.3 Media-wall equivalent coefficient of friction

The effective coefficient of friction between the wall and the vibrationally-fluidized media, mw,

which was needed to apply the appropriate boundary condition to the media-wall interface in the

continuum model, was determined using the three-dimensional DEM model with 8 particle

layers between the glass partitions as used in ref. ‎[19]. The shear forces (parallel to wall) and

normal forces acting on the media were determined during 1 s of simulated operation (equivalent

to 47 tub vibration cycles) at each of 8 locations along the tub wall as shown in Fig. 4.9, using

4d×6d×4d measurement bins. The appropriate simulation time step was calculated to be 0.686 µs

for the 6.3 mm diameter steel balls ‎[19].

Fig. 4.9: Eight measurement locations used to determine average shear and normal forces acting

on the media along the tub wall.

z

x

1

2

3

4

5

6

7

8

HW

LW

BW

Page 116: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

98

Figure 4.10 shows the probability density distributions of the shear to normal force ratio of all

particle-wall contacts, Ft/Fn, at each measurement location for 47 tub vibration cycles. The

maximum value of the predicted Ft/Fn in the distribution was 1.8, which corresponded to the

coefficient of friction that was measured for the 6.3 mm diameter steel balls sliding against the

urethane tub wall in ref. [19] and was used as an input to the DEM. It was observed that most of

particle collisions with the tub wall involved rolling rather than sliding since the most probable

values of Ft/Fn were smaller than 0.7 (Fig. 4.10a). The sliding impact of particles without rolling

is evident as a small peak at the right end of the probability distribution (Fig. 4.10a). Figure

4.10b shows that the probability density distributions of Ft/Fn at the different measurement

points were quite similar. This is a useful observation, because it means that a single relation can

represent the effective friction behaviour on the tub wall.

Page 117: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

99

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.80

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Shear to normal force ratio Ft/Fn

Pro

bab

ilit

y d

en

sit

y d

istr

ibu

tio

n

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.80

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Shear to normal force ratio Ft/Fn

Pro

bab

ilit

y d

en

sit

y d

istr

ibu

tio

n

Fig. 4.10: Probability density distributions of the ratio of the shear to normal forces applied to

the media by the wall (Ft/Fn) at the 8 locations of Fig. 4.9 for: (a) all data (columns ordered from

left to right as positions 1-8), and (b) for Ft/Fn <0.8. Based on all particle-wall impact events

recorded by the DEM during 47 vibration cycles

Figure 4.11 shows how the shear force varied with the normal force considering all impact

events over 47 vibration cycles at location 2 in Fig. 4.9. The slope of the best-fit line was 0.25

(a)

(b)

Page 118: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

100

with a coefficient of determination R2=0.7. This represents the average effective coefficient of

friction between the flowing steel balls and the tub wall at location 2 (µmw). As expected this

was much smaller than the measured coefficient of friction in sliding (µmw=1.80), since rolling

was prominent during contact. This is consistent with observations of Yabuki et al. ‎[28] who

found that abrasive ceramic media did not slide on an aluminum surface in a vibratory finisher,

resulting in an average effective coefficient of friction of about 0.10, compared with a measured

value for sliding of 0.35-0.40 ‎[28]. The average effective coefficient of friction, µmw, at the eight

locations on the tub wall varied from 0.25 to 0.35, giving an overall mean and standard deviation

of 0.28 and 0.02, respectively, as shown in Fig. 4.12. Figure 4.12 also shows the results from a

DEM simulation that used only 4 particle layers between the glass partitions, illustrating that the

predicted shear and tangential forces were independent of the number of layers.

Fig. 4.11: Shear force vs. normal force at location 2 (Fig. 4.9) for all wall impacts over 47

vibration cycles. The best-fit lines represent the mean effective coefficient of friction (µmw=0.25)

with the correlation coefficient, and the measured COF for sliding of a single steel ball against

the polyurethane wall material (µmw=1.8). The line for µmw=0.7 is also shown since it will be

used in the continuum modelling.

Page 119: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

101

Fig. 4.12: The average effective coefficient of friction between the fluidized media and the tub

wall at the 8 locations shown in Fig. 4.9 based on all impacts over 47 vibration cycles for the 4-

layer and 8-layer DEM. The horizontal lines show the average and 95% confidence bounds.

It is noted that the average shear force recorded in these DEM simulations explain the origin of

the bulk flow of the granuar media. Figure 4.13 shows how the average normal and shear forces

applied to the media by the high wall (HW), low wall (LW) and bottom wall (BW) (Fig. 4.9)

changed during a single cycle (average forces over each interval of 1/12 of a cycle from 5 s (235

vibration cycles) of DEM simulation using the 8-layer model). For example, the first point in

Fig. 4.14 represents the average of the impact forces acquired in the first time interval of 1/12 of

a cycle. In Fig. 4.13, shear force acting on the media is positive if it was applied as a counter-

clockwise torque (Fig. 4.13), thereby creating the observed bulk flow in Fig. 4.3.

Page 120: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

102

Fig. 4.13: Grand average shear force acting on the media by the wall over each 1/12 of a cycle

over 235 vibration cycles. Shear forces recorded over the high wall (HW; locations 1-3), low

wall (LW; location 4-6), and bottom wall (BW; locations 7-8) as shown in Fig. 4.9. Positive

shear force produced a counter-clockwise torque on the media.

4.4 Finite element implementation of the continuum model

Both Lagrangian and Eulerian finite element formulations of the continuum model were

investigated using the ABAQUS 6.11 software package (Dassault Systèmes Corp. 2014) ‎[34].

4.4.1 Lagrangian FE continuum model

The explicit plane strain Lagrangian element was used to mesh the 2D domain of the tub

vibratory finisher using two mesh densities and either free triangular elements or free

quadrilateral elements as shown in Fig. 4.14. These element sizes of 2 cm and 1 cm were,

respectively, 5.0% and 2.5% of the domain characteristic length of 40 cm. The mesh was

Page 121: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

103

adapted using 3 to 5 sweeps per increment to correct for excessive mesh distortion ‎[34]. The

mesh aspect ratio was restricted to be larger than 0.1.

The interaction between the media and the tub wall was defined as a surface-to-surface contact

‎[34]. The average effective coefficient of friction and normal contact stiffness were determined

from the DEM as in Sections 4.3.3 and 4.3.2, respectively. The displacement and rotation

components of the tub wall harmonic motion were the same as those used in the DEM (taken

from accelerometer measurements made in ref. ‎[9]). In most previous studies of continuum

modeling of non-fluidized granular flows, the simple no-slip boundary condition was assumed at

the wall surface ‎[25][24, 25, 27, 30]. However, the nature of the wall-media interaction in the

present vibrationally-fluidized bed could not be assumed as being either slip or no-slip, since it

was a combination of rolling and sliding of the media on the wall as described in Section 4.3.3.

Therefore, an effective media-wall coefficient of friction that accounted for media rolling and

sliding, and a normal contact stiffness were considered to apply the appropriate boundary

condition to the wall-media interface. These parameters as well as the media equivalent density,

Young's modulus, and coefficient of internal friction were determined as in Section 4.3, and are

listed in Table 4.3. The only body force exerted on the media was gravity.

Page 122: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

104

Fig. 4.14: The granular media domain meshed using Lagrangian elements: Free triangular

elements of sizes (a) 2 cm and (b) 1 cm; free quadrilateral elements of sizes (c) 2 cm and (d) 1

cm; tub wall meshed using free triangular elements of size 2 cm.

(a) (b)

(c) (d)

Page 123: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

105

Table 4.3: Material and contact properties used in the FE simulations. The media effective

Young’s‎modulus,‎ and‎ the‎ contact‎ parameters‎were‎ evaluated‎ over‎ the‎ indicated‎ ranges.‎ ‎ The‎

values of the media Young's modulus and the media-wall contact stiffness in parentheses were

calculated at the mean volume fraction φmean; i.e. the best estimate.

Material Density, ρb

(kg/m3)

Young's modulus,

Eb (kPa)

Coefficient of

internal friction, µ

Media 4680 257-498

(340 at φmean) 0.12

Tub wall 1200 26000 N/A

Contact Media-wall contact stiffness, K

(MPa/m)

Coefficient of media-wall

friction, µmw

Media-tub wall interaction 15.0-30.2 (19.7 at φmean) 0.1-0.7 (mean=0.3)

As an example, Fig. 4.15 shows the bulk flow velocity distributions obtained from the

Lagrangian analysis using the different elements with the media equivalent Young's modulus

Eb=257 kPa, the effective normal contact stiffness K=30.2 MPa/m, and the media-wall

coefficient of friction µmw=0.3. It is seen that the element type and size did not affect the bulk

flow patterns significantly, and that in all cases the flow fields were qualitatively similar to that

of the equivalent DEM, shown in Fig. 4.3. The mean bulk flow velocity was determined by

averaging the nodal vertical velocities perpendicular to a horizontal line from the center of

circulation to the right wall as shown in Fig. 4.15. The average inclination of the free surface

was the slope of a straight line fitted to the free-surface of the media ‎[19]. Although the flow

fields had the correct form, the mean bulk flow velocity obtained in this Lagrangian FE

continuum model was 4.5 mm/s compared to 14 mm/s from the DE simulation. Section 4.5

describes the introduction of an adjustable parameter to the continuum model in order to improve

the accuracy of the bulk flow velocity.

The present FE model was two-dimensional and did not allow for the out-of-plane particle

motion that can occur in the tub finisher and in the three-dimensional DEM. The DEM

simulations of ref. [19] showed that the average out-of-plane particle impact velocities can be

Page 124: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

106

approximately 20% of the in-plane components. This might be one of the reasons for the

difference between the bulk flow velocities predicted by the FE and DE simulations.

Fig. 4.15: The bulk flow velocity distribution obtained from the Lagrangian analysis: Free

triangular elements of sizes (a) 2 cm and (b) 1 cm; free quad elements of sizes (c) 2 cm and (d) 1

cm. The simulation parameters: K=30.2 MPa/m, Eb=257 kPa and µmw=0.3. Nodal velocity scale

(a) (b)

(c) (d) 2 mm/s

MFV: 4.5 mm/s

Surface angle: 15

MFV: 4.5 mm/s

Surface angle: 15

MFV: 5 mm/s

Surface angle: 17

MFV: 5 mm/s

Surface angle: 17

Page 125: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

107

shown in lower right. Some arrows omitted for clarity. Mean bulk flow velocity (MFV) across

the dark horizontal lines.

4.4.2 Eulerian FE continuum model

In an Eulerian analysis, the mesh is spatially fixed and the material can move through it as

demonstrated in Fig. 4.16 ‎[45]. An Eulerian analysis is normally used when the material

deformation is relatively large and Lagrangian elements become excessively distorted. Although

remeshing can resolve this distortion, it is computationally expensive, and it was of interest to

investigate the accuracy of an Eulerian analysis, which is faster to execute.

In the present case, an Eulerian mesh was generated to contain the tub wall and ensure that the

contact between the media and the tub wall was enforced accurately and the media did not leave

the Eulerian domain (Fig. 4.17) ‎[34]. A Lagrangian mesh was used for the tub wall. A single

layer of 2 cm thick 3D hexagonal elements, the only Eulerian element type available in

ABAQUS, was used for the Eulerian analysis. The elements were assumed to be rectangular in

order to decrease the computational time and to increase the accuracy of the results ‎[34]. Two

element sizes were evaluated, 1 cm and 0.5 cm, which were respectively 2.5% and 1.25% of the

domain characteristic length (Fig. 4.17). Since the wall deformations were very much smaller

than those of the media, a rigid body constraint was applied to the Lagrangian wall domain to

greatly decrease the run time.

Page 126: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

108

Fig. 4.16: Material deformations in the Lagrangian and Eulerian analyses ‎[45].

The volume fraction tool in ABAQUS ‎[34] was used to fill the Eulerian elements occupied by

the media at time zero. By refining the mesh, the calculation of the volume fraction used to

define the material initial location became more accurate, minimizing the artefact of media-wall

overlap ‎[34]. The contact between the media and the tub wall was modelled using the same

effective coefficient of friction and normal contact stiffness considered in the Lagrangian

analysis (Table 4.3). The velocity component normal to the plane of the bulk flow was again

considered to be zero as the flow was two-dimensional.

Lagrangian analysis Eulerian analysis

Page 127: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

109

Fig. 4.17: The media meshed by Eulerian elements of sizes (a) 1 cm and (b) 0.5 cm. The tub

wall was meshed with Lagrangian elements.

Figure 4.18 shows the Eulerian mesh occupied by the media and the bulk flow velocity

distributions using the two sizes of Eulerian rectangular elements and the same parameters as

used in the Lagrangian analysis (K=30.2 MPa/m, Eb=257 kPa and µmw=0.3). The global

circulation observed in Fig. 4.18 was not dependent on the mesh size and was similar to that for

the Lagrangian model (Fig. 4.15). However, the mean bulk flow velocity obtained from the

Eulerian analysis was 90 mm/s, much larger than that computed in the Lagrangian analysis (4.5

mm/s), and in the DEM simulation (14 mm/s). The reason for this much higher bulk flow

velocity is unclear. Consequently, only the Lagrangian FE formulation was used to perform the

investigations of Sections 4.5, 4.6 and 4.7.

(a) (b)

Page 128: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

110

Fig. 4.18: Bulk flow velocity distributions obtained from the Eulerian analysis for rectangular

elements of sizes: (a) 1 cm and (b) 0.5 cm. Nodal velocity scale shown in upper left. Some

arrows omitted for clarity.

4.5 Sensitivity to uncertainty in continuum model parameters

As discussed in Sections 4.3.2 and 4.3.3, the DEM of the vibrationally-fluidized flow showed

that there was considerable uncertainty in the equivalent continuum model parameters in the

vicinity of the media-wall interface. In particular, the media-wall coefficient of friction varied

about µmw=0.3 (Fig. 4.11), while uncertainty in the effective thickness of the media-wall

interaction zone, Lb, resulted in uncertainty in the dependent parameters Eb, the media equivalent

Young's modulus, and K, the media-wall effective contact stiffness (Eq. (4.12) and Fig. 4.6). It

was therefore of interest to assess the sensitivity of the continuum model to these uncertainties.

This was done using the Lagrangian FE analysis since the Eulerian FE simulations predicted

velocities significantly larger than the typical velocities observed in the DE simulations. The

mean bulk flow velocity was chosen as the measure of the continuum model response to these

media-wall parameters. Three values of the effective contact stiffness (K=15.0 MPa/m

(corresponding to Lb=24 mm), 19.7 MPa/m (Lb=16 mm), 30.2 MPa/m (Lb=10 mm)) and two

(a) (b)

100 mm/s

Page 129: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

111

values of the media Young's modulus (Eb=257 (corresponding to Lb=10 mm) kPa, 498 kPa

(Lb=24 mm)) were selected while the coefficient of friction was fixed at µmw=0.3 to study the

sensitivity of the bulk flow behavior to K and Eb (first six combinations in Table 4). Then, to

study the effect of the coefficient of media-wall friction on the bulk behavior, four values of µmw

(0.1, 0.3, 0.5, 0.7) were used with the media equivalent Young's modulus and the media-wall

effective contact stiffness fixed as K=30.2 MPa/m, Eb=257 kPa (last four combinations in Table

4). The media equivalent density, coefficient of internal friction, and Poisson's ratio were held

constant at the values of Table 3. The first six combinations in Table 4 show that the mean bulk

flow velocity increased marginally (from 3 to 5 mm/s) as the effective contact stiffness, K,

increased. Presumably this resulted from the larger coefficient of restitution that followed from

an increase in the contact stiffness. Similarly, the mean bulk flow velocity was insensitive to

changes in the media equivalent Young's modulus, Eb. Much larger changes in the mean bulk

flow velocities (2.5 to 14 mm/s) were seen as the media-wall coefficient of friction, µm-w, was

varied from 0.1 to 0.7 (last four combinations in Table 4). It is noted that these conclusions

pertain to the mean bulk flow velocity, and that different trends may have been observed with

other characteristics of the flow.

Page 130: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

112

Table 4.4: Sensitivity of the Lagrangian FE model mean bulk flow velocity (MFV) and free-

surface angle to changes in the effective media-wall contact stiffness, media equivalent Young's

modulus, and media-wall coefficient of friction.

K (MPa/m), Eb (kPa), µmw

MFV

(mm/s)

Free-surface

angle ()

Eb=257 kPa, µm-w=0.3

K=15.0 MPa/m 3.0 8.0

K=19.7 MPa/m 3.5 11

K=30.2 MPa/m 4.0 15

Eb=498 kPa, µmw =0.3

K=15.0 MPa/m 3.0 10

K=19.7 MPa/m 4.0 12

K=30.2 MPa/m 5.0 15

K=30.2 MPa/m, Eb=257 kPa

µmw=0.1 2.5 11

µmw=0.3 4.5 15

µmw=0.5 7.5 20

µmw=0.7 13 20

4.6 Comparison of FEA and DEM bulk flow velocities

As discussed in Section 4.4, although the flow fields predicted by the FE continuum models were

similar to those of the DEM, there were considerable differences in the predicted mean bulk flow

velocities, particularly with the Eulerian formulation. This is not too surprising given the

uncertainties in the media-wall contact parameters discussed in Section 4.5. It was therefore of

interest to adjust one of these contact parameters to obtain the best fit with the DEM, and then to

assess whether the continuum model with this optimized parameter continued to provide

satisfactory agreement for other vibrationally-fluidized flows.

Page 131: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

113

The results of the Section 4.5 showed that the Lagrangian FE continuum model was most

sensitive to the media-wall coefficient of friction; therefore, it was selected as the adjustable

parameter. Trials with a range of µmw within the plausible range shown in Fig. 4.11 indicated

that the best agreement with the mean bulk flow velocity, MFV, occurred with µmw=0.7. Figure

4.19a compares the predictions of the Lagrangian FE model and the DEM for the actual

measured tub vibration amplitude, A, and for amplitudes of 0.5A, 0.75A and 1.5A. It is seen that

there is good agreement with the MFV for the actual tub amplitude, and that this agreement

persists when the vibrationally-fluidized flow is changed significantly by altering the tub wall

amplitude; i.e. the changes in MFV were approximately in proportion to the changes in the wall

amplitude. Similarly, the free surface angle increased with increasing wall amplitude as shown

in Fig. 4.19b. The angle predicted by the FE model with the optimized µmw=0.7 was 4 larger

than that of the DEM for the actual tub amplitude, A, and this difference increased to 7 when the

amplitude was 1.5A.

Figure 4.20 compares the distribution of the component of the bulk flow velocity perpendicular

to a horizontal line from the center of rotation to the wall as shown in Fig. 4.15. The local bulk

flow velocities obtained from the DEM and the FE simulation with µmw=0.7 were in reasonable

overall agreement, with the under-prediction of the FE model growing near the tub wall. This

may have been due to the rapid change in the bulk flow density and other contact parameters

near the wall (Figs. 4.6 and 4.8).

Overall, it is concluded that using a media-wall coefficient of friction µmw=0.7, optimized using

the DEM for the actual tub wall amplitude, provided reasonably accurate continuum predictions

of bulk flow velocity and free-surface inclination over a wide range of other vibrationally-

fluidized flow fields created by varying the tub wall amplitude.

Page 132: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

114

Fig. 4.19: (a) mean bulk flow velocity and (b) inclination of free surface as a function of tub

amplitude expressed as the fraction of the actual tub amplitude, A. Predictions of the DEM and

of the Lagrangian FE continuum model that used the optimized media-wall effective coefficient

of friction, µmw=0.7. FE predictions with µmw=0.5 included for comparison.

(b)

µmw=0.7

µmw=0.5

µmw=0.7

µmw=0.5

(a)

Page 133: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

115

Fig. 4.20: Distribution of bulk flow velocity perpendicular to the horizontal line defined in Fig.

4.15 in the 4-layer DE and the Lagrangian FE using the optimized µmw=0.7 (K=30.2 MPa/m,

Eb=295 kPa). Zero distance corresponds to the center of circulation and 160 mm is at the tub

wall.

4.7 Comparison of FEA and DEM local impact velocities

The previous sections considered the DEM and the FE continuum model predictions of the bulk

flow velocity and the inclination of the free surface. It is also of great interest to consider

whether the continuum model can yield predictions of the local particle impact velocity, or at

least predictions of parameters related to this velocity, since the erosion and surface deformation

produced in vibratory finishing are proportional to the particle kinetic energy.

The granular temperature, T, of an assembly of N particles is a function of the fluctuations of the

velocity components (u'i, v

'i, w

'i) as shown in Eq. (4.14a) ‎[15], where the fluctuations are the

difference between the instantaneous (ui, vi, wi) and the mean velocities ū, ῡ, w (Eq. (4.14b)).

The granular temperature is the also a function of the mean free path, l, the shear rate and the

coefficient of restitution e (Eq. (4.15a)) ‎[46]. The mean free path is related to the volume

Page 134: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

116

fraction through Eq. (4.15b). Considering the particles to be spherical, the mean free path would

be proportional to the particle diameter, d, divided by the volume fraction, φ.

2 2 2

1

1( )

3

N

i i i

i

T u v wN

(4.14a)

, ,w ,i i i i i iu u u v v v w w (4.14b)

2 2

2(1 )

c lT

e

(4.15a)

2

3

dl

(4.15b)

Combining Eqs. 4.14(a) and 4.14(b), it is seen that the square of the granular temperature, T2, is

proportional to the difference between the average of the instantaneous velocities squared,

2

1

/N

i

i

V N

, and the bulk flow velocity squared, 2V (Eq. (4.16a, b, c)). Since the granular

temperature is proportional to the shear rate squared through Eq. (4.15a), it is seen that a measure

of the local impact velocities, 2

1

/N

i

i

V N

, is related to the shear rate. Therefore, if the shear rate

follows the same trend in the DE and FE simulations, it can be inferred that the impact velocities

would follow the same trend.

2

2 2 2 21

1

1 1( ) ( ) ( )

3 3 3

N

iNi

i i i

i

V

T u u v v w w VN N

(4.16a)

2 2 2 2V u v w (4.16b)

2

21 3

N

i

i

V

T VN

(4.16c)

Figure 4.21a confirms this hypothesis by showing that the mean impact velocity and its

components increased with increasing shear rate calculated at the four locations shown in Fig.

4.4 in the 4-layer DE model. The mean impact velocities were obtained by averaging the

instantaneous particle velocities (the absolute values of the impact velocities) over 10 s, as

Page 135: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

117

described in ref. ‎[19]. Thus the shear rate is a possible proxy for the magnitude of the local

average impact velocity of the media.

The average shear rate was then determined in the Lagrangian FE model (K=30.2 MPa/m,

Eb=295 kPa, µmw=0.7) by following the same procedure described in Section 4.3.1 in the

calculation of the shear rate in the DEM. Figure 4.21b shows that the shear rate obtained from

the DE and FE simulations followed the same trend at the four assessment points in the flow

field. Therefore, it is concluded that the local shear rate predicted by the FE continuum model

correctly reflected the trend of the local impact velocity distribution in the present vibrationally-

fluidized granular flow. This implies that the FE continuum model may prove to be useful as a

means of predicting the impact velocity distribution within a vibratory finisher, and hence the

distribution of the impact energy imparted to workpiece surfaces as they move with the bulk

flow.

Fig. 4.21: (a) Average particle impact velocity, Vimp, and its components vs. the shear rate, , at

points H1, L1, L2, H2 (ordered from left to right) in the 4-layer DEM (Fig. 4.4). (b) Shear rate in

FE and DE simulations at the same points.

(a) H1 L1 L2 H2 (b)

Page 136: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

118

4.8 Conclusions

A continuum finite element model was used successfully to simulate the flow behavior in a tub

vibratory finisher. The effective bulk and media-wall properties for the constitutive equation

governing the visco-plastic continuum model of the spherical, vibrationally-fluidized media were

found using a discrete element model. It was seen the continuum model could be simplified by

using constant, average values of the constitutive properties for the vibrationally-fluidized media

(density, modulus, internal friction), and for the media-wall interaction parameters (friction,

contact stiffness).

The bulk behavior of the vibrationally-fluidized granular flow was modelled using both a

Lagrangian and an Eulerian finite element implementation of the continuum model. The

predicted patterns of the bulk flow velocity distributions agreed with the DEM simulations and

with experimental observations. Good quantitative agreement was obtained by treating the

media-wall coefficient of friction as an adjustable parameter. The generality of this approach

was demonstrated by using this optimized value to predict the bulk flow velocities created by

different tub vibration amplitudes. It was also observed that the local impact velocities of the

vibrationally-fluidized media could be correlated with the local shear rate in the continuum

model through its relation to the granular temperature. It was concluded that the bulk and local

behavior of vibrationally-fluidized flows can be modelled using finite element implementations

of a visco-plastic continuum model. This can lead to significant reductions in computational

effort compared with discrete element modelling.

4.9 References

[1] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, A. Kong, P. J. Narayan, B. C. Hancock,

Granular segregation in discharging cylindrical hoppers: A discrete element and experimental

study, Chem. Eng. Sci. 62 (2007) 6423 – 6439.

[2] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, B. C. Hancock, Predicting the flow mode

from hoppers using the discrete element method, Powder Technol. 195 (2009) 1–10.

Page 137: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

119

[3] R.Y. Yang, R.P. Zou, A.B. Yu, Microdynamic analysis of particle flow in a horizontal

rotating drum, Powder Technol. 130 (2003) 138–146.

[4] V. Jasti, C. F. Higgs, Experimental study of granular flows in a rough annular shear cell,

Phys. Rev. E 78 (2008) 041306-1-8.

[5] N.G. Deen, M. Van Sint Annaland, M.A. Van der Hoef, J. A. M. Kuipers, Review of

discrete particle modeling of fluidized beds, Chem. Eng. Sci. 62 (2007) 28-44.

[6] X. Z. An, C.X. Li, R.Y. Yang, R.P. Zou, A.B. Yu, Experimental study of the packing of

mono-sized spheres subjected to one-dimensional vibration, Powder Technol. 196 (2009)

50–55.

[7] S.E. Naeini, J.K. Spelt, Two-dimensional discrete element modeling of spherical steel media

in a vibrating bed, Powder Technol. 195 (2009) 83-90.

[8] S.E. Naeini, J.K. Spelt, Development of single-cell bulk circulation in granular media in a

vibrating bed, Powder Technol. 211 (2011) 176-186.

[9] K. Hashemnia, A. Mohajerani, J. K. Spelt, Development of a laser displacement probe to

measure particle impact velocities in vibrationally fluidized granular flows, Powder

Technol., 235 (2013) 940-952.

[10] D. Ciampini, M. Papini, J. K. Spelt, Impact velocity measurement of media in a vibratory

finisher, J. Mater. Processing Tech. 183 (2007) 347-357.

[11] S. Wang, R. S. Timsit, and J. K. Spelt, Experimental investigation of vibratory finishing of

aluminum, Wear 243 (2000) 147-156.

[12] A. Yabuki, M.R. Baghbanan, J.K. Spelt, Contact forces and mechanisms in a vibratory

finisher, Wear 252 (2002) 635–643.

[13] EDEM website, http://www.dem-solutions.com/academic, 2012.

[14] A. Munjiza, The combined finite discrete element method, John Wiley & Sons Ltd.,

Chichester, West Sussex, England, 2004.

[15] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of particulate

systems: A review of major applications and findings, Chem. Eng. Sci. 63 (2008) 5728-

5770.

[16] C.H. Tai, S.S. Hsiau, C.A. Kruelle, Density segregation in a vertically vibrated granular bed,

Powder Technol. 204 (2010) 255-262.

[17] M. Majid, P. Walzel, Convection and segregation in vertically vibrated granular beds,

Powder Technol. 192 (2009) 311-317.

Page 138: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

120

[18] R. L. Stewart, J. Bridgwater, Y. C. Zhou, A. B. Yu, Simulated and measured flow of

granules in a bladed mixer - A detailed comparison, Chem. Eng. Sci. 56 (2001) 5457-5471.

[19] K. Hashemnia, J. K. Spelt, Particle impact velocities in a vibrationally fluidized granular

flow: Measurements and discrete element predictions, Chem. Eng. Sci., 109 (2014) 123–

135.

[20] L. Xiang, W. Shuyan, L. Huilin, L. Goudong, C. Juhui, L. Yikun, Numerical simulation of

particle motion in vibrated fluidized beds, Powder Technol. 197 (2010) 25–35.

[21] B. Freireich, J. Litster, C. Wassgren, Using the discrete element method to predict collision-

scale behavior: A sensitivity analysis, Chem. Eng. Sci. 64 (2009) 3407- 3416.

[22] J.J. McCarthy, V. Jasti, M. Marinack, C.F. Higgs, Quantitative validation of the discrete

element method using an annular shear cell, Powder Technol. 203 (2010) 70-77.

[23] W. R. Ketterhagen, J. S. Curtis, C. R. Wassgren, Stress results from two-dimensional

granular shear flow simulations using various collision models, Phys. Rev. E 71 (2005)

061307-1-11.

[24] K. Kamrin, Nonlinear elasto-plastic model for dense granular flow, Int. J. Plasticity, 26

(2010) 167–188.

[25] C. H. Rycroft, K. Kamrin, M. Z. Bazant, Assessing continuum postulates in simulations of

granular flow, J. Mech. Phys. Solids, 57 (2009) 828–839.

[26] P. Jop, Y. Forterre, O. Pouliquen, A constitutive law for dense granular flows, Nature, 441

(2006) 727-730.

[27] Y. Forterre, O. Pouliquen, Flows of dense granular media, Annu. Rev. Fluid Mech. (2008)

1-24.

[28] G. D. R Midi, On dense granular flows, Eur. Phys. J. E (2004) 341-365.

[29] F. Cruz, S. Emam, M. Prochnow, J. N. Roux, and F. Chevoir, Rheophysics of dense

granular materials: Discrete simulation of plane shear flows, Phys. Rev. E 72 (2005)

021309-1-17.

[30] J. E. Andrade, Q. Chen, P. H. Le, C. F. Avila, T. M. Evans, On the rheology of dilative

granular media: Bridging solid- and fluid-like behavior, J. Mech. Phys. Solids 60 (2012)

1122–1136.

[31] J. E. Andrade, C. F. Avila, S. A. Hall, N. Lenoir, G. Viggiani, Multiscale modeling and

characterization of granular matter: From grain kinematics to continuum mechanics, J.

Mech. Phys. Solids 59 (2011) 237–250.

Page 139: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

121

[32] J. E. Andrade, X.Tu, Multiscale framework for behavior prediction in granular media,

Mech. Mater. 41 (2009) 652–669.

[33] Y. Guo, C. Wassgren, W. Ketterhagen, B. Hancock, B. James, J. Curtis, A numerical study

of granular shear flows of rod-like particles using the discrete element method, J. Fluid

Mech. 713 (2012) 1-26.

[34] ABAQUS 6.11 software package and analysis user's manual, http://www.3ds.com/products-

services/simulia, Dassault Systèmes Corp., 2014.

[35] W. M. Lai, D. Rubin, E. Krempl, Introduction to Continuum Mechanics, Fourth Edition,

Elsevier Science, Burlington, MA, US, 2009.

[36] C. S. Campbell, A. Gong, The stress tensor in a two-dimensional granular shear flow, J.

Fluid Mech. 164 (1986), 107-125.

[37] C. S. Campbell, Boundary interactions for two-dimensional granular flows. Part 1. Flat

boundaries, asymmetric stresses and couple stresses, J. Fluid Mech. 247 (1993), 111-136.

[38] J.P. Bardet, I. Vardoulakis, The asymmetry of stress in granular media, Int. J. Solids Struct.

38 (2001) 353-367.

[39] J. Fortin, O. Millet, G. de Saxcé, Construction of an averaged stress tensor for a granular

medium, Eur. J. Mech. A-Solid 22 (2003) 567–582.

[40] W. Gregor, H. Rumpf, Velocity of sound in two-phase media, Int. J. Multiphase Flow, 1

(1975) 753-769.

[41] H. T. Bi, J. R. Grace, J. Zhu, Propagation of pressure waves and forced oscillations in gas-

solid fluidized beds and their influence on diagnostics of local hydrodynamics, Powder

Technol. 82 (1995) 239-253.

[42] S. Sasic, B. Leckner, F. Johnsson, Characterization of fluid dynamics of fluidized beds by

analysis of pressure fluctuations, Prog. Energ. Combust. Sci. 33 (2007) 453–496.

[43] A.B. Stevens, C.M. Hrenya, Comparison of soft-sphere models to measurements of collision

properties during normal impacts, Powder Technol. 154 (2005) 99 –109.

[44] O. Al Hattamleh, B. Muhunthan, H. M. Zbib, Stress distribution in granular heaps using

multi-slip formulation, Int. J. Numer. Anal. Meth. Geomech., 29 (2005) 713–727.

[45] G. Qiu, S. Henke, J. Grabe, Application of a Coupled Eulerian–Lagrangian approach on

geomechanical problems involving large deformations, Comput. Geotech. 38 (2011) 30–39.

[46] I. Goldhirsch, Introduction to granular temperature, Powder Technol. 182 (2008) 130–136.

Page 140: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

122

Chapter 5

5 Conclusions

5.1 Experimental measurements

A laser displacement sensor was used to construct a probe that can be immersed in a granular

flow to measure local impact velocities, bulk flow velocities and a measure of particle packing.

The accuracy of impact velocity measurements was verified using drop tests with the same

granular media. The sensor output was modeled to interpret laser reflections from media passing

before the sensor in different trajectories and to distinguish the media vibrational impact velocity

from the oscillations of the sensor window. The main conclusions of the experimental

measurements were as follows:

The sensor was demonstrated in a tub vibratory finisher with two types of media-steel

and porcelain spheres. In this case, the average bulk flow velocity parallel to the sensor

window in the steel media was almost 40% larger than that in the porcelain media, and

both varied appreciably over the different measurement locations within a single media.

Similarly, the impact velocity and particle packing varied considerably among the various

locations and in different directions. Although it was concluded that there was no direct

proportionality between the impact velocity of particles, the bulk flow velocity, and the

packing density of the media, the smallest packing density and impact velocity occurred

in the wake formed behind the sensor shielding tube. These two parameters were larger

when the sensor was facing the average moving bulk flow.

It was also observed that the impact velocity of the particles was strongly dependent on

depth in the flow, increasing sharply with distance from the free surface.

The average impact velocities in the steel media were approximately 15% smaller than

those in the porcelain media due to differences in the density and coefficient of

restitution.

Page 141: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

123

In addition to providing a means of quantifying the impact energy distribution within a

vibratory bed, the experimental results can be used to validate the predictions of

numerical simulations such as discrete element modeling (DEM). Such models can help

to understand the complex patterns of behavior that have been observed within these

vibrationally fluidized beds.

The laser displacement probe and the current procedures are directly applicable to the

measurement of the local impact velocities of non-spherical particles. The measurement

of the bulk flow velocity in the case of irregular particle shapes would be more difficult,

since the variation in the average displacement signal would be more irregular.

5.2 Discrete element modeling

The impact and bulk flow velocities in a tub vibratory finisher predicted by discrete element

modeling (DEM) were compared to the measured values for steel granular media. This was the

first time measurements of local impact velocities have been compared with DEM predictions in

a vibrationally-fluidized granular flow. The sensitivity of the predicted local impact velocities

and the bulk flow velocities to the DEM contact parameters (coefficients of friction, restitution

and rolling resistance) was investigated both analytically and numerically. The vibration of the

tub finisher walls was essentially two-dimensional, with negligible motion in the transverse

direction. The effect of the number of DE particle layers in the transverse direction was

investigated to determine the minimum number required for accurate predictions of the particle

impact and bulk flow velocities. Moreover, a reduced shear modulus could be used to decrease

the model run time without significantly affecting the accuracy of the predicted impact and bulk

flow velocities. The complete DE model simulated the actual setup used in an earlier

experimental study that measured the bulk flow and local impact velocities in the tub vibratory

finisher. The coefficients of friction and restitution between the steel balls, and between the balls

and the tub walls, were measured in order to obtain the most accurate DE simulations. The main

conclusions drawn from the discrete element simulations were:

Page 142: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

124

It was found that the predictions of both the bulk flow and impact velocities were relatively

insensitive to uncertainties in the contact parameters.

It was concluded that accurate velocities could only be predicted using at least 4-12 layers.

Fewer layers created too much spatial constraint to allow realistic particle motion.

The predicted impact velocities in several directions and at two locations in the tub displayed

the same log-normal probability density distributions as did the previously measured impact

velocity distributions.

Moreover, the predicted average impact velocities were in good agreement with the

experimental measurements, with a maximum error of 19%. The DE predictions of the bulk

flow velocities were also consistent with the experimental measurements, having a maximum

error of 30%.

Therefore, it was concluded that DEM can be used to give reasonably accurate predictions of

both the local impact velocities and the bulk flow of particles in vibrationally-fluidized beds.

This will be useful in predicting the impact energy and force of the media on workpiece

surfaces, and hence the resulting wear and surface deformation during vibratory finishing.

5.3 Continuum modeling

A continuum finite element model was used successfully to simulate the flow behavior in a tub

vibratory finisher. The effective bulk and media-wall properties for the constitutive equation

governing the visco-plastic continuum model of the spherical, vibrationally-fluidized media were

found using a discrete element model. The bulk behavior of the vibrationally-fluidized granular

flow was modelled using both a Lagrangian and an Eulerian finite element implementation of the

continuum model. The main conclusions drawn from the finite element continuum modeling

were as follows:

Page 143: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

125

It was seen the continuum model could be simplified by using constant, average values of

the constitutive properties for the vibrationally-fluidized media (density, modulus,

internal friction), and for the media-wall interaction parameters (friction, contact

stiffness).

The predicted patterns of the bulk flow velocity distributions agreed with the DEM

simulations and with experimental observations. Good quantitative agreement was

obtained by treating the media-wall coefficient of friction as an adjustable parameter.

The generality of this approach was demonstrated by using this optimized value to

predict the bulk flow velocities created by different tub vibration amplitudes.

It was also observed that the local impact velocities of the vibrationally-fluidized media

could be correlated with the local shear rate in the continuum model through its relation

to the granular temperature.

It was concluded that the bulk and local behavior of vibrationally-fluidized flows can be

modelled using finite element implementations of a visco-plastic continuum model. This

can lead to significant reductions in computational effort compared with discrete element

modelling.

Page 144: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

126

6 Future work

The following topics may prove to be interesting areas for future research:

The experimental measurement of impact velocities using the laser displacement sensors

can be extended to the granular flows of non-spherical.

The continuum modeling of flows of non-spherical particles would be useful, although

the calibration of the equivalent bulk properties of non-spherical particles is expected to

be more challenging than spherical particles.

It would be of interest to apply the same procedure followed in the continuum modeling

of the granular flow produced in the vibratory finisher to fluidized beds in other

geometries, such as in a bowl vibratory finisher. It may also be possible to extend the

present single-phase continuum model to a two-phase model that could be applied to the

granular flows that are fluidized by a gas or liquid.

Using the concept of granular temperature to obtain an estimation of the average impact

velocities from the shear rates in the continuum modeling can be explored further.

Page 145: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

127

Appendix A: Analytical sensitivity study in Chapter 3

This Appendix describes the derivation of the equations describing the sensitivity of particle-

particle and particle-wall rebound velocities to uncertainties in the collision coefficients. This

was then be used to determine the required measurement accuracy for these coefficients. The

linear and angular rebound velocities of disks were determined as functions of the contact

parameters using the equations of linear and angular momentum conservation [1]. The

coefficient of rolling resistance is defined as the ratio of the rolling resistance force to the normal

force (or the ratio of the rolling resistance torque to the product of the normal force and the

particle radius), due to the elastic deformation of the bodies in the contact region [2, 3]‎[25].

Figure A.1 of case (1), is described by

1 1 2 2 1 1 2 2n n n nm v m v mV m V (A.1)

1 1 2 2 1 1 2 2t t t tm v m v mV m V (A.2)

V and v were designated for pre-collision and post-collision velocities, and n and t were

designated for normal and tangential directions, respectively. Conservation of angular

momentum gives

1 1 2 2 1 1 2 2J J J J (A.3)

where Ω and ω are the angular speeds before and after the collision, respectively. The normal

and tangential impulse are then defined as

1 1 1 2 2 2n n n n nP m v V m v V

1 1 1 1 1 1 1 2 2 2 2 2 2 2t t r t t tP m v V r m v r V r

(A.4)

The coefficient of rolling resistance ( rµ ) was calculated from Eqs. (A.6) and (A.7) [26].

1 1 1 2 2 2 1 1 1 1 1 1 2 2 2 2 2 2 2n n n n t t tm v V m v V m v r V r m v r V r (A.5)

Page 146: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

128

1 1 1 1t rr P M J (A.6)

2 2 2 2t rr P M J (A.7)

Where µ=Pt/Pn is the coefficient of friction, and * r r nM µ R P is the rolling torque with R* being

the equivalent radius of curvature (Eq. 1b). Equations (A.1-A.7) can be written as Eq. (A.8) with

2

1 2 1 2 1 2, , 1/ 2m m m r r r J J mr .

1

1

2

2

1

2

1 0 1 0 0 0 1 0 1 0 0 0

0 1 0 1 0 0 0 1 0 1 0 0

1 0 1 0 0 0 0 0 0 0

1 1 1

0 2 0 2 0 2 0

0 0

n

t

n

t

r r

v

v

v e e

r r v

r r

µ µ r r µ

1

1

2

2

1

2

1

2

0 0

n

t

n

t

r r

V

V

V

r r V

r r

µ r r

(A.8)

Fig. A.1: Collision of two disks having arbitrary initial linear and angular velocities: (a)

velocities after collision, and (b) impulses during impact.

Equation (A.8) gives the rebound velocities of both disks as functions of their initial velocities

and the collision coefficients. The normal rebound velocity, v1n, the tangential rebound velocity,

v1t, and the angular rebound velocity, ω1, are written in Eq. (A.9) as functions of the contact

parameters and normal impact velocities prior to collision. The sensitivities of v1n, (Eq. (A.9a)),

(a) (b)

Page 147: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

129

v1t, (Eq. (A.9b)), and ω1, (Eq. (A.9c)) to the collision coefficients were determined by

differentiating these velocities with respect to the coefficients as:

1 2 1 2 1

1 1

2 2n n n n nv V V e V V eC

12 1

1

2

nn n

vV V C

e

1 1 0,n n

r

v v

µ µ

(A.9a)

1 2 1

1(1 ) (1 )

2t n nv V V e e C

12 1

1

2

tn n

vV V C

e

12 1

11 1

2

tn n

ve V V e C

1 0,t

r

v

µ

(A.9b)

1 2 1

2 21(1 e) (1 e)

2

r r

n n

µ µV V C

r r

1

2 1

2 21

2

r r

n n

µ µV V C

e r r

1

2 1

2 2 2 21

2n n

e eV V C

r r

1

2 1

1 11

2n n

r

e eV V C

µ r r

(A.9c)

Page 148: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

130

where 2 1

1

2n nC V V is half of the relative normal component of the impact velocity. The

mean values of the rebound velocity components were calculated using the mean values of the

contact parameters; i.e. e=0.675, µ=1.2 and µr=0.05 based on their ranges given in Section 3.3.

The disks were assumed to approach each other with the same speed and in opposite directions

so that the second term in the normal velocity expression in Eq. (A.9a) vanished.

1( ) 0.675C,n avev 1 ,nvC

e

(A.10a)

1 2.01C ,t avev 1 1.2 ,t

ave

vC

e

1 1.68 ,t

ave

vC

(A.10b)

1

4.

10,

Cave r

1 2.45 ,

ave

C

e r

1 3.35 ,

ave

C

r

1 1.73

r ave

C

µ r

(A.10c)

The normalized sensitivities of the rebound velocities to variations in the contact parameters

were then calculated using the derivatives in Eq. (A.9) with the mean values of the contact

parameters, and then dividing by the average rebound velocities of Eq. (A.10).

The same procedure was used for case (2), the collision between a disk and a moving wall. The

average wall velocity was assumed to be one third of the disk approach velocity, based on the

experimental data from the laser probe [4]. The sensitivities of the normal, tangential and angular

velocities of the moving disk in case (2) are shown in Eq. (A.11).

1( ) 1.85C,n avev 1 2nvC

e

(A.11a)

1 4.36C ,t avev 1 2.6 ,t

ave

vC

e

1 3.35 ,t

ave

vC

1 6.7 ,t

r

vC

µ

(A.11b)

1

8.

38,

Cave r

1 5 ,

ave

C

e r

1 6.7 ,

ave

C

r

1 6.7

r ave

C

µ r

(A.11c)

Equations A.9, A.10 and A.11 were used in Section 3.3 to calculate the percentage uncertainties

of the rebound velocities due to 10% uncertainty in the contact parameters.

Page 149: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

131

References

[1] R.M. Brach, Friction, restitution and energy loss in planar impacts, J. Appl. Mech., 51

(1984) 164-170.

[2] J. Ai, J. F. Chen, J. M. Rotter, J. Y. Ooi, Assessment of rolling resistance models in discrete

element simulations, Powder Technol. 206 (2011) 269–282.

[3] T. N. Tang, Input parameters of discrete element methods, J. Eng. Mech., 132 (2006) 723-

729.

[4] K. Hashemnia, A. Mohajerani, J. K. Spelt, Development of a laser displacement probe to

measure particle impact velocities in vibrationally fluidized granular flows, Powder

Technol., 235 (2013) 940-952.

Page 150: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

132

Appendix B: MATLAB Codes of Chapter 2

Appendix B.1. Fast Fourier transformation (FFT) of the tub motion based on the measurements

Inputs: The tub wall displacement signal measured by the laser sensor

Outputs: Dominant frequencies and their corresponding amplitudes

Fs = 10000; % Sampling frequency T = 1/Fs; % Sample time L = 65000; % Length of signal t = (0:L-1)*T; % Time vector % y[]; z=laser(:,2); figure; %plot(Fs*t,z)

NFFT = 2^nextpow2(L); % Next power of 2 from length of y y = fft(z,NFFT)/L; f = Fs/2*linspace(0,1,NFFT/2+1); a=2*real(y); b=-2*imag(y); S=2*abs(y); F=f';

% plot(f,2*abs(y(1:NFFT/2+1))) % % title('Single-Sided Amplitude Spectrum of y-displacement at point A') % xlabel('Frequency (Hz)') % ylabel('|Y(f)|')

% [z,p,k] = butter(6,500/5000,'high'); % [sos,g] = zp2sos(z,p,k); % Convert to SOS form % Hd = dfilt.df2tsos(sos,g); % Create a dfilt object % h = fvtool(Hd); % Plot magnitude response % set(h,'Analysis','freq') % Display frequency response

[B,A]=BUTTER(3,500/5000,'low'); Z=filter(B,A,z);

plot(t,z); hold on plot(t,Z,'red');

title('y-acceleration at point B filtered and unfiltered') xlabel('Time(s)') ylabel('noise')

Y=fft(Z,NFFT)/L;

Page 151: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

133

figure plot(f,2*abs(y(1:NFFT/2+1))); hold on plot(f,2*abs(Y(1:NFFT/2+1)),'r');

title('Single-Sided Amplitude Spectrum of y-acceleration at point B') xlabel('Frequency (Hz)') ylabel('|Y(f)|')

Appendix B.2. Acceleration, velocity and position of the vibratory finisher center of gravity

Inputs: Acceleration of points A and B (Fig. 2.2a) measured by the accelerometers

Outputs: Acceleration, velocity and position of the vibratory finisher center of gravity

% Calculating angular velocity and angular acceleration of the tub with % respect to time and calculating the acceleration and vibration amplitude % of the tub center of gravity (including the media inside it) % angular velocity and acceleration are CCW (counter-clock-wise) by % default. x and y axis are positive Cartesian coordinates % tub vibration frequency: 47 Hz % t=[]; % time vector % AAp=[]; % acceleration of point A (in Volts), dimension n*2; first column

is % x-component and second column is y-component % ABp=[]; % acceleration of point B (in Volts), dimension n*2; first column

is % x-component and second column is y-component clc syms t % Fs=10000; % Sampling frequency % T=1/Fs; % Sample time % L=64000; % Length of signal % t=(0:L-1)*T; % Time vector % After fft analysis and knowing the coefficients

AAXF=-20.933*cos(2*pi*46.39*t)-

3.188*sin(2*pi*46.39*t)+10.522*cos(2*pi*139.3*t)+4.064*sin(2*pi*139.3*t)+10.2

24*cos(2*pi*325.2*t)-2.007*sin(2*pi*325.2*t)-0.167*cos(2*pi*371.7*t)-

14.052*sin(2*pi*371.7*t);

AAYF=-26.837*cos(2*pi*46.39*t)-61.254*sin(2*pi*46.39*t)-

8.391*cos(2*pi*139.3*t)-6.289*sin(2*pi*139.3*t)+7.328*cos(2*pi*185.9*t)-

16.361*sin(2*pi*185.9*t)-20.847*cos(2*pi*232.2*t)-6.868*sin(2*pi*232.2*t)-

19.459*cos(2*pi*325.2*t)-

3.416*sin(2*pi*325.2*t)+9.759*cos(2*pi*371.7*t)+8.150*sin(2*pi*371.7*t);

ABXF=-15.965*cos(2*pi*46.39*t)-

1.706*sin(2*pi*46.39*t)+2.880*cos(2*pi*418.1*t)+8.837*sin(2*pi*418.1*t);

Page 152: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

134

ABYF=59.748*cos(2*pi*46.39*t)-

14.752*sin(2*pi*46.39*t)+5.646*cos(2*pi*92.93*t)

+6.508*sin(2*pi*92.93*t)+5.596*cos(2*pi*325.2*t)+7.035*sin(2*pi*325.2*t)-

4.364*cos(2*pi*418.1*t)+6.870*sin(2*pi*418.1*t);

rAB=.39; % (distance between A and B) theta=atan(19.5/25.4); % x and y components of the tub center of gravity r=.32; % distance between point A and center of gravity % n=length(t); % T1=T; % t1=(0:n-1)*T1; VAXF=int(AAXF,t);% velocity (m/s) of point A-x VAYF=int(AAYF,t);% velocity (m/s) of point A-y VBXF=int(ABXF,t);% velocity (m/s) of point B-x VBYF=int(ABYF,t);% velocity (m/s) of point B-y

XAXF=int(VAXF,t);% position (m) of point A-x XAYF=int(VAYF,t);% position (m) of point A-y XBXF=int(VBXF,t)+rAB;% position (m) of point B-x XBYF=int(VBYF,t);% position (m) of point B-y

% XBX(1,1)=rAB; % initial conditions; the rest are zero.

Lan=atan(XBYF-XAYF)/(XBXF-XAXF); % angle between rAB and x-axis

omegaF=(VBYF-VAYF)/(rAB*cos(Lan)); % angular velocity of the tub

alphaF=(ABYF-AAYF+rAB*(omegaF)^2*sin(Lan))/(rAB*cos(Lan));%angular

acceleration of the tub

% ACG=zeros(n,2); % VCG=zeros(n,2); % XCG=zeros(n,2);

XCGXF=XAXF+r*sin(theta+Lan); XCGYF=XAYF-r*cos(theta+Lan); VCGXF=VAXF+r*omegaF*cos(theta+Lan); VCGYF=VAYF+r*omegaF*sin(theta+Lan); ACGXF=AAXF-r*(omegaF)^2*sin(theta+Lan)+r*alphaF*cos(theta+Lan); ACGYF=AAYF+r*(omegaF)^2*cos(theta+Lan)+r*alphaF*sin(theta+Lan); % AACG=sqrt(ACGX^2+ACGY^2);

% They should be evaluated with respect to t first. Fs=1000; % Sampling frequency T=1/Fs; % Sample time n=1000; % Length of signal t=0; for i=1:n t=t+T; omega(i,1)=eval(omegaF); alpha(i,1)=eval(alphaF); ACGX(i,1)=eval(ACGXF); ACGY(i,1)=eval(ACGYF); VCGX(i,1)=eval(VCGXF); VCGY(i,1)=eval(VCGYF); XCGX(i,1)=eval(XCGXF);

Page 153: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

135

XCGY(i,1)=eval(XCGYF); end figure; plot(t,omega); title('tub angular velocity vs. time') xlabel('Time(s)') ylabel('Omega(rad/s)') figure; plot(t,alpha); title('tub angular acceleration vs. time') xlabel('Time(s)') ylabel('Alpha(rad/s^2)') figure; plot(t,ACGX); title('x-acceleration of tub center of gravity vs. time') xlabel('Time(s)') ylabel('x-acceleration(m/s^2)') figure plot(t,ACGY); title('y-acceleration of tub center of gravity vs. time') xlabel('Time(s)') ylabel('y-acceleration(m/s^2)') figure plot(t,VCGX); title('x-velocity of tub center of gravity vs. time') xlabel('Time(s)') ylabel('x-velocity(m/s)') figure plot(t,VCGY); title('y-velocity of tub center of gravity vs. time') xlabel('Time(s)') ylabel('y-velocity(m/s)') figure plot(XCGX*1000,XCGY*1000); title('y-displacement of tub center of gravity vs. x-displacement') xlabel('x-displacement(mm)') ylabel('y-displacement(mm)') %fft analysis of omega, alpha, ACG(:,1), ACG(:,2), VCG(:,1), VCG(:,2) % Now, the amplitude of this acceleration (AACG) should be determined and % should be divided by square of 47 Hz=295 rad/s in order to obtain the

vibration amplitude

Page 154: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

136

Appendix B.3. Path of the top points of the vibratory finisher (A and B in Fig. 2.2a)

Inputs: Acceleration of points A and B (Fig. 2.2a) measured by the accelerometers

Outputs: Path of points A and B

clc Fs=10000; % Sampling frequency T=1/Fs; % Sample time L=65000; % Length of signal t=(0:L-1)*T; % Time vector AA1=AAp*981; % acceleration (m/s^2) of point A AB1=ABp*981; % acceleration (m/s^2) of point B AA(:,1)=AA1(:,1)-14.01;%12.56; AA(:,2)=AA1(:,2)-13.98;%11.56; AB(:,1)=AB1(:,1)+8.02;%6.85; AB(:,2)=AB1(:,2)-15.03;%13.3; rAB=.39; % (distance between A and B) theta=atan(19.5/25.4); % x and y components of the tub center of gravity r=.32; % distance between point A and center of gravity % A=AB-AA; % acceleration difference between A and B n=length(t); T1=T; t1=(0:n-1)*T1; VA=zeros(n,2);% velocity (m/s) of point A VB=zeros(n,2);% velocity (m/s) of point B % numerical integration for determining velocity of points A and B for i=2:n VA(i,1)=(AA(i,1)+AA(i-1,1))/2*T1+VA(i-1,1); VA(i,2)=(AA(i,2)+AA(i-1,2))/2*T1+VA(i-1,2); % VB(i,1)=(AB(i,1)+AB(i-1,1))/2*T1+VB(i-1,1); % VB(i,2)=(AB(i,2)+AB(i-1,2))/2*T1+VB(i-1,2); end XA=zeros(n,2); XB=zeros(n,2); XB(1,1)=rAB; % numerical integration for determining position of points A and B for i=2:n XA(i,1)=(VA(i,1)+VA(i-1,1))/2*T1+XA(i-1,1); XA(i,2)=(VA(i,2)+VA(i-1,2))/2*T1+XA(i-1,2); % XB(i,1)=(VB(i,1)+VB(i-1,1))/2*T1+XB(i-1,1); % XB(i,2)=(VB(i,2)+VB(i-1,2))/2*T1+XB(i-1,2); end plot(t1,VA(:,1)); title('x-velocity of tub center of gravity vs. time') xlabel('Time(s)') ylabel('x-velocity(m/s)') figure plot(t1,VA(:,2)); title('y-velocity of tub center of gravity vs. time') xlabel('Time(s)') ylabel('y-velocity(m/s)') figure plot(XA(:,1)*1000,XA(:,2)*1000);

Page 155: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

137

title('y-displacement of tub center of gravity vs. x-displacement') xlabel('x-displacement(mm)') ylabel('y-displacement(mm)')

Appendix B.4. The impact velocity calculation based on the laser displacement sensor measurements

Inputs: Displacement signals of the impact velocity probe window and the balls passing behind

the laser beam acquired by the laser sensors

Outputs: Impact velocity of the balls colliding the probe window

clc % y=[]; % ball motion after smoothing % z1=[]; % probe motion after smoothing t=.0005; %input('time increment'); n=65536; p=0; Z=0; % criteria should be changed for upstream 2 (very important) z=-1.446*z1+0.012; % glass target motion after smoothing emax=0.1; %0.12*GAmp % GAmp=0.86 mm is the amplitude of center of gravity (is

to be determined) taramp=0.12/2;% target motion amplitude(mm) depends on each signal pvdata=20; % pvdata is the number of data points from valley to peak (should

be read visually from the signal) min=0.2*pvdata; %should be specified max=0.6*pvdata; %should be specified for i=4:n-4 a1=(z(i-3,1)+z(i-2,1)+z(i-1,1))/3; a2=(z(i+3,1)+z(i+2,1)+z(i+1,1))/3; if z(i,1)>a1 && z(i,1)>a2 p=p+1; Z=Z+z(i,1); end end Zave=Z/p; % average of maximum z (target motion) peak values ll=-0.7; %should be specified visually hl=0.4; %should be specified visually j=0; a=0; c=0; v0=2*pi*47*taramp/2; % target reference velocity (1/3 (pi/6) of the data

points from the bottom points (valleys)) k=0; aa=0; bb=0; for i=30:n-4 a1=(y(i-3,1)+y(i-2,1)+y(i-1,1))/3; % can be changed to average of 2 a2=(y(i+3,1)+y(i+2,1)+y(i+1,1))/3; % can be changed to average of 2

Page 156: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

138

if y(i,1)>a1 && y(i,1)>a2 % maximum points recognition e1=abs(y(i,1)-z(i,1)); if e1<emax % impact point recognition. e is in mm. difference between

peaks of ball and target signals; can be changed for m=i-(pvdata+5):i a11=(y(m-3,1)+y(m-2,1)+y(m-1,1))/3; a22=(y(m+3,1)+y(m+2,1)+y(m+1,1))/3; if y(m,1)<a11 && y(m,1)<a22 % minimum points recognition e2=z(m,1)-y(m,1); if e2<emax/2 % in order to avoid same motions of balls

and target break; else bb=1; end else aa=1; end if aa==1 || bb==1 if y(m,1)>ll && y(m,1)<hl if y(m+1,1)>y(m,1) j=j+1; if j>min && j<max if abs(y(m+1,1)-y(m,1))<0.2 % 0.2 maximum

acceptable distance between the data points in order to avoid the erroneous

data (jumps) a=a+y(m+1,1)-y(m,1); else break; end end else if j>max j=max; end c=j-min; if c>0.25*pvdata %should be specified v=a/(c*t); if v~=0 && v>v0 % k=k+1 % disp(c); V(i)=v; % disp((m-1)*t); end end end end end aa=0; bb=0; end a=0; j=0; end end end j=1; for i=1:length(V)

Page 157: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

139

if V(i)~=0 Vf(j)=V(i); j=j+1; end end for i=1:65536 T(i,1)=(i-1)*.0005; end Vff=Vf'; S=Unique(Vff); plot(T,y); hold; plot(T,z,'r'); % figure; % hist(S,20);

Appendix C: MATLAB Codes of Chapter 3

Appendix C.1. The calculation of the in-plane components of the impact velocity based on DEM simulation results

Inputs: In-plane components of the particles instantaneous velocities acquired over 10 s of a

DEM simulation

Outputs: In-plane components of the mean impact velocity of the particles

% vnx=Input('particles x-relative normal '); % vnz=Input('particles z-relative normal '); % vx=Input('particles x-velocity '); % vz=Input('particles z-velocity ');

%V=zeros(n,4); V=[vnx vnz vx vz]

[n,m]=size(V); % V is the velocity matrix; 1st and 2nd columns: relative

normal x-velocity and z-velocity, 3rd and 4th columns: element A x-velocity

and z-velocity.

d=0.0063; % diameter

theta=zeros(n,1); vn=zeros(n,1); % absolute normal collision velocity (element A) vt=zeros(n,1); % absolute tangential collision velocity (element A)

for i=1:n a=((V(i,1))^2+(V(i,2))^2)^0.5; if a>0.001

Page 158: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

140

theta(i,1)=atan2(V(i,2),V(i,1)); vn(i,1)=V(i,3)*cos(theta(i,1))+V(i,4)*sin(theta(i,1)); vt(i,1)=-V(i,3)*sin(theta(i,1))+V(i,4)*cos(theta(i,1)); else V(i,3)=0; V(i,4)=0; end end

vn1=unique(vn); vt1=unique(vt);

vn2=mean(abs(vn1)); vt2=mean(abs(vt1));

vn2s=std(abs(vn1)); vt2s=std(abs(vt1));

k=0; for i=1:n if V(i,3)~=0 || V(i,4)~=0 k=k+1; Vx(k,1)=V(i,3); Vz(k,1)=V(i,4); end end

vx2=mean(abs(Vx)); vz2=mean(abs(Vz)); % vxBF=mean(Vx); % vzBF=mean(Vz); vx2s=std(abs(Vx)); vz2s=std(abs(Vz)); % vnx=Input('particles x-relative normal ');

Appendix C.2. The calculation of the out- of-plane component of the impact velocity based on DEM simulation results

Inputs: Out-of-plane component of the particles instantaneous velocities acquired over 10 s of a

DEM simulation

Outputs: Out-of-plane component of the mean impact velocity of the particles

%V=zeros(n,3); V=[vny vyA vyB]

[n,m]=size(V); % V is the velocity matrix; 1st column: relative normal x-

velocity, 2nd and 3rd columns: element A and element B y-velocity.

d=0.0063; % diameter

Page 159: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

141

for i=1:n a=V(i,1); if abs(a)<0.001 V(i,2)=0; %V(i,3)=0; end end

k=0; for i=1:n if V(i,2)~=0 %|| V(i,3)~=0 k=k+1; VyA(k,1)=V(i,2); %VyB(k,1)=V(i,3); end end

vyA2=mean(abs(VyA)); %vyB2=mean(abs(VyB)); vyA2s=std(abs(VyA)); %vyB2s=std(abs(VyB));

Appendix C.3. The calculation of coefficient of friction between spherical steel balls based on the linear tribometer measurements

Inputs: The horizontal force applied to a spherical steel ball by a similar ball and its vertical

displacements

Outputs: The tangential force applied to the spherical steel ball

% %T=input('tangential force vector:'); % %y=input('LVDT depth measurement:'); % %t=input('time:'); % n=length(T); % y0=min(y)*ones(n,1); % y1=(y-y0)/1000; % relative depth in mm % R=3.15; % radius in millimeters % for i=1:n % theta(i,1)=acos(1-y1(i,1)/R); % end % W=1; % for i=1:n % f(i,1)=T(i,1)*cos(theta(i,1))+W*sin(theta(i,1));

Page 160: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

142

% N(i,1)=-T(i,1)*sin(theta(i,1))+W*cos(theta(i,1)); % mu(i,1)=f(i,1)/N(i,1); % end % % t(1,1)=1; % % for i=1:n-1 % % t(i+1,1)=t(i,1)+1; % % end % plot(t,mu); % hold % plot(t,T,'r');

% predicting the tangential force measured by tribometer for ball-ball COF % measurements assuming a constant COF; theoretical approach mu=0.6; % constant COF W=1; % weight (N) s=0.9; d=6.3; % stroke and diameter in mm % theta0=asin(s/d); v=0.167; % velocity mm/s y1=0.026; y2=0.033; R=d/2; % depth magnitudes at both ends of the stroke theta1=acos(1-y1/R); theta2=acos(1-y2/R); n=100; for i=1:2*n+1 % theta(i,1)=-theta0+(i-1)*theta0/n; theta(i,1)=-theta1+(i-1)*(theta1+theta2)/(2*n); T(i,1)=(mu*cos(theta(i,1))-

sin(theta(i,1)))/(cos(theta(i,1))+mu*sin(theta(i,1)))*W; % tangential force

(N) t(i,1)=3.15/v*sin(theta(i,1)+theta1);

%t(i,1)=0.3636*(theta(i,1)+theta1)*180/pi; % time (s) end % t=theta*v; % conversion to time plot(theta*180/pi,T); figure plot(t,T);

Page 161: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

143

Appendix D: MATLAB Codes of Chapter 4

Appendix D.1. The calculation of local stress tensor, pressure and the coefficient of internal friction based on the DEM data to calibrate the equivalent properties needed in the continuum model

Inputs: Components of the normal and tangential forces applied to the steel particles in a

measurement location in a DEM simulation and the instantaneous velocities of particles

Outputs: The local stress tensor, local pressure and coefficient of internal friction in the

measurement location

% Multi layer (3D); No angle can be defined (unit vector instead)

[n,m]=size(V); % V is the velocity matrix (6 columns); 1st and 2nd and 3rd

columns: relative normal x-velocity, y-velocity and z-velocity, 4th, 5th and

6th columns: element A x-velocity, y-velocity and z-velocity.

d=0.0063; % diameter

% F is the average collision force matrix (6 columns) (over a single

collision duration); first and second, third columns: x, y and z components

of the normal force, forth, fifth and sixth columns: x, y and z components of

the tangential force.

vn=zeros(n,1); % absolute normal collision velocity (element A) ??? vt=zeros(n,1); % absolute tangential collision velocity (element A) ??? f=zeros(n,1); fn=zeros(n,1);

for i=1:n a=((V(i,1))^2+(V(i,2))^2+(V(i,3))^2)^0.5; if a>0.001 n1(i,1)=V(i,1)/a; n2(i,1)=V(i,2)/a; n3(i,1)=V(i,3)/a; vn(i,1)=V(i,4)*n1(i,1)+V(i,5)*n2(i,1)+V(i,6)*n3(i,1);

f(i,1)=sqrt((F(i,1))^2+(F(i,2))^2+(F(i,3))^2+(F(i,4))^2+(F(i,5))^2+(F(i,6))^2

); fn(i,1)=sqrt((F(i,1))^2+(F(i,2))^2+(F(i,3))^2); else V(i,4)=0; V(i,5)=0; V(i,6)=0; F(i,1)=0; F(i,2)=0; F(i,3)=0; F(i,4)=0; F(i,5)=0; F(i,6)=0; end

Page 162: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

144

end

% vn1=unique(vn); % %vt1=unique(vt); % % vn2=mean(abs(vn1)); % %vt2=mean(abs(vt1)); % % vn2s=std(abs(vn1)); % %vt2s=std(abs(vt1)); % % f1=unique(f); % f2=mean(f1); % fn1=unique(fn); % fn2=mean(fn1);

% added to the previous file

% for i=1:n % a=((V(i,1))^2+(V(i,2))^2+(V(i,3))^2)^0.5; % should the above filtered

velocities be used instead of these velocities ????????????? % n1(i,1)=V(i,1)/a; n2(i,1)=V(i,2)/a; n3(i,1)=V(i,3)/a; % end

k=0; Vol=0.02^2*0.009; % 0.009 m Volume of the measurement bin Tc=zeros(3,3); % collisional stress tensor TS=953; %2*47*10; % number of time steps (snap shots)

for i=1:n if V(i,4)~=0 || V(i,5)~=0 || V(i,6)~=0 k=k+1; Vx(k,1)=V(i,4); Vy(k,1)=V(i,5); Vz(k,1)=V(i,6); end end

vx2=mean(abs(Vx)); vy2=mean(abs(Vy)); vz2=mean(abs(Vz));

vx2s=std(abs(Vx)); vy2s=std(abs(Vy)); vz2s=std(abs(Vz));

% j=0; % for i=1:n % if F(i,1)~=0 || F(i,2)~=0 || F(i,3)~=0 || F(i,4)~=0 || F(i,5)~=0 ||

F(i,6)~=0 % j=j+1; % Fnx(j,1)=F(i,1); % Fny(j,1)=F(i,2); % Fnz(j,1)=F(i,3); % Ftx(j,1)=F(i,4);

Page 163: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

145

% Fty(j,1)=F(i,5); % Ftz(j,1)=F(i,6); % end % end

% considering all of the contact forces for i=1:n Fx(i,1)=F(i,1)+F(i,4); Fy(i,1)=F(i,2)+F(i,5); Fz(i,1)=F(i,3)+F(i,6); dX(i,1)=d/2*n1(i,1); dY(i,1)=d/2*n2(i,1); dZ(i,1)=d/2*n3(i,1);

Tc(1,1)=Tc(1,1)+Fx(i,1)*dX(i,1)/(Vol*TS); Tc(1,2)=Tc(1,2)+Fx(i,1)*dY(i,1)/(Vol*TS); Tc(1,3)=Tc(1,3)+Fx(i,1)*dZ(i,1)/(Vol*TS);

Tc(2,1)=Tc(2,1)+Fy(i,1)*dX(i,1)/(Vol*TS); Tc(2,2)=Tc(2,2)+Fy(i,1)*dY(i,1)/(Vol*TS); Tc(2,3)=Tc(2,3)+Fy(i,1)*dZ(i,1)/(Vol*TS);

Tc(3,1)=Tc(3,1)+Fz(i,1)*dX(i,1)/(Vol*TS); Tc(3,2)=Tc(3,2)+Fz(i,1)*dY(i,1)/(Vol*TS); Tc(3,3)=Tc(3,3)+Fz(i,1)*dZ(i,1)/(Vol*TS); end

Pc=-trace(Tc)/3; Tcs=(Tc+Tc')/2; % symmetric part of the stress tensor OK E=eig(Tcs); mu1=abs((max(E)-min(E))/(max(E)+min(E))); % effective coefficient of friction Tc0s=Tcs+Pc*eye(3); %T0=sqrt(((Tc0(1,1))^2+(Tc0(1,2))^2+(Tc0(1,3))^2+(Tc0(2,1))^2+(Tc0(2,2))^2+(T

c0(2,3))^2+(Tc0(3,1))^2+(Tc0(3,2))^2+(Tc0(3,3))^2)); T0s=norm(Tc0s,'fro'); % OK mu2=T0s/(Pc*sqrt(2));

Appendix D.2. The calculation of local shear rate based on the DEM data to calibrate the equivalent properties needed in the continuum model

Inputs: Components of the instantaneous velocities of particles in a measurement location and its

neighbor bins as described in Chapter 4 in a DEM simulation

Outputs: The local shear rate in the measurement location

% velocity gradient tensor and the inertial number calculation for multilayer

(3d) simulations % L components are calculated from the change of velocity from a bin to the % adjacent bins divided by the distance between the bins centers.

Page 164: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

146

% L12, L21, L23, L32 are close to zero (=zero assumption) and the rest 5

components are non zero. Average velocity in 6 bins or 4 bins (second

version) % and 3 directions should be calculated

% The average values should be calculated in each time step for all the % directions (3D)

% L11=(Vxright-Vxleft)/dx; L13=(Vxtop-Vxbottom)/dz; % L33=(Vztop-Vzbottom)/dz; L22=(Vybackward-Vyforward)/dy; % L31=(Vzright-Vzleft)/dx; L12=(Vxbackward-Vxforward)/dy; % L32=(Vzbackward-Vzforward)/dy; L21=(Vyright-Vyleft)/dx; % L23=(Vytop-Vybottom)/dz; dx=dz=8*d=50 dy=18 in mm % last components should be much smaller than others

% think to include the bin H Vx, Vy, Vz to compare the results of the

velocity % gradient for smaller distances i.e. 25 an 9 mm, respectively. average the % velocities per time step? % OR (second version) % L11=(Vxright-Vx)/dx; L13=(Vxtop-Vx)/dz; % L33=(Vztop-Vz)/dz; L22=(Vybackward-Vy)/dy; % L31=(Vzright-Vz)/dx; L12=(Vxbackward-Vx)/dy; % L32=(Vzbackward-Vz)/dy; L21=(Vyright-Vy)/dx; % L23=(Vytop-Vy)/dz; dx=dz=4*d=25 dy=9 in mm % Vx, Vy and Vz are velocity % components in H (previous)

% for each point e.g. H [n,m]=size(V); % has 18 columns; 1-3 left, 4-6 right, 7-9 bottom, 10-12 top,

13-15 forward, 16-18 backward (x,y,z) velocities, n: number of time steps 18

DAQ % OR % 1-3 H, 4-6 right, 7-9 H, 10-12 top, 13-15 H, 16-18 backward (x,y,z) % velocities 12 DAQ % if the first and the second versions give the same results, the second is % preferable as it is less time consuming %dx=.05; dz=0.05; dy=0.018; % in meters dx=.02; dz=0.02; dy=0.009; % in meters L11=zeros(n,1);L12=zeros(n,1);L13=zeros(n,1);L21=zeros(n,1);L22=zeros(n,1);L2

3=zeros(n,1);L31=zeros(n,1);L32=zeros(n,1);L33=zeros(n,1); % velocity

gradient tensor components Lm=zeros(3,3); % average stress gradient tensor % L=zeros(3*n,n); D=zeros(3*n,n); Ddev=zeros(3*n,n); gamma=zeros(n,1); for i=1:n L11(i,1)=(V(i,4)-V(i,1))/dx; % velocity gradient tensor components at

each time step L12(i,1)=(V(i,16)-V(i,13))/dy; L13(i,1)=(V(i,10)-V(i,7))/dz; L21(i,1)=(V(i,5)-V(i,2))/dx; L22(i,1)=(V(i,17)-V(i,14))/dy; L23(i,1)=(V(i,11)-V(i,8))/dz; L31(i,1)=(V(i,6)-V(i,3))/dx; L32(i,1)=(V(i,18)-V(i,15))/dy; L33(i,1)=(V(i,12)-V(i,9))/dz; Lm(1,1)=Lm(1,1)+L11(i,1)/n; Lm(1,2)=Lm(1,2)+L12(i,1)/n;

Page 165: Vibrationally-Fluidized Granular Flows: Impact and Bulk Velocity … · 2015. 6. 17. · using discrete element modelling (DEM) and compared to the measured values for spherical steel

147

Lm(1,3)=Lm(1,3)+L13(i,1)/n; Lm(2,1)=Lm(2,1)+L21(i,1)/n; Lm(2,2)=Lm(2,2)+L22(i,1)/n; Lm(2,3)=Lm(2,3)+L23(i,1)/n; Lm(3,1)=Lm(3,1)+L31(i,1)/n; Lm(3,2)=Lm(3,2)+L32(i,1)/n; Lm(3,3)=Lm(3,3)+L33(i,1)/n; L=[L11(i,1), L12(i,1), L13(i,1); L21(i,1), L22(i,1), L23(i,1); L31(i,1),

L32(i,1), L33(i,1)]; % velocity gradient tensor at each time step D=(L+L')/2; % symmetric velocity gradient tensor components Ddev=D-trace(D)/3*eye(3); % Deviatoric rate of strain tensor at each time

step

%gamma=sqrt(((Ddev(1,1))^2+2*(Ddev(1,2))^2+(Ddev(2,2))^2+2*(Ddev(1,3))^2+2*(D

dev(2,3))^2+(Ddev(3,3))^2)); gamma(i,1)=sqrt(2)*norm(Ddev,'fro'); % shear rate at each time step %

or *sqrt(2) tr(D)=0 Kamrin's thesis %D0=det(dev(L)) end gammaave=mean(gamma); % average shear rate gamma_std=std(gamma);

Dm=(Lm+Lm')/2; % symmetric velocity gradient tensor components Ddevm=Dm-trace(Dm)/3*eye(3); % Deviatoric rate of strain tensor at each time

ste gammam=sqrt(2)*norm(Ddevm,'fro'); % shear rate at each time step % or

*sqrt(2) tr(D)=0 Kamrin's thesis DmT=trace(Dm); % is it equal to zero? Dilatancy=-DmT/(gammam/sqrt(2)); % Tk should be studied (calculated to br compared to Tc), the previous code

can still be used for Vtot and % GTp